You are on page 1of 466

Endocrinology

Second Edition
Medical Management of Thyroid Disease
Medical
about the book…
For general practitioners and endocrinologists, the new Second Edition of this
bestselling book offers the most up-to-date and practical guidance to diagnose
and manage common and uncommon thyroid diseases.
New to the Second Edition:
• information on thyroid neoplasia, leading to new effective treatments of
advanced thyroid cancer
• important new research on subclinical thyroid disease in the elderly and
Management
of Thyroid
thyroid disorders in pregnancy
• new research on thyroid physiology, pathophysiology, and therapeutics
The new edition is fully evidence-based and updated to include the most current

Disease
treatment and latest findings:
• the screening and case finding for thyroid disease
• the use of calcitonin in the diagnosis of medullary thyroid cancer
• the diagnosis and management of subclinical hyperthyroidism
(mild hyperthyroidism)

Second Edition
• thyroid disease related to interferon therapy and amiodarone therapy
about the editor...
DAVID S. COOPER is Professor of Medicine, The Johns Hopkins University School
of Medicine; Professor of International Health, Johns Hopkins Bloomberg School
of Public Health; and Physician and Director, the Thyroid Clinic, Johns Hopkins
Hospital, Baltimore, Maryland, USA. Dr. Cooper received his M.D. from Tufts
University, Boston, Massachusetts, USA. He is a member of the Endocrine Society
and is a past president of the American Thyroid Association. Dr. Cooper is currently
Deputy Editor of the Journal of Clinical Endocrinology and Metabolism, Editor–in-
Chief of Endocrinology, Up-to-Date, and Contributing Editor of the Journal of the
American Medical Association (JAMA). Dr. Cooper was also the editor of the first
edition of Informa Healthcare’s Medical Management of Thyroid Disease.
Printed in the United States of America

Cooper
H7064

Edited by
David S. Cooper
Medical
Management
of Thyroid
Disease
Medical
Management
of Thyroid
Disease
Second Edition

Edited by
David S. Cooper
The Johns Hopkins University School
of Medicine, Baltimore, Maryland, USA
Informa Healthcare USA, Inc.
52 Vanderbilt Avenue
New York, NY 10017

C 2008 by Informa Healthcare USA, Inc.
Informa Healthcare is an Informa business

No claim to original U.S. Government works


Printed in the United States of America on acid-free paper
10 9 8 7 6 5 4 3 2 1

International Standard Book Number-10: 1-4200-7064-9 (Hardcover)


International Standard Book Number-13: 978-1-4200-7064-4 (Hardcover)

This book contains information obtained from authentic and highly regarded sources. Reprinted
material is quoted with permission, and sources are indicated. A wide variety of references are listed.
Reasonable efforts have been made to publish reliable data and information, but the author and the
publisher cannot assume responsibility for the validity of all materials or for the consequence of their
use.

No part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic,
mechanical, or other means, now known or hereafter invented, including photocopying, microfilming,
and recording, or in any information storage or retrieval system, without written permission from the
publishers.

For permission to photocopy or use material electronically from this work, please access
www.copyright.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc.
(CCC) 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization
that provides licenses and registration for a variety of users. For organizations that have been granted
a photocopy license by the CCC, a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation without intent to infringe.

Library of Congress Cataloging-in-Publication Data


Medical management of thyroid disease / edited by David S. Cooper. – 2nd ed.
p. ; cm.
Includes bibliographical references and index.
ISBN-13: 978-1-4200-7064-4 (hardcover : alk. paper)
ISBN-10: 1-4200-7064-9 (hardcover : alk. paper)
1. Thyroid gland–Diseases. I. Cooper, David S.
[DNLM: 1. Thyroid Diseases–therapy. 2. Thyroid Diseases–diagnosis.
WK 267 M489 2009]

RC655.M434 2009
616.4’4–dc22
2008042809

For Corporate Sales and Reprint Permission call 212-520-2700 or write to: Sales Department,
52 Vanderbilt Avenue, 7th floor, New York, Ny 1017.

Visit the Informa Web site at


www.informa.com

and the Informa Healthcare Web site at


www.informahealthcare.com
Preface to the First Edition

I have been privileged to be a clinical investigator and thyroidologist for almost


25 years. When I was contacted by an editor of Marcel Dekker, Inc., about the
possibility of editing a book on diseases of the thyroid, my first reaction was,
“Why do we need another thyroid book?” There are already several standard
works that exhaustively cover thyroid physiology, pathophysiology, and clinical
thyroid disease. One of these texts is considered to be the “Bible” of thyroidology,
and there are several others that are also notable for their comprehensive treatment
of the subject. Many of them are on my bookshelf, with broken spines and dog-
eared pages from frequent use. However, after further reflection, it occurred to me
that there was a niche for a smaller book that focused purely on the management of
patients with thyroid disease, using a very practical and evidence-based approach.
Therefore, I agreed to develop and edit a book that would cover the basics and
that would appeal to practitioners.
In the last decade, there have been major advances in our understanding of
thyroid gland regulation, thyroid hormone synthesis and action, and the genet-
ics of thyroid neoplasia. While an understanding of molecular mechanisms and
“translational research” is important for optimal patient care, clinicians also need
up-to-date, practical information to help them in the management of common (and
not so common) thyroid problems. In this book, every effort has been made to
cover the universe of thyroidology in as much detail as possible, and yet provide
information that is easily digested and useful. Each chapter considers the epi-
demiology, diagnosis, and treatment of a specific thyroid disorder or disorders,
and serves as a “mini-review” of a particular topic. Because any discussion that
involves the care of patients will engender controversy, areas of debate are identi-
fied and the contributors’ own preferences are clearly stated. Every effort has been
made to provide alternative points of view and to base recommendations on the
best available evidence.
I was extremely fortunate to attract contributors who have international rep-
utations as authors and thyroid researchers, and who are, most importantly, con-
summate physicians. I thank all of them for their diligence, intellectual approach,
and clinical acumen, without which this volume would not have been possible.
And, although this book was written for clinicians, it is my hope that the millions
of patients with thyroid disease will be the primary beneficiaries.

David S. Cooper

iii
Preface

In deciding whether a second edition of Medical Management of Thyroid Disease


would be something to pursue, the most obvious question is whether there have
been enough new developments in the field to warrant it. To put it succinctly:
Would a revised text be sufficiently different from the first edition? The first
edition, published in 2001, clearly relied on information that is now almost a
decade old. Given the exponential growth in medical knowledge, it would only be
logical to assume that there has been a great deal learned in the last decade about
thyroid physiology, pathophysiology, and therapeutics. Some of this knowledge is
entirely new, leading to novel insights and to innovative therapies for previously
untreatable conditions. Some of the knowledge is not necessarily entirely new,
but has led us to challenge previous firmly entrenched concepts, and to reconsider
methods of diagnosis and treatment that had been traditionally taught, but which,
often in retrospect, were based more on expert opinion than on the best sciencific
evidence.
Several examples come to mind.
1. In the realm of basic science, new knowledge about signaling pathways in
thyroid neoplasia is starting to yield startling dividends, especially in the
management of advanced thyroid cancer. Drugs that target specific vascular
endothelial growth factor (VEGF) receptor signaling, epidermal growth factor
(EGF) receptor signaling, and specific oncogenes such as RET and BRAF are
on the horizon, and appear to be quite promising in a group of patients where
up until recently very little could be offered therapeutically.
2. Recent epidemiologic studies have also yielded important information that
impacts directly on patient care. To cite just one example, data from the
Cardiovascular Health Study have shown that older adults with even mild
degrees of thyroid overactivity (subclinical hyperthyroidism) are at risk for
atrial fibrillation. This suggests that intervention should be done earlier than
we previously had thought necessary. This observation could have profound
implications as the number of elderly citizens increases dramatically.
3. In the realm of clinical research, it has long been known that pregnant women
with circulating antithyroid peroxidase (anti-TPO) antibodies are at higher
risk for miscarriage. The reason for this association has been a source of great
interest, but has remained obscure until recently. A recent prospective clinical
trial has shown that women with circulating anti-TPO antibodies and normal
thyroid function likely develop very mild thyroid hormone deficiency during

v
vi Preface

pregnancy. Furthermore, treatment with thyroid hormone reverses any risk of


adverse pregnancy outcomes. This information has major public health and
clinical implications, since 5% to 10% of all women have circulating anti-TPO
antibodies.
I could cite numerous other examples of how information that has been published
in the last few years has led to a change in how we manage a variety of common
thyroid conditions. Therefore, not only is a new edition of this text justified, but it
is long overdue!
The earlier edition was developed to be a practical book on how to man-
age both common and uncommon thyroid diseases. The information was meant to
appeal to practitioners, especially primary care physicians and clinical endocrinol-
ogists. Therefore, the focus was on the clinical presentation, laboratory diagnosis,
and treatment, leaving the basic aspects of thyroid physiology and pathophysiol-
ogy to other more traditional textbooks that take a more encyclopedic approach.
It is my desire that this second edition continue in that tradition. In addition, when
the first edition was published, the concept of “evidence-based medicine” was
not as well accepted as a basis for decision making as it is now. Also, although
the earlier edition was “evidence based” as much as possible, there were few
prospective clinical trials in the field of thyroidology. Over the last decade, there
had been a number of clinical trials and meta-analyses, allowing this book to
be more “evidence based” than before. However, as with all areas of medicine,
expert opinion still prevails in the absence of “good evidence,” and where expert
opinion substitutes for evidence in some circumstances, this will be noted, as will
differences of opinion, controversies, and alternative points of view.

David S. Cooper, M.D.


Contents

Preface to the First Edition . . . . iii


Preface . . . . v
Contributors . . . . ix

1. The Laboratory Approach to Thyroid Disorders 1


Steven I. Sherman

2. Hyperthyroidism Due to Graves’ Disease, Toxic Nodules and Toxic


Multinodular Goiter 39
Kenneth D. Burman and David S. Cooper

3. Thyroiditis and Other More Unusual


Forms of Hyperthyroidism 101
Shon E. Meek and Robert C. Smallridge

4. Hypothyroidism 145
Michael T. McDermott and E. Chester Ridgway

5. Thyroid Nodules and Multinodular Goiter 203


Hossein Gharib

6. Differentiated Thyroid Carcinoma 237


Jennifer A. Sipos and Ernest L. Mazzaferri

7. Medullary Thyroid Carcinoma, Anaplastic Thyroid Carcinoma, and


Thyroid Lymphoma 297
Jennifer A. Sipos and Ernest L. Mazzaferri

8. Pediatric Thyroid Disorders 331


Rosalind S. Brown

9. Thyroid Disease and Pregnancy 369


Susan J. Mandel

10. Thyroid Disease in the Elderly 401


Anne R. Cappola, Myron Miller and Steven R. Gambert

Index . . . . 439

vii
Contributors

Rosalind S. Brown Harvard Medical School, Boston, Massachusetts, U.S.A.


Kenneth D. Burman Washington Hospital Center, Georgetown University,
and the Uniformed Services of the Health Sciences, Washington, D.C., U.S.A.
Anne R. Cappola University of Pennsylvania School of Medicine,
Philadelphia, Pennsylvania, U.S.A.
David S. Cooper The Johns Hopkins University School of Medicine,
Baltimore, Maryland, U.S.A.
Steven R. Gambert Sinai Hospital of Baltimore and The Johns Hopkins
University School of Medicine, Baltimore, Maryland, U.S.A.
Hossein Gharib College of Medicine, Mayo Clinic, Rochester, Minnesota,
U.S.A.
Susan J. Mandel University of Pennsylvania School of Medicine,
Philadelphia, Pennsylvania, U.S.A.
Ernest L. Mazzaferri University of Florida, Gainesville, Florida, U.S.A.
Michael T. McDermott University of Colorado Denver, School of Medicine,
Aurora, Colorado, U.S.A.
Shon E. Meek College of Medicine, Mayo Clinic, Jacksonville, Florida,
U.S.A.
Myron Miller Sinai Hospital of Baltimore and The Johns Hopkins University
School of Medicine, Baltimore, Maryland, U.S.A.
E. Chester Ridgway University of Colorado Denver, School of Medicine,
Aurora, Colorado, U.S.A.
Steven I. Sherman The University of Texas, M. D. Anderson Cancer Center,
Houston, Texas, U.S.A.
Jennifer A. Sipos The Ohio State University, Columbus, Ohio, U.S.A.
Robert C. Smallridge College of Medicine, Mayo Clinic, Jacksonville,
Florida, U.S.A.

ix
1
The Laboratory Approach
to Thyroid Disorders

Steven I. Sherman
The University of Texas, M. D. Anderson Cancer Center, Houston,
Texas, U.S.A.

INTRODUCTION
The central role of the thyroid gland in controlling metabolism was recognized
in the 19th century, but evaluation of the function of the thyroid remains an
evolving science. Initial approaches to the assessment of thyroid function cen-
tered on measuring end-organ responses as biological markers of thyroid hormone
actions. Development of in vitro competitive binding assay methods allowed the
direct quantification of hormone levels in serum, and sensitive immunoassays
have demonstrated the subtleties of pituitary and hypothalamic control of the
thyroid. Abnormalities of hormone binding by serum proteins necessitated sen-
sitive estimation of free hormone levels. With the detection of serum markers
of autoimmune and malignant diseases of the gland has come earlier diagnosis
and improved monitoring of these conditions, often with greater sensitivity than
may be clinically relevant. Limitations to the measurement methods used exist,
however, particularly when underlying assumptions about the comparability of
patient and control specimens are invalid. Nonetheless, the clinician can now
effectively confirm suspected diagnoses of thyroid dysfunction, cost-effectively
screen asymptomatic populations for common diseases, and appropriately monitor
the treatment of patients with disorders of the thyroid.

1
2 Sherman

PHYSIOLOGY OF THE HYPOTHALAMIC-PITUITARY-THYROID AXIS


Excellent reviews and books provide detailed exploration of the physiology of
the hypothalamic-pituitary-thyroid axis, and the reader is invited to delve into
those worthwhile sources. For the purposes of this chapter, a brief review of
the biosynthesis and transport of thyroid hormones and the regulation of thyroid
function by the neurohypophysis will suffice.
The synthesis of thyroxine (T4) and triiodothyronine (T3) begins with the
active transport of iodide into the cell via a sodium-iodine symporter located
in the basal membrane. Following oxidation by thyroid peroxidase (TPO), the
iodide moiety is covalently attached to tyrosyl residues of thyroglobulin, and the
resulting iodotyrosines are coupled and cleaved from thyroglobulin to form T4
and T3, normally in a 10:1 ratio. Thyroid hormone secretion requires endocytosis
and degradation of iodinated thyroglobulin, followed by release of T4 and T3
into the circulation. This process results in the total daily output of 80 to 100 ␮g
of T4. In contrast, only 20% of circulating T3 is produced by the thyroid, the
remaining 80% deriving from the enzymatic outer-ring or 5’-monodeiodination
of T4 in extrathyroidal tissues such as the liver, kidney, brain, muscle, and skin.
Removal of the inner-ring or 5-iodine of T4 forms the inactive metabolite reverse
T3 (rT3). Other inactivating pathways for T4 and T3 include glucuronidation,
sulfation, deamination, and cleavage. The normal daily fractional turnover rates
for T4 and T3 are 10% and 75%, respectively.
In serum, at least 99.95% of T4 and 99.5% of T3 molecules are bound by
the transport proteins thyroxine-binding globulin (TBG), transthyretin (thyroxine-
binding prealbumin), and albumin. Although TBG is present in lower concentra-
tion than either transthyretin or albumin, its greater affinity for thyroid hormones
makes it the predominant serum carrier of T4 and T3. Variations in binding
characteristics among normal and abnormal thyroid hormone-binding proteins
are responsible for much of the methodologic limitations in assays that attempt
to measure concentrations of free T4 and T3. This large pool of protein-bound
hormone provides a stable reservoir that maintains the supply of free, unbound
hormone available for transport into the cells. Once within target cells, T4 is fur-
ther deiodinated to T3, which in the nucleus binds to the thyroid hormone receptor,
modulating transcription of thyroid hormone-responsive genes and producing most
of the clinical effects recognized as the metabolic effects of thyroid hormones.
The primary regulatory influence on thyroid gland function is the circu-
lating level of thyrotropin (thyroid-stimulating hormone or TSH). Produced by
thyrotroph cells of the anterior pituitary, TSH is a two-subunit glycoprotein, the
specificity of which is conferred by its alpha subunit; the beta subunit is struc-
turally similar to that of follicle-stimulating hormone, luteinizing hormone, and
human chorionic gonadotropin. Negative feedback by T4 and T3 influences TSH
synthesis and release, as evidenced by the inverse log–linear relationship of the
concentrations of TSH and free iodothyronine (1,2). It is likely that each indi-
vidual has a genetically determined set point for this TSH/free T4 relationship,
based on twin studies (3,4). TSH levels peak just before nocturnal sleep, and
The Laboratory Approach to Thyroid Disorders 3

the nadir occurs in the late afternoon; this nocturnal surge is lost early in the
course of nonthyroidal illness. TSH levels in various populations conform best to
a log-Gaussian rather than Gaussian distribution (5). The hypothalamic tripeptide
thyrotropin-releasing hormone (TRH) stimulates TSH secretion and modulates
thyrotroph response to altered thyroid hormone levels. In conjunction with the
suppressive effects of dopamine, corticosteroids, somatostatin, androgens, and
endogenous opioids, TRH may be responsible for modulating the set point for
the negative feedback loop that controls thyroid hormone levels. Hypothalamic
production of TRH itself is regulated by circulating thyroid hormones as well as
by multiple central nervous system factors.

LABORATORY EVALUATION OF THYROID FUNCTION


Assays of Thyroid Hormones
Total Serum Iodothyronine Concentrations
When concentrations and binding affinities of thyroid hormone-binding proteins
are normal, there exists at physiologic equilibrium a direct relationship between
levels of total hormone and free hormone (6). Thus, measurement of total iodothy-
ronine concentration can provide a reasonable surrogate for estimating the amount
of free iodothyronine present. Either serum or plasma can be used to assay hor-
mone concentrations, although serum is generally preferred. The most commonly
employed technique for the determination of total T4 and T3 concentrations is
competitive immunoassay, using either polyclonal or monoclonal “capture” anti-
bodies directed against the specific iodothyronine. To ensure measurement of
bound as well as free hormone, inhibitors of iodothyronine binding are added—for
example, 8-anilino-1-naphthalene sulfonic or salicylic acids for TBG and barbital
for thyroxine-binding prealbumin (TBPA). These agents successfully dissociate
the hormone from binding proteins without interfering with hormone binding to
immunoglobulin.
Radioimmunoassay (RIA) depends upon measurement of the distribution
of a tracer quantity of radiolabeled hormone that competes with the endogenous
hormone in the patient’s specimen for binding to a capture antibody. The higher
the serum hormone concentration, the lower the amount of radiolabel that binds to
the antibody. Following the addition of a limited amount of capture antibody and
the radiolabeled iodothyronine to be measured, the antibody–antigen complexes
are separated from the serum. Separation techniques vary, including ammonium
sulfate or second antibody precipitation. Newer methods that facilitate automated
separation include attachment of the anti-T4 antibody to a solid phase, such as the
wall of the assay tube or magnetizable particles. The concentration of either T4
or T3 is then determined by comparison of the amount of antibody-bound radio-
label with a simultaneously derived standard curve. A fundamental assumption,
therefore, is that there is no difference in the assay conditions (including protein
binding and other constituents found in the serum) between the patient’s sample
and the control standards, an assumption that is often invalid.
4 Sherman

Nonisotopic methods avoid reliance upon radioactive reagents and are now
the most commonly used assays. The heterogeneous enzyme-linked immunosor-
bent assay (ELISA) incorporates colorimetric, fluorescent, or luminescent sub-
strates that create a quantitative signal when interacting with a specific enzyme
bound to the tracer hormone—for example, alkaline phosphatase, horseradish
peroxidase, or glucose-6-phosphate dehydrogenase. As in RIA, numerous phys-
ical and chemical approaches exist for separating signal bound to the anti-
iodothyronine antibody from unbound signal. In contrast, homogeneous enzyme
immunoassays do not require a separation step. Instead, the binding of antibody
to a tracer hormone directly affects the activity of the signal-generating enzyme
bound to the tracer. Other technologies have also been applied to shorten the time
requirements for assays and to facilitate automation.
Reference ranges vary to some degree, but commonly cited ranges are 4.5
to 12.6 ␮g/dL (58–160 nmol/L) for total T4 and 80 to 180 ng/dL (1.2–2.7 nmol/L)
for total T3 (7).
As developed by their manufacturers, these assay techniques have similar
performance characteristics, although each may be affected by different sources
of interference. In field use, however, a concerning level of inaccuracy exists even
in total thyroxine immunoassays; a survey of nearly 2000 clinical laboratories
employing 19 different automated thyroxine assays identified that approximately
30% of the laboratories failed to meet accuracy standards defined by the Clinical
Laboratory Improvement Amendments of 1988 (8). Testing an identical speci-
men of fresh frozen serum in 1528 laboratories, 13% of the 16 T4 measurement
methods yielded mean results that differed from the overall mean of all meth-
ods by at least 10% (8). For total T3, the degree of inaccuracy is worse; testing
an identical specimen of fresh frozen serum in 926 laboratories, 58% of the 12
T3 measurement methods yielded mean results that differed from the overall
mean of all methods by at least 10% (8). Contributing factors to measurement
error include qualitative differences between the protein constituents of sample
diluents used for calibration and those found in patient sera, leading to differen-
tial dissociation of hormone from binding proteins, and the lack of harmonized
standards (7).
Determination of Free T4 and T3 Concentrations
Because T4 and T3 are highly bound to serum proteins, alterations in either the
levels of these proteins or their binding characteristics can significantly alter the
concentration of total hormone. As it is the free hormone that is biologically
active, however, techniques are required to permit either direct measurement or
estimation of the serum free hormone levels. All methods that have been developed
face the identical problem: distinguishing between the 3 to 4 orders of magnitude
difference in the concentrations of the free and the protein-bound hormones. In all
free hormone assays, the central assumption is that the effectiveness of separating
the free from the bound hormone is identical in both the patient samples and the
standards used to calibrate the assay, an assumption that is difficult to validate in
The Laboratory Approach to Thyroid Disorders 5

all potential clinical situations. As a result, in a study of 1744 clinical laboratories,


using 13 different free T4 measurement methods to assay an identical sample of
fresh frozen serum, 38% of the methods yielded average concentrations of free T4
that were at least 10% different from the overall group mean and 8% differed by
⬎20% (8).
The reference standard for direct measurement of free T4 and T3 is the
equilibrium dialysis method, although even this methodology lacks a univer-
sal reference measurement procedure (9). Undiluted patient serum is dialyzed
overnight across a membrane with pores that allow free but not protein-bound
hormone to partition, allowing equilibration of the free hormone concentration
across the membrane. A highly sensitive RIA, capable of detecting nanogram (or
picomole) quantities of hormone, is then used to measure the hormone content of
the protein-free dialysate, comparing to a standard curve generated with gravimet-
rically determined amounts of hormone (10). Faster turnaround can be achieved
by using ultrafiltration rather than equilibrium dialysis, but greater variability can
result from minimal amounts of serum proteins that leak through the filtration
device as well as hormone that is adsorbed to either the membrane or container
surface. Such direct measurements are generally expensive, time-consuming, and
not widely used commercially except as calibrators for more commonly used
techniques.
By estimating the fraction of hormone that is free or unbound and multi-
plying by the total concentration of hormone, an indirect method will yield an
estimate of the free hormone level. In the indirect equilibrium dialysis method,
tracer quantities of radiolabeled hormone are added to the patient’s serum, and
the mixture equilibrates across a dialysis membrane. A second dialysis step incor-
porating an anion exchange resin then removes the radioactivity due to labeled
inorganic iodine and other contaminating compounds found in the tracer iodothy-
ronine preparation. Alternatively, the radiolabeled hormone can be precipitated by
addition of magnesium chloride. Regardless of the second separation step used,
the resultant fraction of radiolabel found in the dialysate is proportional to the
ratio of free to total hormone in the patient’s serum. Expected free hormone frac-
tions are approximately 0.02% to 0.04% for T4 and 0.2% to 0.4% for T3. The
free T4 or T3 estimate can then be calculated by multiplying the free fraction
by the total serum hormone concentration obtained using a conventional RIA.
Expected adult values are approximately 0.8 to 2.3 ng/dL for free T4 and 210 to
440 pg/dL for free T3. Although these indirect dialysis methods generally correct
for most protein-binding abnormalities, considerable variability in results still per-
sist, particularly from impurities in the tracer preparation and occasional failure to
achieve equilibrium. The methods are exacting, expensive, and time-consuming,
and other approaches to estimating free hormone concentrations have been devel-
oped for widespread use. However, a recently developed equilibrium dialysis tech-
nique that utilizes liquid chromatography-tandem mass spectrometry to measure
free T4 levels may permit rapid and cost-effective assessment of multiple blood
samples (11).
6 Sherman

Figure 1 Methods for radioimmunoassay estimation of free thyroxine concentration.

Immunoassay methods for estimation of free hormone concentration are now


widely used (Fig. 1). In the “analogue” or “one-step” free T4 method, a labeled T4
analogue that does not bind to serum-binding proteins is added to serum and the
mixture is either incubated with an anti-T4 antibody or allowed to bind to antibody
attached to a solid phase. At equilibrium, the amount of analogue complexed to the
antibody is inversely proportional to the amount of free T4 that is available. One-
step methods require structurally modified analogues that do not displace hormone
from protein-binding sites, but a complete lack of displacement is rarely achieved.
The Laboratory Approach to Thyroid Disorders 7

Therefore, these methods depend on the assumption that there is no difference in


hormone binding affinity for proteins between the sample to be measured and the
assay controls or calibrators, both for the actual analyte as well as the analogue.
This assumption is particularly at risk when there are circulating inhibitors of
hormone binding in serum, such as renal failure or other nonthyroidal illnesses,
or major alterations in hormone-binding protein concentrations (12). Because the
analogues used generally bind to albumin, although not with the same kinetics as
T4 or T3, this method may not be correct for abnormalities in albumin binding.
In “two-step” assays, serum is exposed to a solid phase containing anti-T4
antibody, binding a certain amount of free hormone to the solid phase. By diluting
the specimen and limiting the duration of incubation, there should be minimal
disruption of endogenous hormone binding to serum proteins (10). After removal
of the serum and its proteins, a tracer quantity of radiolabeled T4 is incubated with
the solid phase, equilibrating with the remaining unoccupied antibody molecules.
The amount of radiolabeled T4 complexed to the solid phase is thus inversely
proportional to the free T4 concentration of the serum. Because the label is unable
to interact with serum-binding proteins or endogenous inhibitors of hormone bind-
ing to protein (due to the physical separation step), the “two-step” method has a
good correlation with the free T4 determined by direct equilibrium dialysis. Non-
radioactive assays have also been developed, and automated two-step procedures
are in common use.
For free T3 measurements, methods that rely upon physical separation of
bound from free hormone, such as dialysis or ultrafiltration, are not generally
commercially available. The same technology for “one-step” assays of free T4 is
used to measure free T3. Interference from serum proteins and difficulty avoiding
stripping T3 from its binding proteins is a greater problem than in free T4 assays
(13). In a study of 404 clinical laboratories, using 11 different free T3 measurement
methods to assay an identical sample of fresh frozen serum, 55% of the methods
yielded average concentrations of free T3 that were at least 10% different from
the overall group mean and 27% differed by ⬎20% (8). New methods that use
tandem mass spectrometry following equilibrium dialysis or ultrafiltration may
allow faster and more reliable assays (14).
The thyroid hormone binding ratio (THBR), another calculated value pro-
portional to the fraction of hormone that is free in circulation, derives from mea-
surement of the availability of protein-binding sites in the patient’s serum. In
the traditional uptake method, a tracer quantity of radiolabeled iodothyronine is
added to the serum and allowed to partition between unoccupied specific protein
binding sites and a nonsaturable adsorbent—for example, talc, charcoal, resin, or
anti-iodothyronine antibodies. T3 is generally preferred as the labeled ligand, as it
has a lower affinity for TBG and therefore does not displace T4 its binding sites.
There is an inverse relationship between the amounts of radiolabel adsorbed by the
inert solid phase and unoccupied serum protein-binding sites. The percent uptake
derives from the ratio of tracer bound by the adsorbent to the tracer bound by
serum proteins; an alternative but less reliable formula expresses the ratio as the
8 Sherman

amount of tracer attached to adsorbent to the amount initially added. The THBR
is then calculated as the percent uptake in the patient’s serum and normalized
to that of a control or reference serum; the expected normal range is centered
around unity. The THBR is increased when there are few endogenous binding
sites, which can occur with an increased amount of T4 available to bind (thy-
rotoxicosis); the presence of competing ligands (certain drugs and nonthyroidal
illness); or a decreased amount of binding protein (TBG deficiency). Conversely,
hypothyroidism and TBG excess will produce an increased number of available
binding sites, producing a decreased THBR. As a general rule, true thyroid func-
tion abnormalities produce concordant increases or decreases in the total serum
T4 and THBR, whereas discordant changes in the two tests typically result from
protein-binding abnormalities. Alternate methods use nonisotopic labels, such as
enzyme-linked tracers and light emitters. These all rely on the similar principle of
estimating the partitioning of labeled hormone between serum-binding proteins
and a solid phase. A free hormone index is estimated by multiplying the total
serum hormone concentration by the THBR. In most conditions of endogenous
thyroid function abnormalities or protein-binding alterations, the index corrects
for effects of protein binding on total T4 levels and correlates well with free T4
levels measured by reference methods.
Potential pitfalls in the interpretation of THBR tests occur when there is
a ligand that can interfere with binding to both the solid phase and the serum
proteins, for example, nonthyroidal illness. Falsely elevated free thyroxine index
values can also be present when the protein-binding abnormality is specific for T4
and masked by the use of T3 in the THBR—for example, familial dysalbuminemic
hyperthyroxinemia (FDH), in which an abnormal albumin binds only thyroxine
with high affinity. Similarly derived from the total T3, the “free T3 index” can be
useful in evaluating cases of abnormal serum binding.
Causes of Increased T4 and/or T3 Concentrations
The majority of patients with hyperthyroidism, regardless of the etiology, have
increased total serum concentrations of both T4 and T3, as well as high levels
of the free hormones (Table 1). In a minority of cases, there may be an isolated
elevation of either iodothyronine. T3-toxicosis is especially prominent in patients
with mild and recurrent Graves’ disease or hyperfunctioning adenomas and those
patients overtreated with triiodothyronine-containing thyroid hormone prepara-
tions. The relative magnitude of T3 elevation is often greater than T4 in forms of
hyperthyroidism caused by increased glandular synthesis of hormone; in Graves’
disease, the proportion of circulating T3 that derives from thyroidal production
nearly doubles (15). The opposite—that is, a lower T3:T4 ratio—is true in thyro-
toxicosis due to an inflammatory thyroiditis, in which there is release of previously
formed hormone, iodide-induced hyperthyroidism, and iatrogenic thyrotoxicosis
due to exogenous levothyroxine administration. Mild hyperthyroxinemia can even
be seen in patients being treated with exogenous levothyroxine for hypothyroidism
but whose TSH levels are normal on therapy (16).
The Laboratory Approach to Thyroid Disorders 9

Table 1 Causes of Increased T4 and/or T3 Concentrations


Thyrotoxicosis
Euthyroid hyperthyroxinemia
Increased binding to plasma proteins
TBG excess
Congenital
Hyperestrogenemia: exogenous, endogenous
Acute and chronic active hepatitis
Acute intermittent porphyria
HIV-1 infection
FDH
Transthyretin excess
Congenital
Paraneoplastic
Antithyroxine immunoglobulins
Rheumatoid factor (ex vivo)
Impaired T4 to T3 conversion
Iodinated contrast agents
Amiodarone
Propylthiouracil
Glucocorticoids
Propranolol
Congenital
Generalized resistance to thyroid hormones
Nonthyroidal illness
Acute psychosis
Acute medical/surgical illness
Hyperemesis gravidarum
Lead intoxication
Drugs
Clofibrate
5-fluorouracil
Perphenazine
Methadone
Heroin
L-thyroxine therapy

Increased total T4 concentrations without thyrotoxicosis, termed euthyroid


hyperthyroxinemia, result from both acquired and congenital etiologies. One com-
monly encountered situation is acquired TBG excess due to hyperestrogenemia.
Elevated hepatic exposure to estrogen leads to increased sialylation of carbohy-
drate side chains of TBG, thereby decreasing clearance of the glycoprotein and
increasing serum TBG levels. This effect is seen within several weeks of the onset
of hyperestrogenemia and can occur with exogenous administration of estrogens,
increased endogenous production (e.g., pregnancy), and even administration of
10 Sherman

selective estrogen receptor modulators, such as tamoxifen and raloxifene (17,18).


Exogenous estrogen administered transdermally, by avoiding first pass metabolism
in the liver, does not cause elevated TBG levels and hyperthyroxinemia (19).
Acquired TBG excess may also be responsible for the slight increase in T4 levels
reported in male cigarette smokers (20). X-linked inherited TBG excess occurs
with a frequency of 1 in 25,000 newborns, and can cause up to 2.5-fold elevations
in the total serum concentration of T4. Other abnormal serum-binding proteins can
contribute to euthyroid hyperthyroxinemia. In the autosomal dominant condition
FDH, one or more abnormal species of albumin contain a high-affinity binding
site for thyroxine. Because the defect is specific for T4 and does not affect T3
binding, these patients have an elevated total T4; a normal THBR using T3, but a
decreased THBR using T4 as the ligand; a normal total T3; and either a normal
or increased free T4, depending on the type of direct assay used. Equilibrium
dialysis typically yields normal levels of free T4 in this syndrome. The diagnosis
is established by paper or gel electrophoresis of serum enriched with radiolabeled
T4, which permits identification of the abnormal binding proteins.
Elevations of free T4 concentrations can occur as a result of interference
in binding to serum proteins. In vivo, hormone can be displaced from protein by
medications such as furosemide, causing a true, albeit rapidly reversible, minimal
hyperthyroxinemia after rapid intravenous administration of the diuretic. Activa-
tion of lipases by both low– and high–molecular weight heparins leads to increased
levels of free fatty acids that displace thyroid hormones both in vivo and ex vivo,
the latter situation causing an artifactual elevation of measured free hormone (21).
In autoimmune thyroid diseases and monoclonal gammopathies, endoge-
nous serum anti-T4 or anti-T3 antibodies bind thyroid hormones, increasing the
serum concentrations of protein-bound hormones. More commonly, however, anti-
iodothyronine autoantibodies have negligible in vivo effects on hormone binding,
but interfere with immunoassay measurements (22). In a classic RIA for total
hormone concentration, the autoantibody will compete with the capture antibody
for radiolabeled ligand, reducing the amount of signal available to be measured
and leading to a false high value. A similar spuriously increased result can occur
in one-step free T4 assay, in which the autoantibody binds the labeled T4 ana-
logue, preventing it from being measured and yielding a falsely increased free T4
level; this is avoided in a two-step assay in which the labeled ligand is unable to
interact with the serum autoantibodies. Another autoantibody that interferes with
immunoassays is the rheumatoid factor, an IgM directed against the Fc fragment of
human IgG. Because rheumatoid factor is weakly heterophilic, it appears to bind
to the nonhuman capture antibody, preventing interaction with the radiolabeled
ligand and leading to a falsely increased hormone concentration (23). Preincuba-
tion of the serum specimen with a nonspecific animal immunoglobulin, ethanol,
or polyethylene glycol reduces this antibody-mediated interference.
Decreased function of the 5’-monodeiodinase causes impaired conversion
of T4 to T3, decreasing T4 clearance and increasing T4 levels. Iodinated radio-
contrast dyes—for example, sodium ipodate—are potent inhibitors of T4 to T3
The Laboratory Approach to Thyroid Disorders 11

conversion and have been used therapeutically in severely hyperthyroid patients,


but are no longer commercially available in the United States. Amiodarone, a
highly iodinated antiarrhythmic agent, also interferes with T4 deiodination. Since
amiodarone-induced hyperthyroidism can also occur, great care must be taken in
interpreting hyperthyroxinemia in patients receiving iodinated medications (24).
An inherited defect in 5’-monodeiodinase function, due to mutation in a seleno-
cysteine insertion sequence binding protein, has recently been described, and is
probably responsible for hyperthyroxinemia observed in these patients (25).
Patients with resistance to thyroid hormones have an inherited partial defect
in tissue responsiveness to thyroid hormones. Serum concentrations of total and
free thyroid hormones are both increased as compensation for partial resistance.
Most kindreds that have been evaluated have been found to have a dominant neg-
ative mutation in a single allele of the thyroid hormone receptor beta. Although
affected individuals are generally described as being clinically euthyroid, consid-
erable variation exists in the measurable degrees of hormone resistance among
specific target organs for thyroid hormone (26).
Transient elevations of total serum T4 and, less frequently, free T4 levels
occur in patients with acute medical and psychiatric illnesses. Although some
patients develop increased levels of both T4 and T3 when the nonthyroidal illness
resolves, consistent with coexistent hyperthyroidism, in most of these patients
normal thyroid hormone levels are restored with recovery (27). Transient increases
in total and free T4 and T3 can be seen in 8% to 33% of patients admitted for acute
psychiatric disorders (28,29). TSH concentrations have been reported as increased
in up to 10% of acutely psychotic patients (30), but they are frequently suppressed
in severely depressed outpatients as well as those suffering from posttraumatic
stress disorders (31,32).
Causes of Decreased T4 and/or T3 Concentrations
Reduced serum levels of total and free T4 and T3 are typically seen in patients with
overt hypothyroidism, reflecting impairment of hormone synthesis and release by
the gland (Table 2). Due to TSH stimulation of residual gland function and eleva-
tion in the fractional conversion of T4 to T3 by 5’-monodeiodinase in both thyroid
and peripheral tissues, 30% of patients with primary hypothyroidism maintain
normal T3 levels despite decreases in T4. Thyroxine synthesis is also suppressed
in patients receiving T3 exogenously or with autonomous T3 overproduction.
Euthyroid hypothyroxinemia can be due to a variety of mechanisms. Anal-
ogous to the abnormalities that can cause hyperthyroxinemia, defects in hormone
binding to serum proteins can lead to decreases in T4 levels. Partial deficiency
of TBG, caused by impaired production or accelerated degradation of unstable
variants, occurs in 1 in 4000 births. X-linked complete TBG deficiency is less
common, found in 1 in 15,000 male births; female heterozygotes have TBG lev-
els that are partially reduced. Numerous variants of TBG with reduced affinity
for thyroid hormones have been described, with varying frequencies in different
populations (33). Acquired impairment of hormone binding develops secondary
12 Sherman

Table 2 Causes of Decreased T4 and/or T3 Concentrations


Hypothyroidism
Euthyroid hypothyroxinemia
Decreased binding to serum proteins
TBG deficiency
Chronic liver disease
Congenital
Cushing’s syndrome
Drugs
L-Asparaginase
Androgens
Nicotinic acid
Growth hormone excess
Nephrosis
Protein-losing enteropathy
TBG and transthyretin variants with reduced affinity
Inhibition of T4 binding by drugs
Carbamazepine
Diphenylhydantoin
Fenclofenac
Furosemide
Heparin
Meclofenamic acid
Mefenamic acid
Salicylates
Sertraline
Exogenous T3 or triiodothyroacetic acid administration
Nonthyroidal illnesses

to decreases in binding protein levels, due to either reduced production (as occurs
in hyperthyroidism) or increased clearance (as from nephrotic syndrome). In most
patients with quantitative or qualitative defects in TBG, direct and indirect esti-
mates of free T4 levels are normal. In the extreme case of complete deficiency, lack
of a linear relationship between free T4 fraction and THBR leads to falsely low
free T4 index results, and values of free T4 can be either normal or underestimated
by two-step and direct measurements.
Hypothyroxinemia and hypotriiodothyroninemia are common findings in
patients with nonthyroidal illness, with more severe reductions in total hormone
levels associated with more severe or critical illness (34,35). Milder degrees of
illness are typically accompanied by reductions in T4 to T3 conversion, resulting
in a low T3 state but preservation of T4 levels. In addition to deficiency of albu-
min and transthyretin, other proposed mechanisms include inhibition of hormone
binding to TBG, perhaps due to certain free fatty acids released from damaged
tissues or cytokines, such as tumor necrosis factor (36). Numerous medications
The Laboratory Approach to Thyroid Disorders 13

interfere with thyroid hormone binding to serum proteins, including diphenyl-


hydantoin, furosemide, heparin, sertraline, and certain nonsteroidal anti-
inflammatory agents (37,38). Inhibition of 5’-monodeiodinase activity in non-
thyroidal tissues accelerates clearance of T4 through nondeiodinative mecha-
nisms, particularly in nonthyroidal illness and starvation, and may be secondary to
increased levels of interleukin-6; the production rate of T3 declines as a result of
this monodeiodinase inhibition, but no change is seen in T3 metabolic clearance
(39). Medications such as glucocorticoids, amiodarone, oral radiocontrast agents,
gold, and high-dose propranolol and propylthiouracil (PTU) also inhibit T4 deio-
dination to T3; however, clinical signs of hypothyroidism are unlikely to develop.
Hypothyroxinemia has been described in patients treated with novel anticancer
agents that inhibit vascular endothelial growth factor receptors, with evidence
of multiple potential mechanisms that include primary thyroid dysfunction, but
also effects on either thyroid hormone absorption or metabolic clearance (40,41).
Pituitary TSH production is suppressed by endogenous and/or exogenous gluco-
corticoids, dopamine, somatostatin, and endorphins and may also be mediated
by reduced hypothalamic TRH secretion (42). Alteration of TSH sialylation and
bioactivity may occur in critical illness as well (43). With increasing severity of
nonthyroidal illness, all of the proposed mechanisms presumably result in a low
T4, low T3 state. Often, the decrease in protein binding is reflected by a decreased
T4 and increased THBR, yielding a normal free thyroxine index. However, in
many instances, the presence of a binding inhibitor (such as heparin or free fatty
acids released in inflammation) interferes with hormone attachment to the solid
phase, leading to a slightly lower value for the THBR and a falsely low estimate
of the free thyroxine index. Most analogue and some two-step procedures for
measuring free T4 are also adversely affected by binding inhibition in nonthy-
roidal illness (6,12). These laboratory abnormalities reverse with recovery from
the nonthyroidal illness or discontinuation of the interfering medication. Although
most of the effects of nonthyroidal illness may represent energy-conserving adap-
tive mechanisms, the traditional view of these patients as being euthyroid is not
universally held (44). However, no benefit from thyroid hormone supplementation
has yet been demonstrated.

Assays of Thyroid-Stimulating Hormone


Early TSH assays used a single polyclonal antibody in a radioimmunoassay and
were capable of detecting elevated levels of TSH in patients who have primary
hypothyroidism. With a sensitivity of approximately 1 mU/L, these tests were
unable to distinguish the low-normal TSH levels in serum of 25% of euthyroid
individuals from subnormal concentrations. With the introduction of immunomet-
ric (IMA) methods that use two or more antibodies directed at different antigenic
determinants on the TSH molecule, assay sensitivities have been improved by 10-
to 200-fold. The first antibody, usually a mouse monoclonal construct, is linked to
a solid phase, permitting the target molecule to be separated from the serum with
14 Sherman

high affinity; the second antibody, which may be polyclonal, is labeled, providing
a signal proportional to the amount of ligand bound. With these more sensitive
assays, hyperthyroid patients can be identified on the basis of low or undetectable
levels of TSH in IMAs, analogous to detection of primary hypothyroidism with
elevated TSH levels. Even more sensitive determinations of low TSH values have
been obtained in an assay using a chemiluminescent acridinium ester to gen-
erate the antibody-linked signal. High intra-assay and interassay precision with
chemiluminometric methods may permit routine detection of TSH levels as low as
0.01 mU/L or lower.
The ability of TSH assays to measure accurately low concentrations of
the hormone is termed the “functional sensitivity” of the assay, defined as the
concentration at which the interassay coefficient of variation is 20%. This contrasts
with the “analytical sensitivity,” which is based on intra-assay measurements of the
blank calibrator, and does not reflect a clinically meaningful result (7). Whereas the
original RIA methods have been termed “first generation” assays, the newer, more
sensitive TSH assays, which provide a sufficient separation in serum TSH values
between hyperthyroid and euthyroid patients, are defined as “second generation”
when the functional sensitivity is 0.1 mU/L, and “third generation” when the
functional sensitivity is 0.01 mU/L (45).
Multiple sources contribute to the total variation observed in TSH assay
results (46). Endogenous, biologic variation exists due to heterogeneity of TSH
isoforms, based on posttranslational modifications that can alter the immunoreac-
tivity as well as the bioactivity of the molecule; this potentially may be overcome
with use of variants of recombinant TSH that mimic these individual modifications
(47,48). Circadian and seasonal effects contribute to within-person variation as
well. But, within-person variation during serial measurements is relatively min-
imal compared with between-person variation, raising concern that population
reference standards may be inadequate to distinguish a healthy from diseased
state (46,49,50).
Debate now exists about the optimal reference range for TSH assays. Typi-
cally, the lower and upper limits of a population reference range of the analyte’s
concentrations are the 2.5th and 97.5th percentiles (the 95% confidence interval),
measured in a rigorously defined normal cohort without any evidence of relevant
disease. Applying this criterion to TSH levels, as determined in the recent U.
S. National Health and Nutrition Examination Survey (NHANES III), the pop-
ulation reference range would be 0.45 to 4.12 mU/L (51). Similar ranges have
been reported in other populations, differing to some degree due to variations
in iodine intake, ages, gender, and even the time of day that blood is sampled
(52). As most functional thyroid disorders are due to autoimmune thyroid dis-
ease, the relationship between levels of thyroid autoantibodies and TSH has also
been evaluated, demonstrating a U-shaped curve with the lowest prevalence of
autoantibodies at TSH levels between 0.1 and 1.5 mU/L in women and 0.1 and
2.0 mU/L in men (53). Additionally, the likelihood of eventual development of
overt primary hypothyroidism has been reported to be markedly higher in the
The Laboratory Approach to Thyroid Disorders 15

setting of a TSH level of at least 2.0 mU/L and elevated levels of anti-TPO anti-
bodies (54). Therefore, it has been proposed that the upper limit of the population
reference range should in fact be as low as 2.5 or 3.0 mU/L (55,56). Other studies
have suggested that age-specific reference ranges would be appropriate, with the
97.5th percentile being well above 4.5 mU/L with successively increasing deciles
of age (57). However, in the absence of definitive evidence that defining hypothy-
roidism as a TSH ⬎2.5 leads to unequivocal clinical benefit from treatment with
thyroid hormone, and given the overall concern that the population reference range
may not be optimal for defining a disease state when interindividual variation is
relatively large, it does not appear that changing the population reference range is
appropriate at this time (58).
Interference with TSH immunoassays is uncommon. Patients with endoge-
nous heterophilic antibodies directed against mouse immunoglobulin can have
falsely elevated TSH levels, as the heterophilic antibody can substitute for TSH
and bridge between the two antibodies in the assay (59). This problem has been
eliminated from most commercially available kits by addition of an excess of
mouse immunoglobulin. If interference with the assay is suspected, measurement
of serial dilutions of the sample may show a nonlinear relationship; alternatively,
the sample can be tested using another manufacturer’s assay (7).
Causes of Hypothyrotropinemia
When conventional TSH radioimmunoassays were used, the inability to differen-
tiate hyperthyroid from euthyroid patients was overcome by measurement of the
TSH response to stimulation with exogenous TRH. After the intravenous admin-
istration of 500 ␮g of TRH, TSH levels normally increase to 7 to 20 mU/L at
30 minutes. TSH levels usually remain undetectable in hyperthyroid patients when
measured by conventional RIA after TRH stimulation. However, recent studies
with sensitive assays have demonstrated that there is persistence of the logarith-
mic relationship between basal and TRH stimulated TSH levels even within the
hyperthyroid range (45). In severe hyperthyroidism, TSH levels remain below the
functional sensitivity of even third or fourth generation assays, but such degrees
of suppression are not seen in other causes of low TSH levels. Therefore, in most
cases, the presence of a basal serum TSH level of ⬍0.01 mU/L obviates the need
to perform a TRH-stimulation test to diagnose hyperthyroidism. Subnormal but
detectable TSH levels can be seen in patients who have mild or asymptomatic
hyperthyroidism of any etiology, or they may be due to TSH suppression from
nonthyroidal illness. TRH stimulation testing in acutely ill patients is usually
unable to distinguish mild hyperthyroidism from the effects of nonthyroidal ill-
ness. More sensitive TSH immunoassays provide adequate separation between
hospitalized hyperthyroid patients with medical illness, in whom basal TSH lev-
els generally remain undetectably low, and euthyroid patients with nonthyroidal
illness, in whom basal TSH levels are usually but not always ⬎0.01 mU/L.
In hypothyroidism due to hypothalamic or pituitary disease, low levels of
basal TSH may occur. It was initially suggested that TRH stimulation testing was
16 Sherman

able to classify central hypothyroidism as either hypothalamic or pituitary in origin,


with the former showing a delayed and exaggerated TSH response 30 to 60 minutes
after TRH administration, and the latter having a persistently blunted response.
In practice, there was an overlapping spectrum of TRH responses in patients with
pituitary and hypothalamic diseases, and TRH is no longer readily available for this
purpose. Hypothyroidism due to pituitary or hypothalamic disease can also present
with inappropriately normal or even slightly elevated levels of immunologically
intact but biologically inactive TSH secondary to alternative glycosylation of the
protein (60,61). Among the drugs that can affect TSH production, the rexinoid
bexarotene appears to suppress TSH gene transcription directly and causes a
dose-dependent central hypothyroidism (62,63). Recently, the hypoglycemic drug
metformin was reported to lower TSH levels by an as yet unknown mechanism
(64,65).
Causes of Hyperthyrotropinemia
Elevated serum TSH values are the cornerstone of the diagnosis of primary
hypothyroidism. Due to the extreme sensitivity of the hypothalamic-pituitary-
thyroid negative feedback loop, small decrements in circulating thyroid hormone
levels produce logarithmic increases in serum TSH levels (45). At one end of
the spectrum are patients with frankly symptomatic thyroid hormone deficiency,
whose free T4 levels are subnormal and whose TSH levels are typically
⬎20 mU/L. But, even patients with the earliest stages of thyroid gland impairment
can have elevated TSH concentrations. These patients with so-called subclinical
hypothyroidism have T4 and T3 levels within the normal range associated
with increased serum TSH concentrations. Although the clinical management
of such patients remains controversial, those individuals with a predisposition
to developing clinical hypothyroidism—for example, those with autoimmune
thyroiditis or a history of thyroid irradiation or surgery, should be treated or
followed longitudinally for development of overt hypothyroidism. Medications
that have been associated with hyperthyrotropinemia include cytokines that can
cause autoimmune thyroiditis (such as interferon-alpha) and tamoxifen, although
the latter appears to induce a mild and transient increase in TSH levels that is
not seen with raloxifene (17,66). In neonates, various maternal causes of fetal
distress, including preeclampsia and gestational diabetes mellitus, are associated
with elevated TSH levels in cord blood, but whether this reflects transient primary
hypothyroidism or a central stimulation of TSH production is unknown (67).
The differential diagnosis of hyperthyrotropinemia also includes conditions
associated with inappropriate TSH secretion, as in patients whose TSH levels are
higher than would be predicted from their circulating free thyroid hormone levels.
Patients with TSH-secreting pituitary adenomas may have normal or increased
TSH levels in the setting of increased T4 concentrations. These patients usually
present with a goiter and clinical evidence of thyrotoxicosis, with or without
clinical evidence of a sellar mass lesion. In half of cases, there is cosecretion of
other anterior pituitary hormones (e.g., growth hormone or prolactin) and the beta
The Laboratory Approach to Thyroid Disorders 17

subunit of TSH is commonly overproduced. A molar ratio of alpha subunit to


intact TSH that is greater than unity is strongly suggestive of a pituitary adenoma.
Although thyrotoxic with elevated levels of TSH and T4, patients with the rarer
syndrome of isolated pituitary resistance to thyroid hormone do not have radio-
graphic evidence of a pituitary tumor, and their ratio of alpha subunit to intact
TSH is usually ⬍1 (68).
Resistance of the thyroid to TSH, presenting with nongoitrous congenital
hypothyroidism and elevated TSH levels, has been described both in isolated form
and in pseudohypoparathyroidism type Ia (68). In this latter congenital condition,
deficiency of the stimulatory subunit of the guanine nucleotide-binding proteins
that mediate activation of adenylate cyclase can cause resistance to multiple hor-
mones, including TSH and parathyroid hormone.
In infants, exposure to cold temperatures immediately following birth or
during hypothermic surgery causes TSH concentrations to rise as high as 50 to
100 mU/L, which is thought to reflect the immaturity of the hypothalamic-
pituitary-thyroid axis (69). Adults, on the other hand, do not demonstrate altered
TSH levels after brief periods of cold exposure, despite increases in the concen-
trations and fractional clearance rates of circulating free T4 and T3.

Specialized Studies of Thyroid Function


Thyroglobulin
In most forms of thyroid disease, thyroglobulin (Tg) is released from thyroid fol-
licular cells proportional to the synthesis and release of T4 and T3, increasing size
of the gland, and the degree of cytotoxic inflammation. The reference range in
subjects with intact thyroid glands and normal TSH levels is approximately 3 to
40 ng/mL. Markedly elevated levels are seen in most patients with hyperthyroidism
and thyroiditis, but mild increases are also observed in cigarette smokers despite
slightly lower TSH levels (70). In determining the cause of hyperthyroidism, an
undetectable serum thyroglobulin suggests factitious or iatrogenic thyrotoxico-
sis. Undetectable levels are also seen in hypothyroid patients with congenital or
acquired absence of the thyroid. Presently, the primary indication for measurement
of serum Tg concentrations is as a tumor marker for the longitudinal follow-up
of patients with differentiated thyroid carcinoma, which necessitates greater func-
tional sensitivity at lower concentrations than the euthyroid reference range (71).
Although introduced more than 15 years ago, these assays are now being increas-
ingly used to detect Tg in fine-needle aspirations of neck masses or cystic lesions
as an adjunct to cytologic interpretation to diagnose recurrent or metastatic cancer
(72).
Serum Tg is generally measured by either two-antibody immunometric assay
or single-antibody immunoassay. The newer immunometric assays require shorter
incubation times and have greater sensitivity (≤1 ng/mL) than the immunoassays,
but several problems persist. The greatest limitation is the potential for interfer-
ence by anti-Tg autoantibodies, which can be found in up to 25% of differentiated
18 Sherman

thyroid carcinoma patients. In the immunometric assays, the serum Tg concen-


tration can be falsely lowered by autoantibodies that bind Tg and effectively
remove it from the serum, thus making it incapable of binding to the assay’s
reporter antibodies. Attempts to abrogate this effect by use of monoclonal anti-
bodies directed against epitopes of Tg that do not react with the autoantibodies
have been ineffective. Detecting the presence and degree of autoantibody inter-
ference in an immunoassay may also be difficult. Although “recovery tests” that
determine whether the addition of known quantities of exogenous Tg to a serum
specimen that contains anti-Tg antibodies results in an appropriate increase in the
concentration were initially purported to accurately detect the degree of interfer-
ence, subsequent studies have demonstrated that such assessments are unreliable
(73). Conversely, in radioimmunoassays, anti-Tg autoantibodies can cause falsely
high values because they bind radiolabeled Tg; as a result, less is available to
bind to the assay antibody. Thus, in the presence of anti-Tg antibodies, discordant
findings of an undetectable Tg in an immunometric assay and a concentration of
at least 2 ng/dL in a radioimmunoassay may suggest the presence of antibody
interference, but cannot be used to quantify the problem. Measure of serum Tg
should therefore always be preceded by a test for anti-Tg antibodies, and it is rec-
ommended that laboratories withhold reporting low results of Tg immunometric
assays when autoantibodies are identified (7). Of note, recent reports demonstrate
that the presence of anti-Tg antibodies may not preclude identification of the high
concentration of Tg seen in fine-needle aspiration specimens (74).
Despite a trend toward assay standardization, the variability of results using
differing assays remains at least 25%, due to variations in the antithyroglobulin
antibodies used and the molecular heterogeneity of Tg. Occasionally, immuno-
metric assays may fail to detect very high serum Tg concentrations due to the
so-called hook effect, in which the high concentrations of Tg bind to one antibody,
preventing the formation of the two-antibody sandwich upon which the assay
depends. If this effect is suspected, the sample should be reanalyzed after dilu-
tion. Another cause of a false-negative Tg in patients with differentiated thyroid
cancer can be tumor production of variants of Tg that fail to be recognized by the
antibodies used in an assay (75).
The negative predictive value of an undetectable Tg level to identify differ-
entiated thyroid cancer patients who have no remaining evidence of disease has
been markedly improved by the now routine practice of using TSH to try to stim-
ulate Tg production by any remaining tumor cells (76). Whereas higher Tg levels
result from more prolonged endogenous TSH stimulation during withdrawal from
thyroid hormone, quality of life is preserved without significant loss of diagnos-
tic utility when stimulation is provided by exogenous recombinant human TSH
administered to a patient taking thyroid hormone (77,78). Alternatively, the posi-
tive predictive value is limited in the presence of remnant normal thyroid cells left
after thyroidectomy, and thus one indication for postsurgical adjuvant radioiodine
therapy is to eliminate such normal sources of Tg (79). False-positive Tg results
can also be caused by heterophilic antibodies, a problem in many immunometric
The Laboratory Approach to Thyroid Disorders 19

assays that has only been partially resolved by the addition of blocking antibodies,
but rare false-negative results have also been reported (80,81).
Tg assays with functional sensitivity as low as 0.1 ng/dL or less have recently
been introduced (82,83). Such more sensitive assays have been proposed as a
replacement for use of recombinant TSH stimulation testing, but a limiting factor
remains the reduction in specificity that accompanies the higher sensitivity of
these procedures, with the possible requirement for multiple additional testing
procedures based upon false-positive Tg results.
Thyroid Autoantibodies
Antibodies directed against the cell surface (TSH receptor), intracellular compo-
nents (microsomal membranes, thyroglobulin), and extracellular antigens (T4, T3)
are often found in sera of patients with autoimmune thyroid diseases. Although
autoantibodies tend to target fewer antigenic epitopes than heterologous antibod-
ies, these autoantibodies can still be quite heterogeneous mixtures of proteins,
leading to problems with both specificity and sensitivity in assays.
In Hashimoto’s disease, cytotoxic antibodies may bind to a thyroid micro-
somal antigen that is expressed on the apical cell surface, and these antibodies
subsequently fix complement. These antithyroid microsomal antibodies can be
detected by sensitive hemagglutination techniques in the sera of 95% of patients
with histologically proven Hashimoto’s disease, as compared with only 55% for
non-complement-fixing antithyroglobulin antibodies. Among commercially avail-
able assays, immunometric procedures, including RIA, immunoradiometric assay
(IRMA), and enzyme-linked and fluorescent methods, are superior to routine
hemagglutination techniques. Marginal improvements in sensitivity and specificity
have been obtained using monoclonal antibodies directed against TPO, consistent
with evidence that this enzyme is the primary microsomal antigen. Factors limit-
ing the accuracy of anti-TPO assays include the presence of splicing variants of
the enzyme as well as differences in posttranslational modifications (84). Interna-
tional standardization now exists against a specific reference preparation, Medical
Research Council (MRC) 66/387, permitting reporting of results in “international
units,” but concordance among multiple assays remains suboptimal (85). Refer-
ence ranges vary widely among different assays, with manufacturers often citing
levels ⬎10 kIU/L as being clinically relevant predictors of autoimmune thyroid
disease. However, long-term follow-up studies that identified antimicrosomal anti-
bodies as being predictive of eventual hypothyroidism were likely based on far
less sensitive assays, and similar studies will be required to determine whether
such minimally detectable levels are also predictive (54,84).
Antithyroglobulin antibodies are less specific for autoimmune thyroiditis,
but have achieved greater significance for their potential to interfere with thy-
roglobulin assays in thyroid cancer patients. Contemporary immunoassays are
considerably more sensitive and specific than older, agglutination methods, and
can detect antithyroglobulin antibodies in up to 10% of the clinically disease-free
population and 3.4% of those who lack anti-TPO antibodies (51). Nevertheless,
20 Sherman

reference preparations for standardization of these assays still vary considerably,


and even use of the accepted international standard reference MRC 65/93 has not
resulted in interchangeabilty of assays (84). As with anti-TPO antibody measure-
ments, differences exist in the definitions used for reference ranges. Assays that
report detectable levels of antithyroglobulin antibodies ⬍10 kIU/L as abnormal
may have low specificity both for actual pathology and for antibodies that can
interfere with thyroglobulin assays (86).
No correlation exists between the severity of hypothyroidism and titers of
antithyroid antibodies, and low levels can be seen in patients with no demonstrable
thyroid dysfunction. Anti-TPO and antithyroglobulin antibodies are also present
in Graves’ disease, albeit less frequently (85% and 25%, respectively), and may
predict the subsequent development of hypothyroidism in some patients with this
condition. With appropriate treatment of the thyroid hormone excess or deficiency,
antithyroid antibody titers often decrease but are not clinically useful measures of
disease activity.
Multiple procedures have been developed to measure the TSH-receptor
stimulatory immunoglobulins that are pathogenetic for Graves’ disease, detecting
either stimulation of biochemical functions in thyroid cells (thyroid-stimulating
immunoglobulins) or blockade of receptor binding by TSH (TSH-binding
inhibitors). The original long-acting thyroid stimulator (LATS) assay of Adams
and Purves had been largely replaced by quantitation of cyclic adenosine
monophosphate (AMP) production, iodine uptake, or thymidine incorporation
into DNA in cell lines that derive from a rat thyroid follicular cell. False-negative
results can occur due to human IgG molecules that do not interact with the rat TSH
receptor—an outcome that can be circumvented by use of human thyroid cells in
primary culture or Chinese hamster ovary cells expressing a human TSH receptor
(87). TSH-binding inhibitors can be detected by quantitation of radiolabeled TSH
binding to solubilized porcine thyroid membranes (or more recently recombinant
human TSH receptors) in the presence of serum, followed by polyethylene gly-
col precipitation to separate bound from unbound radiolabel (88). Alternatively,
recombinant TSH receptors can be affinity-immobilized on an antibody-coated
tube, which is then incubated with TSH with an attached radioactive or chemilumi-
nescent label (89). In general, the most sensitive of these assays can detect thyroid-
stimulating immunoglobulins in up to 95% of hyperthyroid Graves’ sera, and
TSH-binding inhibitors in 60% to 85%. Recent reports differ as to the magnitude
of correlation between these two immunoglobulin types as well as their respective
utilities for diagnosis and monitoring of response to therapy for Graves’ disease,
despite recent adoption of an international standard MRC 90/672 (90). However,
thyroid-stimulating immunoglobulins levels may be more useful for identifying
Graves’ disease as the cause of exophthalmos (91). Blocking antibodies that bind
to but do not stimulate the TSH receptor have also been identified in hypothyroid
and euthyroid patients with autoimmune thyroiditis or Graves’ disease.
The measurement of thyroid autoantibodies is of value in selected clinical
situations. The presence of thyroid-stimulating immunoglobulins in patients in
The Laboratory Approach to Thyroid Disorders 21

whom the etiology of hyperthyroidism is uncertain can lead to a diagnosis of


Graves’ disease, although anti-TPO antibodies are also common in this condition
and may be a more cost-effective test. With the second generation radioreceptor
assays that use recombinant TSH receptors, anti-TSH receptor antibody levels
before treatment are better predictors of the likelihood of response to antithy-
roid drug therapy than earlier assays (92). Persistence of high levels of thyroid-
stimulating immunoglobulins in Graves’ disease following therapy is associated
with increased rates of recurrence (93,94). When detected during the third trimester
of pregnancy in a woman with Graves’ disease, significant increases in either TSH-
binding inhibitors or thyroid-stimulating immunoglobulins titers correlate with the
development of intrauterine and neonatal hyperthyroidism due to transplacental
passage of immunoglobulins. Recent evidence-based guidelines suggest that the
more easily measured TSH-binding inhibitors level is the preferred method for
predicting risk for neonatal hyperthyroidism (31).
In the setting of asymptomatic hyperthyrotropinemia with normal thyroid
hormone levels, significant anti-TPO antibody levels are prognostic for the devel-
opment of overt hypothyroidism in approximately 5% of patients per year; the
likelihood of thyroidal failure is higher in those patients with higher levels and
baseline TSH concentrations ⬎2 mU/L (54). The presence of serum anti-TPO
antibodies in a euthyroid pregnant woman greatly increases her risk of developing
symptomatic postpartum thyroiditis, as well as her risk for fetal loss (95).
Reverse T3
Radioimmunoassays for reverse T3 (rT3) have been developed, but have limited
clinical usefulness. Levels are increased in nonthyroidal illness due to impairment
of 5’-monodeiodination, a major degradative step for rT3, but this is insufficiently
reliable to distinguish euthyroid patients from those with coexisting hypothy-
roidism (96).
Tissue Responses to Thyroid Hormone Action
Before the availability of hormone immunoassays, measurement of the end-organ
responses—for example, the basal metabolic rate—was the primary means of
evaluating thyroid hormone function. Today, regulation of serum TSH levels by
T4 and T3 is the most precisely measurable and useful response by tissues to
the action of thyroid hormones. Measurements of thyroid hormone effects in
extrapituitary tissues are occasionally used to evaluate patients in whom there
is a discordance among the clinical evaluation, thyroid hormone levels, and the
concentration of TSH (97).
Numerous serum constituents have altered levels in hyperthyroidism and
hypothyroidism, mostly reflecting changes in synthesis and/or clearance of these
substances (Table 3). There is considerable overlap between the normal ranges and
values seen in thyroid gland dysfunction. However, they remain useful markers
of thyroid hormone effects, especially with serial determination during therapy
of underlying thyroid disorders and in the evaluation of patients with discordant
22 Sherman

Table 3 How Various Serum Constituents Are Altered in Hyperthyroidism


and Hypothyroidism

Increased Decreased

Hyperthyroidism
Alkaline phosphatase Cholesterol (total, LDL)
Angiotensin converting enzyme Apolipoprotein b, apo (a)
Calcium Corticosteroid-binding globulin
Factor VIII
Ferritin
Osteocalcin
Sex hormone–binding globulin
Urine nitrogen excretion
Urine pyridinoline cross-links
Hypothyroidism
Carcinoembryonic antigen Aldosterone
Cholesterol (LDL and HDL fractions) Angiotensin converting enzyme
Creatine phosphokinase Factor VIII
Creatinine Osmolarity
Lactic dehydrogenase Sex hormone–binding globulin
Myoglobin
Norepinephrine
Prolactin
Corticosteroid-binding globulin

thyroid function tests. Combinations of biophysical and serum parameters of


thyroid hormone action are particularly useful in the evaluation of patients with
possible thyroid hormone resistance states. To characterize the presence and
extent of resistance, parameters of pituitary and peripheral tissue response are
measured before and during administration of increasing doses of T3 (50, 100, and
200 ␮g/day). Among the various tests performed, changes in sex hormone binding
globulin, basal metabolic rate, and body weight provide the strongest distinction
between normal responsiveness and generalized resistance to thyroid hormones
(98).

LABORATORY EVALUATION FOR THYROID DISEASE


Distinct strategies for use of thyroid function tests should be designed to satisfy
four distinct purposes: screening for the presence of clinically unsuspected disease
in an asymptomatic general population, case finding to detect thyroid disease in
patients whose symptoms and signs are sufficiently subtle that the examining
clinician may not suspect thyroid dysfunction as the etiology, diagnosis to prove
the presence of clinically suspected disease, and optimization of management of
proven thyroid disease.
The Laboratory Approach to Thyroid Disorders 23

Screening and Case Finding


Population screening is generally warranted if the prevalence of such disease is
not small, the health consequences of undiagnosed disease are substantial, and
treatment is effective. With these criteria in mind, there is considerable contro-
versy about the appropriateness of screening asymptomatic adults for thyroid
dysfunction (99,100). The Whickham study demonstrated an annual incidence of
thyroid hormone excess and deficiency of 0.5% in women and 0.06% in men in
the United Kingdom (54). The hazard rate for developing thyroid dysfunction was
higher in women with advancing age, but not men. Using a logit model to evaluate
contributors to risk, only the presence of antithyroid antibodies and a baseline
TSH of at least 2.0 mU/L were predictive of eventual overt hypothyroidism. More
at issue than prevalence, however, is the question of whether undiagnosed mild
hypothyroidism or hyperthyroidism has significant enough consequences to jus-
tify the costs of screening. Using a decision analysis model, adding a serum TSH
determination to the quintennial cholesterol screening recommended starting at
age 35 was found to be reasonably cost-effective (101). Deferring periodic TSH
screening until older ages and decreasing cost for TSH assays are key factors in
improving cost-effectiveness even further. As a result, three endocrine professional
organizations (the American Thyroid Association, American Association of Clin-
ical Endocrinologists, and The Endocrine Society) all support routine screening
of asymptomatic adults (102). Conversely, other organizations with broader focus
than these endocrine groups do not recommend screening for thyroid dysfunc-
tion, including the U. S. Preventive Services Task Force, American College of
Physicians, Royal College of Physicians, and Institute of Medicine (103–106).
There is uniform agreement, however, that screening for neonatal hypothy-
roidism is necessary (107). Neonatal hypothyroidism occurs with a frequency of
1 in 4000 live births and is associated with significant neurological and develop-
mental morbidity, much of which can be prevented by early treatment with thyroid
hormone replacement. Mandatory neonatal screening is based upon the measure-
ment of total (not free) T4 in whole blood collected on filter paper. Determination
of TSH concentration is performed if the T4 level is below the 10th percentile, and
serum assays are used to confirm a diagnosis of hypothyroidism. The advantage
of a T4-first strategy is the ability to detect central hypothyroidism and minimized
impact of the neonatal TSH surge (108). An alternative strategy employs pri-
mary TSH screening, followed by confirmatory T4 testing; this approach is more
commonly used in Europe and in areas of iodine deficiency (109).
Case finding is best reserved for patients whose clinical assessment may be
sufficiently complex as to obscure suspicion for thyroid dysfunction. Often, these
patients are elderly and their symptoms may be primarily constitutional, neu-
ropsychiatric, or cardiovascular. Although dementia is an uncommon presentation
of hypothyroidism, the relative ease of diagnosis and treatment of this condition
warrants inclusion of a thyroid function test in the evaluation of such patients. As
an initial test for case finding, a sensitive TSH assay has excellent sensitivity and
24 Sherman

specificity for both hyperthyroidism and hypothyroidism. In contrast, hospitalized


patients with acute illnesses have a high frequency of transient thyroid function
abnormalities and are unlikely to have primary thyroid disease diagnosed on the
basis of routine tests. In the absence of strong clinical evidence of thyroid dys-
function, patients hospitalized with acute illnesses should probably not undergo
thyroid testing for case finding (34).
Postpartum women have a high frequency of transient thyroid dysfunc-
tion, especially those with preexisting euthyroid autoimmune thyroiditis. Within
the first three months after delivery, at least 5% of women develop postpartum
thyroiditis, a painless inflammatory condition that can cause thyrotoxicosis and/or
hypothyroidism. More than one-half of these patients require therapeutic interven-
tion. Furthermore, 25% of women with postpartum thyroiditis eventually develop
chronic hypothyroidism, thus requiring lifelong therapy. Case finding with serum
TSH measurements three and six months after delivery is recommended for women
with type I diabetes mellitus, personal history of postpartum thyroiditis, or those
known to have elevated levels of anti-TPO antibodies (110).

Hypothyroidism
Diagnosis
Clinical manifestations, hypothyroxinemia, and an increased serum TSH establish
the diagnosis of primary hypothyroidism. Nonspecific clinical symptoms and signs
alone are a poor predictor of hypothyroidism. Among ambulatory patients evalu-
ated for possible hypothyroidism—with symptoms of weight gain, fatigue, men-
strual irregularities, depression, cold intolerance, constipation, or galactorrhea—
only 4% have increased TSH levels and fewer than half of these have values
⬎5 mU/L above normal (111). The presence of a goiter on examination increases
the likelihood of hypothyroidism sixfold. However, the number and severity of
symptoms are poorly correlated with presence or severity of hypothyroidism.
Given this low prevalence of disease in symptomatic patients, first-line laboratory
testing for suspected hypothyroidism should be a highly sensitive, inexpensive test
capable of excluding patients who would not benefit from therapy.
In patients with hypothyroid symptoms who lack a history of known thy-
roid or pituitary disease, the initial diagnostic test should be the serum TSH,
with confirmation of diagnosis provided by subsequent determination of the
free T4 level. Although rare patients with secondary hypothyroidism may not be
detected, a serum TSH assay will accurately diagnose hypothyroidism in almost
all patients, including those with subclinical hypothyroidism. An alternative strat-
egy for patients who are not suspected to have thyroxine binding or clearance
abnormalities would be measurement of the total T4 with an attenuated normal
range, but this strategy is not generally recommended. Confirmation of the diag-
nosis of primary hypothyroidism can then be made with a TSH assay, a test with
specificity and sensitivity for primary hypothyroidism of ⬎95%.
The Laboratory Approach to Thyroid Disorders 25

Other clinical settings present different testing requirements. Hypothyrox-


inemia can be due to decreased serum binding to proteins, nonthyroidal illness
or certain medications. Similarly, hypothyroid patients with increased serum thy-
roxine binding due to increased TBG may have normal total T4 levels that mask
the complexity of the underlying thyroid disorders. If abnormal serum binding is
suspected, for example, in patients taking estrogen therapy, measurement of serum
TSH is the most appropriate initial test. When the prevalence of primary hypothy-
roidism is higher than that in the general ambulatory population, as in patients
attending an endocrinology clinic or those who present with a goiter or a history of
thyroid disease, lithium therapy, neck surgery, or high-dose radiation exposure, the
sensitive TSH assay may be the most efficient initial test to perform, followed by
free T4 when TSH is elevated. Hypothyroidism due to pituitary or hypothalamic
disease is more common as well in an endocrinology clinic, and therefore a TSH
assay as the sole evaluation of the pituitary-thyroid axis can be misleading given
the frequent lack of elevation of TSH concentrations in this disorder (112). When
secondary or tertiary hypothyroidism is also a consideration, the combination of
serum free T4 and TSH should be measured.
Additional tests can occasionally be helpful in establishing the diagnosis and
cause of hypothyroidism. The diagnosis of autoimmune thyroiditis, either chronic
or transient, is confirmed by the presence of serum anti-TPO antibodies. Although
most hypothyroid patients who develop nonthyroidal illness remain hyperthy-
rotropinemic, individuals with mild hypothyroidism may have normal TSH con-
centrations in this situation, especially when exposed to the TSH-suppressive
effects of exogenous and endogenous glucocorticoids, dopamine, somatostatin,
and opioids. Confirmation of the diagnosis may require follow-up testing when
the patient has recovered from the acute illness. If pituitary or hypothalamic disease
is suspected, other abnormalities of anterior pituitary function are often found—
for example, inappropriately normal gonadotropin levels in a postmenopausal
woman, radiological imaging of the sella is also often indicated. When the clinical
diagnosis of hypothyroidism is uncertain, supporting evidence can be obtained
by examining tissue responses to thyroid hormones—for example, sex hormone
binding globulin.
Treatment
Thyroid hormone replacement therapy for the treatment of primary hypothy-
roidism should be monitored with a sensitive TSH assay. The goal of therapy
for hypothyroidism is to make the patient clinically and biochemically euthyroid,
which is usually associated with normalization of the serum TSH level; the target
TSH range most commonly recommended by thyroidologists is 0.5 to 2.0 mU/L
in the nonelderly and 0.5 to 4.0 mU/L in the elderly (113). When levothyroxine
sodium is administered daily by mouth, decreasing serum TSH concentrations
plateau within six to eight weeks, and measurement of the TSH level should await
this new equilibrium. Patients with persistently elevated serum TSH levels gener-
ally require an increased dose of levothyroxine, whereas those patients with a low
26 Sherman

serum TSH concentration usually require a decrease in dose. Serum T4 and T3


levels are usually normal during therapy, although mild hyperthyroxinemia can
occur in as many as 20% of patients who are otherwise euthyroid. Once a hypothy-
roid patient becomes euthyroid, follow-up evaluation should be performed after
several months, given the gradual increase of T4 clearance that occurs in these
patients. Subsequent evaluations generally are recommended annually, although
stable patients may be monitored less frequently (114). To monitor therapy in
patients with secondary or tertiary hypothyroidism, the patient’s clinical status
and serum free T4 levels should be assessed. The goal of therapy should be attain-
ment of a clinically euthyroid state, with normal or high-normal free T4 levels,
since the serum TSH concentration is not a reliable indicator of thyroid hormone
status in this setting (115).
Patients receiving therapy with T3-containing preparations, which are
shorter-acting than levothyroxine sodium alone, may be more difficult to monitor
(116). Individuals who take desiccated thyroid or other formulations containing
triiodothyronine are likely to have variable T3 levels during the day.

Hyperthyroidism
Diagnosis
Clinical manifestations, increased serum thyroid hormone concentrations, and a
subnormal serum TSH level typically establish the diagnosis of thyrotoxicosis. In
the evaluation of the patient with hyperthyroid symptoms, two alternative testing
strategies can be followed. Measurement of free T4, if near or above the upper
range of normal, can detect hyperthyroxinemic patients in whom a low serum
TSH level can then confirm the diagnosis of hyperthyroidism. With this approach,
most patients with euthyroid hyperthyroxinemia due to increased hormone binding
to serum proteins are identified on the basis of a normal free T4. Patients with
FDH, nonthyroidal illness, and generalized resistance to thyroid hormones may be
recognized by their usually normal serum TSH concentrations. However, patients
with T3 toxicosis, due to mild or recurrent Graves’ disease or hyperfunctioning
adenomas, can present with normal or occasionally low serum free T4, and the
diagnosis of hyperthyroidism may be missed using the T4-first strategy.
Alternatively, the serum TSH level, measured in an assay with a detection
limit of ⬍0.1 mU/L, is a more sensitive initial test for the diagnosis of thyro-
toxicosis. In patients with subnormal TSH levels, increased serum T4 and/or T3
levels confirm the diagnosis of thyrotoxicosis, high-normal T4 and/or T3 sug-
gest subclinical thyrotoxicosis, and low-normal or low–thyroid hormone levels
suggest either T3 thyrotoxicosis, nonthyroidal illness, or central hypothyroidism
(58). Patients with the rare causes of TSH-mediated hyperthyroidism who may
have normal or even elevated TSH levels—that is, TSH-secreting pituitary ade-
noma or isolated pituitary resistance to thyroid hormones—may be undiagnosed
with this strategy. Thus, in settings where unusual causes of hyperthyroidism are
more prevalent, such as an endocrinology or neurosurgical clinic, simultaneous
The Laboratory Approach to Thyroid Disorders 27

determinations of serum TSH and thyroid hormone concentrations may be the


most appropriate initial tests to perform.
Additional tests are occasionally useful in establishing the diagnosis and
etiology of thyrotoxicosis. When the clinical and biochemical evidence for thyro-
toxicosis are discordant, supplemental evidence for excess thyroid hormone levels
can be obtained by examining tissue responses to thyroid hormones that are altered
in thyrotoxicosis, such as sex hormone–binding globulin. Once the diagnosis is
certain, the clinical presentation of hyperthyroidism may suggest the need for
ancillary testing. In a patient with a tender goiter and hyperthyroidism of short
duration, an elevated erythrocyte sedimentation rate is characteristic of subacute
thyroiditis. An increased radioiodine uptake is typically found in Graves’ disease
and hyperfunctioning nodular disease but is not seen in subacute or lymphocytic
thyroiditis. Similarly, the serum thyroglobulin and radioiodine uptake are typi-
cally low in factitious and iatrogenic thyrotoxicosis. When clinical findings do
not provide adequate diagnostic distinction, testing for anti-TPO antibodies or
thyroid-stimulating immunoglobulins/thyrotropin receptor antibodies may assist
in the discrimination between Graves’ disease and toxic multinodular goiter. Radi-
ologic imaging of the sella and a serum alpha subunit level would be indicated in
the evaluation of the patient with TSH-mediated hyperthyroidism.
Treatment
The goal of therapy for hyperthyroidism is a clinically euthyroid patient with
normal serum thyroid hormone and TSH levels. However, because normalization
of the serum TSH concentration may lag behind thyroid hormone levels by several
months, initial therapy should be directed to reducing thyroid hormone levels. Once
the TSH level is detectable, it is again a sensitive indicator of thyroid hormone
status. Given the increased T3:T4 ratio typical of increased glandular synthesis of
hormone, the T4 level may become normal or even low while the T3 level remains
elevated, producing persistent T3 toxicosis. In addition, an elevated T3 level may
be the earliest sign of recurrent hyperthyroidism. Thus, both the T3 and free T4
levels should be monitored during treatment of hyperthyroidism. In patients with
TSH-mediated hyperthyroidism, clinical assessment and thyroid hormone levels
are the only measures of use, as the serum TSH level does not reflect thyroid
hormone status.

THYROID NEOPLASIA
Alterations in thyroid function are unusual in the presentation of thyroid neoplasia.
Most neoplastic lesions, both benign and malignant, organify iodine and produce
thyroid hormones inefficiently. Unless most of the nonneoplastic follicular tissue is
destroyed by infiltrating tumor or is ablated during the treatment of the neoplasm,
thyroid hormone levels are normal. Autonomously hyperfunctioning thyroid ade-
nomas can produce hyperthyroidism, thereby suppressing serum TSH levels and
reducing hormone production by nonadenomatous tissue. However, only 5% of
28 Sherman

thyroid nodules are hyperfunctioning, and therefore most patients with thyroid
nodules are euthyroid. Fine-needle aspiration biopsy is the most important diag-
nostic test for the solitary or dominant thyroid nodule, capable of distinguishing
benign from malignant lesions in a majority of patients.
In a patient with differentiated thyroid carcinoma who has undergone total
thyroidectomy or remnant ablation with radioactive iodine, serial thyroglobulin
determinations can assist in the recognition of residual or metastatic disease. In
general, the sensitivity of detecting thyroid carcinoma by measurement of serum
Tg after discontinuation of thyroid hormone therapy is 85% to 95% but may be as
low as 50% during therapy. Levels of thyroglobulin following recombinant human
TSH stimulation may be approximately 50% as high as following endogenous TSH
stimulation after thyroid hormone withdrawal, when performed in conjunction
with diagnostic radioiodine scanning (117). The results are most likely to be
falsely negative in patients with small nodal metastases of papillary carcinoma
and in those with tumor dedifferentiation. In immunometric assays, false-negative
Tg results can also occur due to interference from antithyroglobulin antibodies.
Antithyroglobulin antibodies that persist or are rising beyond the first several
years of follow-up after primary therapy can indicate higher risk for residual
disease (86). However, persistence of antibodies has been reported as long as
18 years after initial treatment without clinical evidence of disease in patients who
had coexistent Hashimoto’s disease in their thyroidectomy specimen (118). It is
difficult to generalize from the results of either Tg or antithyroglobulin antibody
testing in one center because of interlaboratory variations in assay sensitivity
and specificity. With increasing assay sensitivity, Tg levels at least 1 to 2 ng/mL
following TSH stimulation should justify further diagnostic evaluation for possible
residual or metastatic disease (119). Tg levels that are minimally detectable during
TSH suppressive therapy but that do not rise following TSH stimulation may
suggest the presence of heterophilic antibodies causing false elevations of the
assay (81).

C CELLS, CALCITONIN, AND MEDULLARY THYROID CARCINOMA


The 32–amino acid peptide calcitonin is principally produced by the thyroid C
cells; smaller amounts can be found in neuroendocrine cells throughout the body.
Physiological stimuli to calcitonin secretion from the thyroid include hypercal-
cemia, hypergastrinemia, hypermagnesemia, and beta-adrenergic agonists. Sup-
pression of hormone release is produced by dopamine, somatostatin, and 1, 25-
dihydroxyvitamin D. In pharmacological amounts, calcitonin induces inhibition
of osteoclast-mediated bone resorption, leading to hypocalcemia and hypophos-
phatemia. However, in humans, neither an excess amount nor the absence of
calcitonin leads to demonstrable alterations in calcium or bone metabolism. Other
pharmacological effects of calcitonin include a diminution of pancreatic exocrine
and endocrine secretion and impaired renal tubular reabsorption of calcium and
phosphate.
The Laboratory Approach to Thyroid Disorders 29

Single- or double-antibody immunoassay techniques are used for the routine


measurement of serum calcitonin levels, although the latter are markedly more sen-
sitive. Difficulties in the interpretation of these assay results arise from calcitonin
species that are immunologically detectable but biologically inactive, such as poly-
meric forms of the molecule. Passage of the serum through a silica cartridge prior
to immunoassay yields mostly monomeric calcitonin, with a normal concentration
of ⬍10 pg/mL; calcitonin levels can be undetectable in more than half of normal
subjects tested (120). Additional uncertainty can be due to the numerous sites
of calcitonin synthesis, including pulmonary neuroendocrine cells, the pancreas,
and the small intestine. Thus, thyroidectomized patients may still have detectable
levels of calcitonin, and mild elevations occur in diseases that do not involve the
C cells of the thyroid. The differential diagnosis of hypercalcitoninemia includes
chronic and acute pulmonary diseases (such as pneumonitis, chronic obstructive
pulmonary disease, and sarcoidosis), gastrointestinal disorders (e.g., pancreatitis,
ileitis, and hypergastrinemic diseases), renal failure (due to impaired clearance),
cigarette smoking, and heterophilic antibodies (120,121). In addition to tumors of
neuroendocrine derivation (e.g., small-cell lung cancer, carcinoid, and pheochro-
mocytoma), many malignancies are associated with hypercalcitoninemia, although
frequently the calcitonins produced by these cancers are biologically inactive. In
contrast, the “hook effect” can also affect calcitonin immunoassays, causing false
low assay results despite marked elevation of calcitonin concentrations (122).
Medullary thyroid carcinoma (MTC) is a malignancy of the C cells that
produces marked hypercalcitoninemia. The disease can occur in several clinical
settings: sporadic tumors, multiple endocrine neoplasia syndromes (MEN), and
familial non-MEN medullary carcinoma. In the MEN syndromes, the disease
is bilateral and associated with primary hyperparathyroidism and pheochro-
mocytoma (MEN 2A), or mucosal ganglioneuromata, marfanoid habitus, and
pheochromocytoma (MEN 2B). Elevated serum levels of calcitonin (⬎200 pg/mL)
are almost always found with palpable medullary carcinomas, and the degree of
elevation usually corresponds to the size of the tumor. However, small tumors and
MEN patients with the premalignant lesion C-cell hyperplasia may have normal
basal levels of calcitonin, thus necessitating stimulation testing. The procedure
of greatest discriminant value in this setting is combined calcium-pentagastrin
testing, but pentagastrin is no longer available in the United States. To use
calcium as a sole stimulatory agent, elemental calcium (2 mg/kg) is administered
intravenously in 50 mL of 0.9% saline over 60 seconds, and plasma for calcitonin
measurements is collected 5 and 10 minutes after the infusion (123). An alternative
provocative test relies upon the calcitonin-stimulatory effects of omeprazole-
induced hypergastrinemia (124). C-cell hyperplasia and medullary carcinoma
produce greater than fivefold elevations in the serum calcitonin concentration.
Additional tumor markers, such as carcinoembryonic antigen, can be useful in the
long-term follow-up of patients with MTC. During follow-up, most patients whose
calcitonin levels are doubling in less than 6 months die within 5 years; doubling
time less than 24 months is associated with progressive disease (125,126).
30 Sherman

Routine measurement of the serum calcitonin concentration is not recom-


mended as a screen for MTC in a patient with a solitary nodule. However, recent
reports suggest that perhaps as many as 3% of patients with nodular thyroid dis-
ease will have an elevated serum calcitonin level when measured in a sensitive
immunometric assay, many of whom will prove to have MTC at thyroidectomy
(127). The cost-effectiveness of adding a single serum calcitonin determination
to the evaluation of a patient with nodular thyroid disease has been estimated as
nearly $12,000 per life-year saved, and is sensitive to variation in disease preva-
lence, specificity of fine-needle aspiration, and cost of testing (128). Nonetheless,
routine calcitonin measurements have not been recommended in the guidelines
for thyroid nodule evaluation by the American Thyroid Association (129).
For patients in known kindreds with inherited MTC, prospective family
screening increasingly identifies disease carriers long before clinical symptoms or
signs are noted. Using the traditional approach of stimulated secretion of calcitonin
by either pentagastrin or calcium infusion, 65% of MEN 2A gene carriers will have
abnormal calcitonin levels by age 20 years and 95% by age 35 (130). Compared
with sporadic disease, the typical age of presentation for familial disease is the third
decade, without gender preference. In MEN 2A, it is uncommon for the signs or
symptoms of hyperparathyroidism or pheochromocytoma to present before those
of MTC, even in the absence of prospective screening. All familial forms of
MTC and MEN 2 are inherited in an autosomal dominant fashion. In almost all
known kindreds, the disease is caused by a germline mutation in the ret proto-
oncogene, a 21-exon gene located near the centromere on chromosome 10. Ret
codes for a cell membrane–associated tyrosine kinase receptor for glial cell line-
derived neurotrophic factor, a circulating ligand that promotes development of
various central and peripheral nervous system neurons. Mutations associated with
MEN 2A and familial medullary thyroid carcinoma (FMTC) have been primarily
identified in several codons of the cysteine-rich extracellular domains of exon 8,
10, 11, and 13, whereas MEN 2B and some FMTC mutations are found within the
intracellular exons 15 and 16 (131).
Of note, nearly half of patients with sporadic disease have somatic ret
mutations, typically involving the codon 918 mutation that also causes MEN 2B
when inherited. Further, approximately 6% of patients with clinically sporadic
MTC carry a germline mutation in ret, leading to identification of new kindreds
with multiple previously undiagnosed affected individuals. Genetic testing for
ret proto-oncogene mutations should be offered to all patients newly diagnosed
with clinically apparent sporadic MTC, as well as for screening children and
adults in known kindreds with inherited forms of MTC. Given the frequency of
mutations in certain exons, a sensible strategy for mutational analysis would start
with examination of exon 11, followed sequentially by exons 10, 16, 13, 14, 15,
and 8 (132). Although common mutations can be identified by broadly available
commercial testing sources, only a limited number of sites perform the more
thorough analyses that are required to identify the less common mutations.
Studies of genotype:phenotype correlations of various ret mutations have
demonstrated the presence of three levels of risk, defined as age of earliest
The Laboratory Approach to Thyroid Disorders 31

diagnosis of malignant disease and risk for mortality due to the cancer (131). At the
highest risk are the group 3 mutations, codons 918 and 883 that cause MEN 2B and
metastatic disease as early as age 6 months. At higher risk are the group 2 muta-
tions, which include the common mutations at codons 611, 618, 620, and 634 that
are associated with malignant disease as early as age 5. At high risk are the group
1 mutations, which include those at codons 609, 768, 790, 791, 804, and 891; with
some of the group 1 mutations, fatal MTC has yet to be reported. Recommendations
for the aggressiveness and timing of therapeutic intervention are tied to these risk
levels (132).

REFERENCES

1. Burmeister LA, Goumaz MO, Mariash CN, et al. Levothyroxine dose requirements
for thyrotropin suppression in the treatment of differentiated thyroid cancer. J Clin
Endocrinol Metab 1992; 75(2):344–350.
2. Spencer CA, LoPresti JS, Patel A, et al. Applications of a new chemiluminomet-
ric thyrotropin assay to subnormal measurement. J Clin Endocrinol Metab 1990;
70(2):453–460.
3. Meikle AW, Stringham JD, Woodward MG, et al. Hereditary and environmental influ-
ences on the variation of thyroid hormones in normal male twins. J Clin Endocrinol
Metab 1988; 66(3):588–592.
4. Hansen PS, Brix TH, Iachine I, et al. Genetic and environmental interrelations
between measurements of thyroid function in a healthy Danish twin population. Am
J Physiol Endocrinol Metab 2007; 292(3):E765–E770.
5. Jensen E, Hyltoft Petersen P, Blaabjerg O, et al. Establishment of a serum thyroid
stimulating hormone (TSH) reference interval in healthy adults. The importance
of environmental factors, including thyroid antibodies. Clin Chem Lab Med 2004;
42(7):824–832.
6. Midgley JE. Direct and indirect free thyroxine assay methods: Theory and practice.
Clin Chem 2001; 47(8):1353–1363.
7. Baloch Z, Carayon P, Conte-Devolx B, et al. Laboratory medicine practice guidelines.
Laboratory support for the diagnosis and monitoring of thyroid disease. Thyroid
2003; 13(1):3–126.
8. Steele BW, Wang E, Klee GG, et al. Analytic bias of thyroid function tests: Analysis
of a College of American Pathologists fresh frozen serum pool by 3900 clinical
laboratories. Arch Pathol Lab Med 2005; 129(3):310–317.
9. Thienpont LM, Beastall G, Christofides ND, et al. Proposal of a candidate inter-
national conventional reference measurement procedure for free thyroxine in
serum 1): International Federation of Clinical Chemistry and Laboratory Medicine
(IFCC) 2) IFCC Scientific Division. Clin Chem Lab Med 2007; 45(7):934–
936.
10. Holm SS, Hansen SH, Faber J, et al. Reference methods for the measurement of
free thyroid hormones in blood: Evaluation of potential reference methods for free
thyroxine. Clin Biochem 2004; 37(2):85–93.
11. Yue B, Rockwood AL, Sandrock T, et al. Free thyroid hormones in serum by
direct equilibrium dialysis and online solid-phase extraction-liquid chromatogra-
phy/tandem mass spectrometry. Clin Chem 2008; 54(4):642–651.
32 Sherman

12. Wang R, Nelson JC, Weiss RM, et al. Accuracy of free thyroxine measurements
across natural ranges of thyroxine binding to serum proteins. Thyroid 2000; 10:31–
39.
13. Toldy E, Locsei Z, Szabolcs I, et al. Protein interference in thyroid assays: An in
vitro study with in vivo consequences. Clin Chim Acta 2005; 352(1–2):93–104.
14. Gu J, Soldin OP, Soldin SJ. Simultaneous quantification of free triiodothyronine and
free thyroxine by isotope dilution tandem mass spectrometry. Clin Biochem 2007;
40(18):1386–1391.
15. Woeber KA. Triiodothyronine production in Graves’ hyperthyroidism. Thyroid
2006; 16(7):687–690.
16. Jonklaas J, Davidson B, Bhagat S, et al. Triiodothyronine levels in athyreotic indi-
viduals during levothyroxine therapy. JAMA 2008; 299(7):769–777.
17. Ceresini G, Morganti S, Rebecchi I, et al. A one-year follow-up on the effects
of raloxifene on thyroid function in postmenopausal women. Menopause 2004;
11(2):176–179.
18. Brent GA. Maternal thyroid function: Interpretation of thyroid function tests in
pregnancy. Clin Obstet Gynecol 1997; 40(1):3–15.
19. Shifren JL, Desindes S, McIlwain M, et al. A randomized, open-label, crossover study
comparing the effects of oral versus transdermal estrogen therapy on serum andro-
gens, thyroid hormones, and adrenal hormones in naturally menopausal women.
Menopause 2007; 14(6):985–994.
20. Fisher CL, Mannino DM, Herman WH, et al. Cigarette smoking and thyroid hormone
levels in males. Int J Epidemiol 1997; 26(5):972–977.
21. Stevenson HP, Archbold GP, Johnston P, et al. Misleading serum free thyroxine
results during low molecular weight heparin treatment. Clin Chem 1998; 44(5):1002–
1007.
22. Klee GG. Interferences in hormone immunoassays. Clin Lab Med 2004; 24(1):1–18.
23. Norden AG, Jackson RA, Norden LE, et al. Misleading results from immunoassays
of serum free thyroxine in the presence of rheumatoid factor. Clin Chem 1997; 43(6
Pt 1):957–962.
24. Martino E, Bartalena L, Bogazzi F, et al. The effects of amiodarone on the thyroid.
Endocr Rev 2001; 22(2):240–254.
25. Dumitrescu AM, Liao XH, Abdullah MSY, et al. Mutations in SECISBP2 result in
abnormal thyroid hormone metabolism. Nat Genet 2005; 37:1247–1252.
26. Refetoff S, Dumitrescu AM. Syndromes of reduced sensitivity to thyroid hormone:
Genetic defects in hormone receptors, cell transporters and deiodination. Best Pract
Res Clin Endocrinol Metab 2007; 21(2):277–305.
27. Attia J, Margetts P, Guyatt G. Diagnosis of thyroid disease in hospitalized patients:
A systematic review. Arch Intern Med 1999; 159(7):658–665.
28. Chopra IJ, Solomon DH, Huang TS. Serum thyrotropin in hospitalized psychiatric
patients: Evidence for hyperthyrotropinemia as measured by an ultrasensitive thy-
rotropin assay. Metabolism 1990; 39(5):538–543.
29. Nader S, Warner MD, Doyle S, et al. Euthyroid sick syndrome in psychiatric inpa-
tients. Biol Psychiatry 1996; 40(12):1288–1293.
30. Woolf PD, Nichols D, Porsteinsson A, et al. Thyroid evaluation of hospitalized
psychiatric patients: The role of TSH screening for thyroid dysfunction. Thyroid
1996; 6(5):451–456.
The Laboratory Approach to Thyroid Disorders 33

31. Brouwer JP, Appelhof BC, Hoogendijk WJG, et al. Thyroid and adrenal axis in major
depression: A controlled study in outpatients. Eur J Endocrinol 2005; 152(2):185–
191.
32. Olff M, Güzelcan Y, de Vries G-J, et al. HPA- and HPT-axis alterations in chronic
posttraumatic stress disorder. Psychoneuroendocrinology 2006; 31(10):1220–1230.
33. Schussler GC. The thyroxine-binding proteins. Thyroid 2000; 10(2):141–149.
34. Langton JE, Brent GA. Nonthyroidal illness syndrome: Evaluation of thyroid func-
tion in sick patients. Endocrinol Metab Clin North Am 2002; 31(1):159–172.
35. Adler SM, Wartofsky L. The nonthyroidal illness syndrome. Endocrinol Metab Clin
North Am 2007; 36(3):657–672.
36. Feelders RA, Swaak AJ, Romijn JA, et al. Characteristics of recovery from the
euthyroid sick syndrome induced by tumor necrosis factor alpha in cancer patients.
Metabolism 1999; 48(3):324–329.
37. Harel Z, Biro FM, Tedford WL. Effects of long term treatment with sertraline (Zoloft)
simulating hypothyroidism in an adolescent. J Adolesc Health 1995; 16(3):232–234.
38. Samuels MH, Pillote K, Asher D, et al. Variable effects of nonsteroidal antiinflamma-
tory agents on thyroid test results. J Clin Endocrinol Metab 2003; 88(12):5710–5716.
39. Torpy DJ, Tsigos C, Lotsikas AJ, et al. Acute and delayed effects of a single-dose
injection of interleukin-6 on thyroid function in healthy humans. Metabolism 1998;
47(10):1289–1293.
40. Sherman SI, Schlumberger MJ, Elisei R, et al. Exacerbation of postsurgical hypothy-
roidism during treatment of thyroid carcinoma with motesanib diphosphate (AMG
706). In: 89th Annual Meeting of The Endocrine Society; 2007 June 4; Toronto, ON;
2007.
41. Tamaskar I, Bukowski R, Elson P, et al. Thyroid function test abnormalities in
patients with metastatic renal cell carcinoma treated with sorafenib. Ann Oncol
2008; 19(2):265–268.
42. Fliers E, Alkemade A, Wiersinga WM. The hypothalamic-pituitary-thyroid axis in
critical illness. Best Pract Res Clin Endocrinol Metab 2001; 15(4):453–464.
43. Magner J, Roy P, Fainter L, et al. Transiently decreased sialylation of thyrotropin
(TSH) in a patient with the euthyroid sick syndrome. Thyroid 1997; 7(5):807–808.
44. DeGroot LJ. Non-thyroidal illness syndrome is functional central hypothyroidism,
and if severe, hormone replacement is appropriate in light of present knowledge. J
Endocrinol Invest 2003; 26(12):1163–1170.
45. Spencer CA, Schwarzbein D, Guttler RB, et al. Thyrotropin (TSH)-releasing hor-
mone stimulation test responses employing third and fourth generation TSH assays.
J Clin Endocrinol Metab 1993; 76:494–498.
46. Andersen S, Bruun NH, Pedersen KM, et al. Biologic variation is important for
interpretation of thyroid function tests. Thyroid 2003; 13(11):1069–1078.
47. Persani L, Borgato S, Romoli R, et al. Changes in the degree of sialylation of
carbohydrate chains modify the biological properties of circulating thyrotropin iso-
forms in various physiological and pathological states. J Clin Endo Metab 1998;
83(7):2486–2492.
48. Donadio S, Pascual A, Thijssen JHH, et al. Feasibility study of new calibrators
for thyroid-stimulating hormone (TSH) immunoprocedures based on remodeling of
recombinant TSH to mimic glycoforms circulating in patients with thyroid disorders.
Clin Chem 2006; 52(2):286–297.
34 Sherman

49. Harris EK. Effects of intra- and interindividual variation on the appropriate use of
normal ranges. Clin Chem 1974; 20(12):1535–1542.
50. Brabant G, Beck-Peccoz P, Jarzab B, et al. Is there a need to redefine the upper
normal limit of TSH? Eur J Endocrinol 2006; 154(5):633–637.
51. Hollowell JG, Staehling NW, Flanders WD, et al. Serum TSH, T(4), and thyroid anti-
bodies in the United States population (1988 to 1994): National Health and Nutrition
Examination Survey (NHANES III). J Clin Endocrinol Metab 2002; 87(2):489–499.
52. Jensen E, Blaabjerg O, Petersen PH, et al. Sampling time is important but may
be overlooked in establishment and use of thyroid-stimulating hormone reference
intervals. Clin Chem 2007; 53(2):355–356.
53. Spencer CA, Hollowell JG, Kazarosyan M, et al. National Health and Nutrition
Examination Survey III thyroid-stimulating hormone (TSH)-thyroperoxidase anti-
body relationships demonstrate that TSH upper reference limits may be skewed by
occult thyroid dysfunction. J Clin Endocrinol Metab 2007; 92(11):4236–4240.
54. Vanderpump MP, Tunbridge WM, French JM, et al. The incidence of thyroid dis-
orders in the community: A twenty-year follow-up of the Whickham Survey. Clin
Endocrinol (Oxf) 1995; 43(1):55–68.
55. AACE. American Association of Clinical Endocrinologists medical guidelines for
clinical practice for the evaluation and treatment of hyperthyroidism and hypothy-
roidism. Endocr Pract 2002; 8(6):457–469.
56. Wartofsky L, Dickey RA. The evidence for a narrower thyrotropin reference range
is compelling. J Clin Endocrinol Metab 2005; 90(9):5483–5488.
57. Surks MI, Hollowell JG. Age-specific distribution of serum thyrotropin and antithy-
roid antibodies in the US population: Implications for the prevalence of subclinical
hypothyroidism. J Clin Endocrinol Metab 2007; 92(12):4575–4582.
58. Biondi B, Cooper DS. The clinical significance of subclinical thyroid dysfunction.
Endocr Rev 2008; 29(1):76–131.
59. Despres N, Grant AM. Antibody interference in thyroid assays: A potential for
clinical misinformation. Clin Chem 1998; 44(3):440–454.
60. Oliveira JHA, Persani L, Beck-Peccoz P, et al. Investigating the paradox of hypothy-
roidism and increased serum thyrotropin (TSH) levels in Sheehan’s syndrome: Char-
acterization of TSH carbohydrate content and bioactivity. J Clin Endo Metab 2001;
86(4):1694–1699.
61. Persani L, Ferretti E, Borgato S, et al. Circulating thyrotropin bioactivity in sporadic
central hypothyroidism. J Clin Endo Metab 2000; 85(10):3631–3635.
62. Sherman SI, Gopal J, Haugen BR, et al. Central hypothyroidism associated with
retinoid X receptor-selective ligands. N Engl J Med 1999; 340(14):1075–1079.
63. Golden WM, Weber KB, Hernandez TL, et al. Single-dose rexinoid rapidly and
specifically suppresses serum thyrotropin in normal subjects. J Clin Endocrinol
Metab 2007; 92(1):124–130.
64. Vigersky RA, Filmore-Nassar A, Glass AR. Thyrotropin suppression by metformin.
J Clin Endocrinol Metab 2006; 91(1):225–227.
65. LinksIsidro ML, Penı́n MA, Nemiña R, et al. Metformin reduces thyrotropin levels
in obese, diabetic women with primary hypothyroidism on thyroxine replacement
therapy. Endocrine 2007; 32(1):79–82.
66. Zidan J, Rubenstein W. Effect of adjuvant tamoxifen therapy on thyroid function in
postmenopausal women with breast cancer. Oncology 1999; 56(1):43–45.
The Laboratory Approach to Thyroid Disorders 35

67. Chan LY-S, Chiu PY, Lau TK. Cord blood thyroid-stimulating hormone level in
high-risk pregnancies. Eur J Obstet Gynecol Reprod Biol 2003; 108(2):142–145.
68. Beck-Peccoz P, Persani L, Calebiro D, et al. Syndromes of hormone resistance in the
hypothalamic-pituitary-thyroid axis. Best Pract Res Clin Endocrinol Metab 2006;
20(4):529–546.
69. Fisher DA, Nelson JC, Carlton EI, et al. Maturation of human hypothalamic-
pituitary-thyroid function and control. Thyroid 2000; 10(3):229–234.
70. Bertelsen JB, Hegedus L. Cigarette smoking and the thyroid. Thyroid 1994;
4(3):327–331.
71. Spencer CA, Lopresti JS. Measuring thyroglobulin and thyroglobulin autoantibody
in patients with differentiated thyroid cancer. Nat Clin Pract Endocrinol Metab 2008;
4(4):223–233.
72. Baloch ZW, Barroeta JE, Walsh J, et al. Utility of thyroglobulin measurement in
fine-needle aspiration biopsy specimens of lymph nodes in the diagnosis of recurrent
thyroid carcinoma. Cytojournal 2008; 5:1–5.
73. Spencer CA, Takeuchi M, Kazarosyan M, et al. Serum thyroglobulin autoantibodies:
Prevalence, influence on serum thyroglobulin measurement, and prognostic signif-
icance in patients with differentiated thyroid carcinoma. J Clin Endocrinol Metab
1998; 83(4):1121–1127.
74. Boi F, Baghino G, Atzeni F, et al. The diagnostic value for differentiated thyroid
carcinoma metastases of thyroglobulin (Tg) measurement in washout fluid from
fine-needle aspiration biopsy of neck lymph nodes is maintained in the presence of
circulating anti-Tg antibodies. J Clin Endocrinol Metab 2006; 91(4):1364–1369.
75. Prentice L, Kiso Y, Fukuma N, et al. Monoclonal thyroglobulin autoantibodies:
Variable region analysis and epitope recognition. J Clin Endocrinol Metab 1995;
80(3):977–986.
76. Kloos RT, Mazzaferri EL. A single recombinant human thyrotropin-stimulated serum
thyroglobulin measurement predicts differentiated thyroid carcinoma metastases
three to five years later. J Clin Endocrinol Metab 2005; 90(9):5047–5057.
77. Schroeder PR, Haugen BR, Pacini F, et al. A comparison of short-term changes in
health-related quality of life in thyroid carcinoma patients undergoing diagnostic
evaluation with recombinant human thyrotropin compared with thyroid hormone
withdrawal. J Clin Endocrinol Metab 2006; 91(3):878–884.
78. Robbins RJ, Srivastava S, Shaha A, et al. Factors influencing the basal and recombi-
nant human thyrotropin-stimulated serum thyroglobulin in patients with metastatic
thyroid carcinoma. J Clin Endocrinol Metab 2004; 89(12):6010–6016.
79. Torlontano M, Crocetti U, Augello G, et al. Comparative evaluation of recombinant
human thyrotropin-stimulated thyroglobulin levels, 131I whole-body scintigraphy,
and neck ultrasonography in the follow-up of patients with papillary thyroid micro-
carcinoma who have not undergone radioiodine therapy. J Clin Endocrinol Metab
2006; 91(1):60–63.
80. Giovanella L, Ghelfo A. Undetectable serum thyroglobulin due to negative inter-
ference of heterophile antibodies in relapsing thyroid carcinoma. Clin Chem 2007;
53(10):1871–1872.
81. Preissner CM, O’Kane DJ, Singh RJ, et al. Phantoms in the assay tube: Heterophile
antibody interferences in serum thyroglobulin assays. J Clin Endocrinol Metab 2003;
88(7):3069–3074.
36 Sherman

82. Mazzaferri EL. Will highly sensitive thyroglobulin assays change the management
of thyroid cancer? Clin Endocrinol (Oxf) 2007; 67(3):321–323.
83. Schlumberger M, Hitzel A, Toubert ME, et al. Comparison of seven serum thy-
roglobulin assays in the follow-up of papillary and follicular thyroid cancer patients.
J Clin Endocrinol Metab 2007; 92(7):2487–2495.
84. Sinclair D. Analytical aspects of thyroid antibodies estimation. Autoimmunity 2008;
41(1):46–54.
85. Dherbomez M, Sapin R, Gasser F, et al. Concordance of eight kits for antithyroid
peroxidase autoantibodies determination. Clin Chem Lab Med 2000; 38(6):561–566.
86. Spencer CA, Takeuchi M, Kazarosyan M, et al. Serum thyroglobulin autoantibodies:
Prevalence, influence on serum thyroglobulin measurement, and prognostic signif-
icance in patients with differentiated thyroid carcinoma. J Clin Endocrinol Metab
1998; 83(4):1121–1127.
87. Morgenthaler NG. New assay systems for thyrotropin receptor antibodies. Curr Opin
Endocrinol Diabetes Metab 1999; 6(4):251–260.
88. Schott M, Feldkamp J, Bathan C, et al. Detecting TSH-receptor antibodies with the
recombinant TBII assay: Technical and clinical evaluation. Horm Metab Res 2000;
32(10):429–435.
89. Costagliola S, Morgenthaler NG, Hoermann R, et al. Second generation assay for
thyrotropin receptor antibodies has superior diagnostic sensitivity for Graves’ dis-
ease. J Clin Endo Metab 1999; 84(1):90–97.
90. Gupta MK. Thyrotropin-receptor antibodies in thyroid diseases: Advances in detec-
tion techniques and clinical applications. Clin Chem Acta 2000; 293(1–2):1–29.
91. Yamano Y, Takamatsu J, Sakane S, et al. Differences between changes in serum
thyrotropin-binding inhibitory antibodies and thyroid-stimulating antibodies in the
course of antithyroid drug therapy for Graves’ disease. Thyroid 1999; 9(8):769–773.
92. Carella C, Mazziotti G, Sorvillo F, et al. Serum thyrotropin receptor antibodies
concentrations in patients with Graves’ disease before, at the end of methimazole
treatment, and after drug withdrawal: Evidence that the activity of thyrotropin recep-
tor antibody and/or thyroid response modify during the observation period. Thyroid
2006; 16(3):295–302.
93. Schott M, Morgenthaler NG, Fritzen R, et al. Levels of autoantibodies against human
TSH receptor predict relapse of hyperthyroidism in Graves’ disease. Horm Metab
Res 2004; 36(2):92–96.
94. Giovanella L, Ceriani L, Garancini S. Clinical applications of the 2nd generation
assay for anti-TSH receptor antibodies in Graves’ disease: Evaluation in patients
with negative 1st generation test. Clin Chem Lab Med 2001; 39(1):25–28.
95. Stagnaro-Green A, Roman SH, Cobin RH, et al. Detection of at-risk pregnancy
by means of highly sensitive assays for thyroid autoantibodies. JAMA 1990;
264(11):1422–1425.
96. Burmeister LA. Reverse T3 does not reliably differentiate hypothyroid sick syndrome
from euthyroid sick syndrome. Thyroid 1995; 5(6):435–441.
97. Klein I. Clinical, metabolic, and organ-specific indices of thyroid function.
Endocrinol Metab Clin North Am 2001; 30(2):415–427.
98. Refetoff S. Resistance to thyroid hormone. Clin Lab Med 1993; 13(3):563–581.
99. Surks MI, Ortiz E, Daniels GH,et al. Subclinical thyroid disease: Scientific review
and guidelines for diagnosis and management. JAMA 2004; 291(2):228–238.
The Laboratory Approach to Thyroid Disorders 37

100. Helfand M. Screening for subclinical thyroid dysfunction in nonpregnant adults: A


summary of the evidence for the U. S. Preventive Services Task Force. Ann Intern
Med 2004; 140(2):128–141.
101. Danese MD, Powe NR, Sawin CT, et al. Screening for mild thyroid failure at the
periodic health examination. A decision and cost-effectiveness analysis. JAMA 1996;
276(4):285–292.
102. Gharib H, Tuttle RM, Baskin HJ,et al. Subclinical thyroid dysfunction: A joint state-
ment on management from the American Association of Clinical Endocrinologists,
the American Thyroid Association, and the Endocrine Society. J Clin Endocrinol
Metab 2005; 90(1):581–585.
103. Stone MB, Wallace RB. Committee on the Medicare Coverage of Routine Thy-
roid Screening BoHCS. Institute of Medicine. Medicare Coverage of Routine
Screening for Thyroid Dysfunction. Washington, DC: National Academies Press,
2003.
104. Vanderpump MPJ, Ahlquist JAO, Franklyn JA, et al. Consensus statement for good
practice and audit measures in the management of hypothyroidism and hyperthy-
roidism. Bmj 1996; 313(7056):539–544.
105. U. S. Preventive Services Task Force. Screening for thyroid disease: Recommenda-
tion statement. Ann Intern Med 2004; 140(2):125–127.
106. Screening for thyroid disease. Ann Intern Med 1998; 129(2):141–143.
107. LaFranchi S, Dussault JH, Fisher DA, et al. Newborn screening for congenital
hypothyroidism: Recommended guidelines. Pediatrics 1993; 152:974–975.
108. Hanna CE, Krainz PL, Skeels MR, et al. Detection of congenital hypopituitary
hypothyroidism: Ten-year experience in the northwest regional screening program.
J Pediatr 1986; 109(6):959–964.
109. Delange F. Screening for congenital hypothyroidism used as an indicator of the
degree of iodine deficiency and of its control. Thyroid 1998; 8(12):1185–1192.
110. Abalovich M, Amino N, Barbour LA, et al. Management of thyroid dysfunction
during pregnancy and postpartum: An Endocrine Society Clinical Practice Guideline.
J Clin Endocrinol Metab 2007; 92(8 suppl):s1–s47.
111. Schectman JM, Kallenberg GA, Shumacher RJ, et al. Yield of hypothyroidism in
symptomatic primary care patients. Arch Intern Med 1989; 149:861–864.
112. Wardle CA, Fraser WD, Squire CR. Pitfalls in the use of thyrotropin concentration
as a first-line thyroid-function test. Lancet 2001; 357(9261):1013–1014.
113. McDermott MT, Woodmansee WW, Haugen BR, et al. The management of subclin-
ical hyperthyroidism by thyroid specialists. Thyroid 2003; 13(12):1133–1139.
114. Viswanath AK, Avenell A, Philip S, et al. Is annual surveillance of all treated
hypothyroid patients necessary? BMC Endocr Disord 2007; 7:4.
115. Ferretti E, Persani L, Jaffrain-Rea ML, et al. Evaluation of the adequacy of
levothyroxine replacement therapy in patients with central hypothyroidism. J Clin
Endocrinol Metab 1999; 84(3):924–929.
116. Saravanan P, Siddique H, Simmons DJ, et al. Twenty-four hour hormone profiles of
TSH, Free T3 and free T4 in hypothyroid patients on combined T3/T4 therapy. Exp
Clin Endocrinol Diabetes 2007; 115(4):261–267.
117. Haugen BR, Pacini F, Reiners C, et al. A comparison of recombinant human thy-
rotropin and thyroid hormone withdrawal for the detection of thyroid remnant or
cancer. J Clin Endocrinol Metab 1999; 84(11):3877–3885.
38 Sherman

118. Chiovato L, Latrofa F, Braverman LE, et al. Disappearance of humoral thyroid


autoimmunity after complete removal of thyroid antigens. Ann Intern Med 2003;
139(5 Pt 1):346–351.
119. Mazzaferri EL, Robbins RJ, Spencer CA, et al. A consensus report of the role of
serum thyroglobulin as a monitoring method for low-risk patients with papillary
thyroid carcinoma. J Clin Endocrinol Metab 2003; 88(4):1433–1441.
120. d’Herbomez M, Caron P, Bauters C, et al. Reference range of serum calcitonin
levels in humans: Influence of calcitonin assays, sex, age, and cigarette smoking.
Eur J Endocrinol 2007; 157(6):749–755.
121. Papapetrou PD, Polymeris A, Karga H, et al. Heterophilic antibodies causing falsely
high serum calcitonin values. J Endocrinol Invest 2006; 29(10):919–923.
122. Tommasi M, Raspanti S. Hook effect in calcitonin immunoradiometric assay. Clin
Chem Lab Med 2007; 45(8):1073–1074.
123. Parthemore JG, Bronzert D, Roberts G, et al. A short calcium infusion in the diagnosis
of medullary thyroid carcinoma. J Clin Endocrinol Metab 1974; 39(1):108–111.
124. Vitale G, Ciccarelli A, Caraglia M, et al. Comparison of two provocative tests for
calcitonin in medullary thyroid carcinoma: Omeprazole vs pentagastrin. Clin Chem
2002; 48(9):1505–1510.
125. Barbet J, Campion L, Kraeber-Bodere F, et al. Prognostic impact of serum calcitonin
and carcinoembryonic antigen doubling-times in patients with medullary thyroid
carcinoma. J Clin Endo Metab 2005; 90(11):6077–6084.
126. Laure Giraudet A, Al Ghulzan A, Auperin A, et al. Progression of medullary thy-
roid carcinoma: Assessment with calcitonin and carcinoembryonic antigen doubling
times. Eur J Endocrinol 2008; 158(2):239–246.
127. Elisei R, Bottici V, Luchetti F, et al. Impact of routine measurement of serum
calcitonin on the diagnosis and outcome of medullary thyroid cancer: Experience
in 10,864 patients with nodular thyroid disorders. J Clin Endocrinol Metab 2004;
89(1):163–168.
128. Cheung K, Roman SA, Wang TS, et al. Calcitonin measurement in the evaluation
of thyroid nodules in the United States: A cost-effectiveness and decision analysis.
J Clin Endocrinol Metab 2008; 93:2173–2180.
129. Cooper DS, Doherty GM, Haugen BR, et al. Management guidelines for patients with
thyroid nodules and differentiated thyroid cancer. Thyroid 2006; 16(2):109–142.
130. Ponder BA, Ponder MA, Coffey R, et al. Risk estimation and screening in families
of patients with medullary thyroid carcinoma. Lancet 1988; 1(8582):397–401.
131. Evans DB, Shapiro SE, Cote GJ. Invited commentary: Medullary thyroid cancer:
The importance of RET testing. Surgery 2007; 141(1):96–99.
132. Sherman SI, Angelos P, Ball DW, et al. Thyroid carcinoma. J Natl Compr Canc Netw
2007; 5(6):568–621.
2
Hyperthyroidism Due to Graves’ Disease,
Toxic Nodules and Toxic Multinodular
Goiter

Kenneth D. Burman
Washington Hospital Center, Georgetown University, and the Uniformed
Services of the Health Sciences, Washington, D.C., U.S.A.

David S. Cooper
The Johns Hopkins University School of Medicine, Baltimore,
Maryland, U.S.A.

GRAVES’ DISEASE
Introduction
Graves’ disease is an autoimmune thyroid disorder characterized by clinical hyper-
thyroidism and the presence of autoantibodies directed against the thyrotropin
(TSH) receptor (1,2). The presentation of this disease varies with age. Younger
patients manifest nervousness, weight loss, anxiety, heat intolerance, hyperdefeca-
tion, inability to concentrate, and tremulousness, while older patients may manifest
few if any of these typical symptoms (3). Circulating TSH receptor–stimulating
antibodies are present in at least 90% of patients and are responsible, in large
part, for the thyroidal hyperactivity (4). An interesting aspect of Graves’ dis-
ease is its association with ophthalmopathy, which can cause tearing, burning,
itching, proptosis, double vision, and/or (rarely) visual impairment (5,6). The

39
40 Burman and Cooper

etiology of Graves’ hyperthyroidism and ophthalmopathy remains unclear. There


are abnormalities in T-cell function that allow the TSH receptor antibodies to
develop; these antibodies not only stimulate TSH receptor action in thyrocytes
but may cross-react with orbital antigens (e.g., fibroblasts) as well (5,7–13).
The term Graves’ disease commonly refers to patients who manifest clinical
and biochemical hyperthyroidism (1), but in certain circumstances this term may
require further delineation. For example, patients with Graves’ disease and hyper-
thyroidism may become spontaneously euthyroid or hypothyroid because, over
time, the stimulatory TSH receptor antibodies have become blocking antibod-
ies (7,10) or because the chronic inflammatory process has resulted in thyroid
failure.

Epidemiology
Although it may present in patients of any age, Graves’ disease occurs more
commonly in women than in men, especially in women between the ages of
about 20 and 50 years (3). Graves’ disease is rare in young children; when
it occurs in neonates, it is almost always related to transplacental passage of
TSH receptor–stimulating immunoglobulins, a condition that typically persists
for several weeks until the IgG antibodies are cleared from the neonate’s circu-
lation (14–17). In adults, the annual incidence of new cases of Graves’ disease
is 1 to 10 per 100,000, although, of course, these numbers vary depending upon
the method of detection and the iodine content in the geographic area (18–23).
People living in all areas of the world are affected with Graves’ disease. It is
believed that the incidence correlates directly with the amount of iodine in the
diet. Increased iodine intake has been shown to be associated with an increased
frequency of hyperthyroidism, possibly related to enhanced thyroid hormone syn-
thesis and possibly because iodine sufficiency is important for maximal antibody
production (19,20,24–32). Graves’ disease may present more commonly in the
spring and summer months, possibly because warmer weather increases the per-
ception of symptoms rather than a true increased incidence in disease activity.
Cigarette smoking (33) and stressful life events (20) have also been linked to
the etiology of Graves’ disease. The use of certain drugs, especially interferon-
alfa, has been associated with the development of Graves’ disease during
therapy (34).

Pathophysiology
Although much has been learned about the immune dysregulation that character-
izes Graves’ disease, the precise cause is unknown. There are defects in antigen-
specific T cells that result in B-cell production of many antibodies, most notably
stimulatory TSH receptor antibodies. The thyroid glands of patients with Graves’
disease are infiltrated with these antigen-specific T cells. Whether the disease is
caused by abnormal clones of autoreactive T cells or the initial trigger is abnormal
antigen presentation by thyrocytes is not known (35,36).
Hyperthyroidism 41

Table 1 Clinical Effects of Hyperthyroidism


System Effects

General Nervousness, insomnia, fatigue, tremulousness, heat intolerance,


weight loss
Skin Warmth, moistness, hyperidrosis, alopecia, increased
pigmentation, onycholysis, acropachy, pretibial myxedema,
urticaria, pruritus, vitiligo
Eyes Exophthalmos, conjunctivitis, chemosis, diplopia, decreased
vision
Cardiovascular Tachycardia, dyspnea, palpitations, atrial fibrillation, heart block,
congestive failure, angina pectoris
Gastrointestinal Hyperphagia, diarrhea or hyperdefecation, elevated liver
function tests, hepatosplenomegaly
Metabolic Elevated serum calcium, decreased serum magnesium, increased
bone alkaline phosphatase, hypercalciuria
Neuromuscular Fine hand tremor, proximal muscle weakness, myopathy, muscle
atrophy, creatinuria, periodic paralysis
Osseous Osteoporosis, osteopenia
Neurologic Fever, delirium, stupor, coma, syncope, choreaathetosis,
hemiballismus
Reproductive/sexual Irregular menses, amenorrhea, gynecomastia, decreased fertility
Hematopoietic Normochromic normocytic anemia, lymphocytosis,
lymphadenopathy, enlarged thymus, splenomegaly
Mental Restlessness, irritability anxiety, inability to concentrate,
emotional lability, depression, psychosis

Source: From Ref. 2. This table is not intended to be all-inclusive but rather representative.

Diagnosis
Signs and Symptoms
The typical signs and symptoms of Graves’ hyperthyroidism do not differ sig-
nificantly from those of any other type of hyperthyroidism (Table 1) (3). The
main features of hyperthyroidism relate to the action of excess thyroid hormone
at the cellular level and enhanced beta-adrenergic activity. Typical manifestations
include weakness, fatigue, anxiety, tremulousness, heat intolerance, and weight
loss. Any organ system may be involved. The skin may be warm, smooth, and
moist. Tachycardia is common, but atrial arrhythmias, heart block, or high or low
cardiac output may occur, especially in older individuals. Mitral valve prolapse, a
systolic flow murmur, an S3 gallop, or a Means-Lerman “scratch” murmur may be
present. The latter systolic sound is best heard along the left intercostal space dur-
ing expiration. It is thought to result from either turbulent pulmonic artery blood
flow or to friction between the pericardial and pleural surface in a hyperdynamic
heart. Recent studies have also shown that older patients with thyrotoxicosis may
develop congestive heart failure with evidence of a reversible cardiomyopathy
42 Burman and Cooper

and normal or low ejection fraction (37). In addition, some patients may develop
reversible, usually asymptomatic, pulmonary hypertension related to increased
cardiac output or to left atrial diastolic dysfunction (38).
Patients may complain that they are eating voraciously but are still losing
weight; hyperdefecation is more frequent than diarrhea. Liver function tests may
be elevated secondary to the hyperthyroid process or, less frequently, to related
autoimmune disorders, such as primary biliary cirrhosis, systemic lupus erythe-
matosus, or scleroderma. Additional manifestations of hyperthyroidism actually
relate to the underlying immunologic abnormalities. Vitiligo or prematurely gray
hair may be present, indicating the presence of antimelanocyte autoantibodies.
Patients with Graves’ ophthalmopathy may present with or have burning, itching,
proptosis, photophobia, or diplopia. Uncommonly, there may be proptosis or optic
nerve compression, resulting in decreased visual acuity (Figs. 1–4).
Elevated serum calcium, probably related to a direct effect of thyroid hor-
mones on osteoclasts, may occur in approximately 10% of patients. Some believe
that the likelihood of having a coincidental parathyroid adenoma is increased in
Graves’ disease patients (39). Patients who do have overt hyperthyroidism of any
type, especially if chronic, may have lower bone density values than they would
otherwise. Even mildly elevated thyroid hormone levels or subclinical hyperthy-
roidism may be associated with decreased bone mineral density, most notably in
postmenopausal women (40–44). The issue of whether Graves’ hyperthyroidism
causes sufficient bone loss to induce or aggravate an increased risk of fractures
is controversial (40–42,45–47). The risk of fractures is probably increased in

Figure 1 (A)Patient with Graves’ ophthalmopathy demonstrating proptosis and severe


exposure keratitis. This patient had received external radiation therapy, which, along with
inability to close the lid completely, contributed to dryness and resultant keratitis (8).
Hyperthyroidism 43

Figure 2 An orbital CT scan (coronal view) showing diffuse orbital muscle involvement
with enlargement, most notably of the right medial and left inferior recti muscles in a patient
with Graves’ ophthalmopathy (8).

postmenopausal women with a history of hyperthyroidism as shown in a recent


meta-analysis, although further studies are needed in this area (48,49).
Hand tremor and generalized proximal muscle weakness are common.
Rarely, hypokalemic periodic paralysis may occur, most frequently in Asian males
(50–52). Attacks are precipitated by high carbohydrate intake and heavy exercise.
In a recent analysis of thyrotoxic periodic paralysis (51), hypokalemia was present
in all 24 initial episodes and serum potassium levels varied from 1.1 to 3.4 mmol/L.

Figure 3 A patient with Graves’ ophthalmopathy with predominant superior limbal ker-
atopathy (8).
44 Burman and Cooper

A B

Figure 4 A patient with Graves’ disease demonstrating bilateral exophthalmos and


achropachy (clubbing) (left panel). The right panel shows a radiograph of the same patient
demonstrating phalangeal periosteal reaction (arrow) (8).

Hypophosphatemia was present initially in 12 (80%) of 15 episodes. No patient


had a recurrent episode of paralysis after becoming euthyroid. The precise patho-
physiology of these events is unknown, but patients may have a genetic predisposi-
tion to activation of Na/K-ATPase activity, which is enhanced in hyperthyroidism
(53). One suggested treatment regimen is to administer 30 mEq potassium orally
every two hours with close monitoring of serum potassium and cardiac status,
as rebound hyperkalemia occurs commonly (51). Propranolol, given either orally
or intravenously can reverse or prevent attacks (54). Restoration of the euthyroid
state prevents future attacks.
The central nervous system (CNS) manifestations of Graves’ hyperthy-
roidism are varied but include restlessness, irritability, nervousness, and impa-
tience. Some patients may realize that they have a decreased ability to concen-
trate and remember facts; occasionally, they may have demonstrable personality
changes (55). These features are difficult to quantify, but relatives may help to
identify them. After patients are treated and euthyroidism is restored, patients will
frequently comment upon their previous personality changes. Cognitive function
may be more severely impaired, especially in hospitalized elderly patients (56),
and seizures and coma can be the presenting features of “thyroid storm” (see
below). Depression and irrational or even criminal behavior are very unusual. It
is difficult to prove that serious personality disorders or criminal behavior are
directly associated with hyperthyroidism per se, although hyperthyroidism has
been implicated in these circumstances (57). Rarely, hyperthyroid patients may
present with other neurologic findings, such as chorea (58).
Women may have irregular menses and decreased fertility, but amennorhea
is rare (59). Men may have decreased libido and gynecomastia, thought to be
Hyperthyroidism 45

Figure 5 A patient with Graves’ disease showing gynecomastia. The cause of gyneco-
mastia in this circumstance is thought to be related to increased conversion of testosterone
to estradiol. Other causes of gynecomastia, such as hCG-secreting tumors, should also be
considered (8).

related to increased estrogen production (Fig. 5) (60). Total serum estrogen levels
are usually increased, in part related to increased sex hormone–binding globulin
levels. Serum LH concentrations are increased and there may be Leydig cell
failure associated with impaired spermatogenesis (61). A recent study suggested
that up to 50% of hyperthyroid men have some aspect of sexual dysfunction
that is recovered with therapy (62). Generalized lymphadenopathy, splenomegaly,
and thymic enlargement may occur, although other causes should be excluded.
A normochromic normocytic anemia has been described, probably related to
decreased ability to incorporate iron into red blood cell precursors.
Pretibial myxedema results from excessive lymphocyte infiltration in the
pretibial area, with resultant mucopolysaccharide secretion and deposition by
fibroblasts (63–65). The clinical result may simply be a small area of raised dis-
coloration in the pretibial area. Rarely, a large area of induration and nonpitting
46 Burman and Cooper

A B

Figure 6 (A and B) Two different patient with pretibial myxedema demonstrating vary-
ing degrees of involvement. The left panel illustrates more minimal involvement with skin
thickening, while the right panel shows severe thickening, which would make daily activ-
ities, such as walking with shoes, difficult. The patient on the right also has a patch of
vitiligo (8).

edema may develop, sometimes involving the entire lower leg. In this circum-
stance, the patient may have difficulty wearing shoes and the area may be pruritic
and even painful (Fig. 6). Although the cause of pretibial myxedema is unknown, it
seems to be related to anti-TSH receptor antibody levels (63). Pretibial myxedema
usually does not occur unless a patient has clinical evidence of ophthalmopa-
thy, and pretibial myxedema may occur in other anatomic sites, such as the feet,
face, or preradial area. Topical steroids, usually recommended to be used under
an occlusive dressing is the most effective therapy, but the response is poor in
patients with more severe disease.
The manifestations of Graves’ hyperthyroidism that occur in younger indi-
viduals may be different from those in older subjects (66–68). Younger patients
tend to have more classic findings, such as nervousness, weight loss, anxiousness,
tachycardia, and heat intolerance. Older patients may have none of these manifes-
tations but may only present with weight loss or a cardiac abnormality, especially
atrial fibrillation. The explanation for these differences is unknown, and, of course,
these comments should be taken as generalizations with many exceptions. A more
complete discussion of this topic is found in chapter 10, “Practical Management
of Thyroid Disease in the Elderly.”
Hyperthyroidism 47

Laboratory Diagnosis
Thyroid Hormone and TSH Levels
In the past, common thyroid hormone measures included total T4, total T3, resin
T3 uptake, and TSH. However, recent advances in techniques now allow the direct
measurement of free T4 (FT4). This analysis is preferred to the total T4, as ⬎99%
of T4 is bound to circulating proteins (thyroxine-binding globulin, albumin, and
prealbumin) and ⬍1% is unbound and available to enter cells, and, after conversion
to T3, to bind specific nuclear receptors and mediate biologic activity. Total T4
measurements are affected by factors that influence thyroid hormone–binding
proteins, including drugs (estrogens, birth control pills, and androgens, opiates),
and medical conditions such as hepatitis, cirrhosis, and nephrotic syndrome. FT4
levels remain normal in these situations (3). The resin T3 uptake test was designed
to indirectly assess the quantity of thyroid hormone–binding proteins in serum and
was used as an adjunct to the total T4 concentrations. However, now that we have
the ability to measure FT4 directly, rapidly, and inexpensively, the measurement
of FT4 is preferred. Total T3 is still measured because FT3 cannot yet be measured
in most laboratories in a reliable, precise, cost-effective manner. Approximately
5% of patients will have a normal serum free T4 level and elevated serum T3 level
(“T3 toxicosis”), and some patients, especially the elderly can have “T4 toxicosis”
with normal serum T3 levels. Like total T4, total T3 levels are altered by situations
that change thyroid hormone–binding proteins. TSH assays have also improved,
and “third generation” assays can measure ⬍0.01 mU/L in serum, with the normal
range being approximately 0.5 to 4.5 mU/L (51–57). These improvements in
sensitivity result from utilizing chemiluminescent techniques. All patients with
conventional forms of hyperthyroidism should have an undetectable TSH level in
third-generation assays, although many commercial laboratories only report that
a value is ⬍0.1 mU/L.

24-Hour Radioiodine Uptake


Serum measurements of thyroid hormone and TSH are the cornerstone of the
diagnosis of hyperthyroidism, but they do not assess biologic activity or the tissue
effects of the circulating thyroid hormone levels. The capacity of the thyroid gland
to concentrate radioactive iodine is a physiologic test representing in vivo events.
A normal subject will concentrate approximately 8% to 30% of radioactive
iodine administered when determined at 24 hours. Patients with hyperthyroidism
will usually concentrate higher amounts of radioactive iodine than normal,
reflecting the heightened ability of the gland to concentrate iodine. On the other
hand, patients with thyrotoxicosis and a low radioiodine uptake generally have
a problem associated with increased release into the circulation of preformed
thyroid hormone, for example, the various form of thyroiditis. (see chap. 3,
“Hyperthyroidism.”) Radiocontrast dyes and other sources of exogenous iodine
such as amiodarone will interfere with this test, because the enormous amounts of
48 Burman and Cooper

Table 2 Anti-TSH Receptor Antibody Measurements


Method Nomenclature Frequency Advantage Disadvantage

cAMP TSIa 80–100% Stimulatory Difficult to perform


generation
125
I-TSH TBIIb 70–90% Easy to perform Detects all TSH
binding receptor antibodies
inhibition
a TSI, thyroid stimulating immunoglobulins. This test detects cAMP generation in vitro from thyroid
cells or cells transfected with TSH receptor.
b TBII, thyroid binding inhibitory immunoglobulins. This test detects the ability of serum samples to

displace 125 I-TSH binding from thyroid cell membranes or cells transfected with TSH receptor.
Source: From Ref. 2.

unlabeled iodine in these compounds dilute out the radioactive tag, resulting in less
radioactive iodine being concentrated by the thyroid gland and a very low 24-hour
uptake (69). An elevated radioactive iodine uptake can also be seen in euthyroid
or even hypothyroid patients with an organification defect, most notably patients
with Hashimoto’s thyroiditis (70,71). Therefore, an elevated radioactive iodine
test is not specific for hyperthyroidism. Since dietary iodine intake has decreased
over time (72), the normal range for the 24-hour uptake may have increased
compared to values obtained 20 or 30 years ago. Variations in geographic and
individual dietary iodine intake of bread, pastry, seafood, salt, and dairy products
may also contribute to changes in the 24-hour uptake. Most institutions have not
reassessed their normal range for this test in many years, mainly because it is
difficult to justify the administration of radioactive materials to normal subjects.
Therefore, the normal range should be considered as a guide and not an absolute
limit. Some clinicians feel that a scan should be performed whenever a radioiodine
uptake is ordered to assess whether an undetected cold nodule also may be present.

TSH Receptor Antibody Measurements


Anti-TSH receptor antibody measurements can be performed with one of two
possible assays (Table 2) (4,10). Stimulatory TSH receptor immunoglobulins (TSI)
are measured in vitro by testing the ability of serum or IgG from a possible
Graves’ patient to stimulate thyroid cells and generate cAMP or enhance T3
and T4 secretion (4,73–75). These responses are compared to those of normal
control serum or IgG. A ⬎30% increase above control serum is considered to
be a positive cAMP response. Usually, sera from Graves’ disease patients will
stimulate cAMP by more than two- or threefold. The advantage of the TSI assay
is that it measures TSH receptor–stimulating antibodies, which are relevant to
hyperthyroidism; but its disadvantages are that it is relatively expensive, and the
sensitivity of commercial assays vary. Another test, called a thyrotropin-binding
Hyperthyroidism 49

inhibitory immunoglobulin assay (TBII assay), measures the total conglomerate


amount of TSH receptor antibodies in serum. The ability of serum or IgG from
hyperthyroid patients to inhibit radiolabeled TSH binding to bovine, human, or
recombinant TSH receptors is compared to control normal serum or IgG (76).
Graves’ disease patients’ serum samples cause more than 10% inhibition of binding
compared to controls, with values frequently being as high as 50% to 70% in
patients with active disease. The potential disadvantage of TBII measurements is
that they do not distinguish stimulatory from inhibitory antibodies.
There may be wide variability in the nature and type of TSH receptor anti-
bodies identified in different patients (77–79). Di Cerbo et al. (78) have suggested
that, in addition to TSH receptor antibodies that stimulate cAMP production, other
antibodies may be capable of stimulating alternative pathways. For example, 72%
of Graves’ patients’ serum contained immunoglobulins that could stimulate the
phospholipase A2 pathway.
The measurement of TSH receptor antibodies in clinical settings is only
occasionally indicated (4). They can help differentiate Graves’ disease from other
causes of hyperthyroidism when this differentiation cannot be made clinically
and when it is relevant to do so. Anti-TSH receptor antibody measurements may
also be useful to help confirm the presence of Graves’ ophthalmopathy from
other nonendocrine causes of proptosis in euthyroid patients. Anti-TSH receptor
antibody measurements may also help predict if a patient with Graves’ disease
is in remission at the end of a course of antithyroid drug therapy, but this test
is not very sensitive or specific (80). Most importantly, anti-TSH receptor anti-
bodies may be elevated in the sera of pregnant women who have (or have had)
active autoimmune thyroid disease, such as Graves’ disease. If the TSH recep-
tor antibodies are markedly elevated (e.g., two- to threefold above the upper
limit of normal), there is an increased likelihood that these IgG antibodies will
cross the placenta and cause neonatal hyperthyroidism (81,82). Anti-TSH recep-
tor antibodies can be found in the sera of pregnant women with Graves’ disease,
but they may also be present in the sera of patients who have had Graves’ dis-
ease and even in women who have had Hashimoto’s thyroiditis. Many experts
suggest measuring these antibodies in the third trimester in women with active
Graves’ disease or Graves’ ophthalmopathy, in women treated for Graves’ disease
in the recent past, and in women with high-titer antithyroid peroxidase or thy-
roglobulin antibodies who have been diagnosed recently with autoimmune thyroid
disease.
There are several additional tests that may occasionally be useful in patients
with Graves’ disease. The measurement of serum antithyroid peroxidase and
thyroglobulin antibody levels should not be measured in the majority of patients
with Graves’ disease. However, elevated antibody titers gives information about
the coexistence of Hashimoto’s thyroiditis with Graves’ disease (79,83). High-titer
antibodies may help to predict remissions after a course of antithyroid drugs or
permanent hypothyroidism following radioiodine therapy or surgery.
50 Burman and Cooper

Pitfalls

There are several pitfalls that should be avoided in the laboratory assessment
of Graves’ disease patients. The assays for iodothyronines and TSH are specific
and accurate; as a result, there are few reasons for artifactual results except for
mislabeling and rare laboratory errors. Specific antibodies against T3 or T4 may
alter their respective measurements. Although unusual, such antibodies can occur
in patients with autoimmune thyroid disease, in those who work with animals, and
occasionally for no apparent reason. In a study of 115 patients with antithyroid
hormone autoantibodies, approximately 42% of patients had antibodies against
triiodothyronine, 33% against thyroxine, and 25% of patients had both anti-T3
and anti-T4 antibodies (84). Although 44% of these patients were considered to
be euthyroid, 16% were hyperthyroid and almost 40% were hypothyroid. While
the effect of antibodies on T4 and T3 measurements depends upon the method of
measurement, in general, they cause a laboratory result that is incongruent with
the clinical state (84).
It is important to ensure that patients with Graves’ hyperthyroidism have an
undetectable serum TSH level, so that rare individuals with peripheral hormone
resistance or TSH secreting pituitary tumors are not misdiagnosed as having
Graves’ disease. Also, certain patients hay harbor heterophilic antibodies that
can result in falsely elevated serum TSH levels, causing diagnostic confusion in
patients with hyperthyroidism (85). It is also important to exclude other coexistent
autoimmune disorders that may be confusing the clinical picture. For example, a
patient with Graves’ hyperthyroidism who complains of inordinate weakness and
tiredness may, in fact, have coexistent Addison’s disease.
The patient’s clinical assessment and history must be integrated with the
thyroid function tests. Thyroid function tests should ideally be determined twice
prior to treatment and, when possible, measurement of FT4 and T3 is preferred.
In patients with mild hyperthyroidism and no clinical features of Graves’ disease,
it is important to obtain a radioactive iodine uptake test prior to treatment to
confirm high-uptake thyroid disease. Even if the diagnosis of Graves’ disease
seems obvious, at least one TSH measurement should be obtained to rule out the
unlikely TSH-secreting pituitary tumor.
The term euthyroid sick syndrome refers to the thyroid hormone changes that
occur in patients with a wide variety of systemic illnesses (86–88). The presence
of the euthyroid sick syndrome may interfere with the interpretation of thyroid
function tests in a patient being evaluated for possible hyperthyroidism. Early in
the euthyroid sick syndrome, T4 and FT4 remain normal, although occasionally
they may be decreased or even slightly increased. Total and FT3 are decreased
due to diminished T4 to T3 conversion. Serum TSH generally remains within the
normal range but may be decreased slightly or may increase during the recovery
phase of systemic illness. In this circumstance a systemically ill patient with
moderate or severe hyperthyroidism may have few symptoms, and the serum T3
Hyperthyroidism 51

may be inappropriately within the normal range, because of decreased T4-to-T3


conversion (so-called T4 toxicosis).

Treatment
Ideally, the treatment of any medical condition is directed at its cause, but the
cause of the immune dysregulation in Graves’ disease remains obscure (3,36).
Therefore, the available treatments are directed at the thyroid gland rather than the
underlying autoimmunity. The therapies that are available to the clinician in the
21st century are the same as those that were available 50 years ago: Antithyroid
drugs, radioiodine, and surgery. Although some patients, especially those who
are relatively asymptomatic, may wonder whether specific treatment is necessary,
overtly hyperthyroid individuals usually require restoration of a euthyroid state
because of potentially deleterious skeletal, cardiovascular, and psychologic effects.

Antithyroid Drug Therapy


Antithyroid drugs remain the first choice for initial therapy of children, adoles-
cents, and young adults in the United States (89,90) and are the usual treatment
for almost all patients in the rest of the world (91). Antithyroid drugs are generally
safe and effective in controlling the hyperthyroid state. However, they have limi-
tations and toxicities that are important to recognize, and their proper use requires
knowledge of their pharmacology as well as clinical experience.

Clinical Pharmacology of Antithyroid Drugs


Antithyroid drugs do not directly affect iodine uptake or hormone release by the
thyroid; hence, contrary to popular belief, the 24-hour radioiodine uptake is not
affected very much by antithyroid drug therapy. Within the thyroid, both propy-
lthiouracil (PTU) and methimazole (Tapazole) inhibit thyroid hormone synthesis
by interfering with intrathyroidal iodine utilization and the iodotyrosine coupling
reaction, both of which are catalyzed by thyroid peroxidase. Extrathyroidally,
PTU, but not methimazole, inhibits the conversion of T4 to T3 in peripheral tis-
sues. Although some feel that this difference confers an advantage of PTU over
methimazole, there are no data to support this supposition; in fact, comparative
studies show that methimazole generally normalizes serum T4 and T3 levels faster
than PTU (92–94). There are also in vitro and in vivo data pointing to possible
beneficial effects of both drugs on the immune system, although it is far from clear
whether this is important clinically in terms of remission rates with antithyroid
drug therapy (89).
Antithyroid agents are well absorbed from the gastrointestinal tract. In the
circulation, PTU is heavily protein-bound, mainly to albumin, while methimazole
binding to proteins is negligible. This characteristic may affect antithyroid drug
choice in pregnancy and lactation, since PTU would be expected to cross the pla-
centa and breast epithelium less readily than methimazole. The serum half-lives
52 Burman and Cooper

of PTU and methimazole are 1 and 4 to 6 hr, respectively. However, the intrathy-
roidal duration of action of both drugs is longer than that, making the determination
of drug blood levels not particularly helpful clinically. Although both drugs are
metabolized in the liver and metabolites are excreted by the kidney, in the absence
of data to the contrary, the doses used to treat hyperthyroidism do not generally
need to be altered in patients with liver or kidney disease.

Antithyroid Drugs in Clinical Practice


Antithyroid drugs are used in two ways in the therapy of hyperthyroidism (89).
They can be employed as primary therapy, and are usually given for one to
two years in the hope that the patient will achieve a remission (a remission is
usually defined arbitrarily as biochemical euthyroidism for one year following
cessation of the antithyroid drug); or, they are used for a few months to “cool the
patient down” prior to ablative therapy with radioiodine or surgery. Unfortunately,
patients are often started on antithyroid drugs without a clear goal in mind and
then remain on them either continuously or intermittently for protracted periods
of time. Antithyroid drugs are also mistakenly used in the long-term treatment of
toxic nodules or toxic multinodular goiter, situations in which remission is highly
unlikely.

Antithyroid Drugs for Primary Therapy of Graves’ Disease


Prior to initiating antithyroid drug therapy, the physician should carefully discuss
the options with patients and their family. Unlike radioiodine and surgery, antithy-
roid drug treatment will not cause permanent hypothyroidism, but the chances of
remission are ⬍50% for an average patient. Even if a remission occurs, the chances
of permanent remission are ⬍50%, and late hypothyroidism may develop in up to
20% (95). Also, the potential for allergic reactions is often underestimated or not
discussed. Patient preferences are important to take into consideration, even if the
decision for or against a particular therapy seems to be based more on emotion
than fact. No therapy has been shown to be superior to any other in terms of effi-
cacy or patient satisfaction (96). However, the physician can help the patient make
an informed choice. Remissions are less likely with large goiters and in severe
disease, especially when the serum T3 concentration is ⬎900 ng/dL (97). In one
report that used a proportional hazards analysis to examine various potential fac-
tors influencing the outcome of antithyroid drug therapy, only the baseline serum
T3 level emerged as an independent risk factor for relapse (97). Some studies have
found that “T3 predominant” disease, in which the unitless serum T3/T4 ⬎20,
also makes remissions less likely (98), but others have not confirmed this obser-
vation (96). A prior history of relapse is another factor that would argue against
antithyroid drug use as first-line therapy. On the other hand, a small gland and
mild biochemical changes would favor a remission, and in some studies the rates
may be as high as 70% to 90%. A negative TSI titer at the beginning of therapy
has been shown to predict a high rate of remission (99), but negative titers occur
in only approximately 10% of patients, so it is probably not cost-effective to order
Hyperthyroidism 53

TSI titers routinely. Age, sex, family history of Graves’ disease, the presence of
ophthalmopathy, and smoking behavior are not reliably or consistently predictors
of remission.
Family planning is another factor that should be considered in women. First,
many clinicians feel that if pregnancy is desired in the following one to two years,
antithyroid drugs are less appropriate, since the patient may be pregnant while
taking a drug that could harm the fetus. Also, in patients who have had a remission
after a course of antithyroid drugs, relapse is very common in the postpartum
period (100). Therefore, some clinicians feel that women desirous of pregnancy in
the near future are not optimal candidates for long-term antithyroid drug treatment.
Once antithyroid drugs are selected as initial treatment, a choice between
PTU and methimazole must be made. Methimazole has a number of advantages
over PTU. First, it is a once-a-day drug (101), which improves compliance, and the
numbers of methimazole tablets that a typical patient takes daily is fewer, which is
important to some patients. Second, the toxicity of methimazole is probably more
predictable, in that the frequency of side effects is dose-related; many patients can
be treated with doses as low as 5 to 15 mg/day, a dose range in which side effects
are few (102). There is no dose relationship with side effects for PTU. Also, the rare
side effects of drug-induced lupus, vasculitis, and hepatitis are far more common
with PTU (94). Because of its protein binding and lack of teratogenic effects,
PTU is preferred in pregnancy and lactation, although methimazole has been used
safely. Also, PTU may be preferable in thyroid storm or severe hyperthyroidism
because of its ability to block T4-to-T3 conversion. High-dose methimazole is
more expensive than PTU, but low doses have a comparable cost. On an average,
methimazole costs approximately $80/mo (30 mg/day) as compared to the cost of
PTU ($30/mo; 300 mg/day).
In mild to moderate thyrotoxicosis, methimazole is usually started at a dose
of 10 to 30 mg/day as a single daily dose, and PTU is begun at a dose of 100 mg
three times a day. The rapidity of response depends on the severity of the under-
lying thyroid problem, the size of the gland and its hormonal stores, the dose
and frequency of the drugs, and, of course, compliance. Most patients become
euthyroid within 6 to 12 weeks, with methimazole possibly acting faster than PTU
to normalize both T3 and T4 serum levels (92), especially in patients with more
severe degrees of thyrotoxicosis (94).
Although it may take longer to achieve control than with higher doses
(103), initial doses as low as 10 mg/day can control hyperthyroidism in many
patients (102). Although older studies had suggested that remission rates might be
increased by using high doses of antithyroid drug, more recent randomized trials
have not found this to be the case (102,104). High-dose therapy has the decided
disadvantage of being associated with higher rates of side effects (94,102,105).
Once antithyroid drugs have been started, thyroid function should be moni-
tored every four to six weeks, at least for the first six months and less frequently
thereafter. Normalization of serum T3 can lag behind serum T4, so it is important
to follow both measurements. During treatment, some patients can have startling
54 Burman and Cooper

degrees of “T3 predominance,” with serum T3 levels two to three times above the
upper limit of normal and serum T4 values that are subnormal (106). Also, the
serum TSH level can remain suppressed long after the patient has become euthy-
roid or even hypothyroid, which limits its value early in the course of treatment.
In many patients, the drug can be tapered to a lower dose after a few months,
once the patient has become biochemically euthyroid. If this tapering is not done,
hypothyroidism will often ensue (103). In patients who are hyperthyroid on a
low drug dose but hypothyroid on a larger dose, some physicians use a “block-
replacement” regimen. In this method, a dose of antithyroid drug that would cause
hypothyroidism is employed in conjunction with thyroxine supplementation to
maintain a euthyroid state. This method of treating patients may also be useful in
the pediatric population.
The “block-replacement’’ regimen has also been used in the hope that the
combination will yield a higher remission rate than antithyroid drugs alone (107).
This strategy is based on the theory that thyroxine supplementation will suppress
the serum TSH, thereby diminishing the immune system’s exposure to thyroidal
antigens, especially the TSH receptor. While one well-done, randomized, clinical
trial from Japan did show a higher remission rate using combined therapy (107), a
number of equally rigorous trials have failed to confirm these original observations
(108–112). At the present time, most clinicians have abandoned the idea that
combined antithyroid drug–thyroxine therapy enhances the chances of remission.
We prefer simply to taper the antithyroid agent dose as required and not to add
L-thyroxine, which would also make it more difficult to interpret and follow T4
and T3 levels as a measure of endogenous thyroid function.
Once a patient has been placed on long-term therapy with an antithyroid
drug, what is the optimal duration of therapy before a remission has been achieved
and the drug can be discontinued? Older retrospective data suggested that the
longer a person remained on therapy, the more likely a remission would be
achieved once the drug was stopped (113,114). More recently, prospective tri-
als have not shown longer treatment periods (e.g., ⬎12–18 months) to be more
effective (115,116). Therefore, treatment for a year is reasonable, but data sup-
porting longer periods of time are lacking.
After one to two years of therapy, most clinicians discontinue antithyroid
drugs to see whether a remission has occurred. If the goiter has gotten smaller
and the patient’s disease has been controlled on diminishing doses of drug, the
chances of remission are greater. On the other hand, a large goiter, continued
requirement for large doses of drugs, and a persistently high-T3/T4 ratio are all
poor prognostic signs. Although a number of tests have been proposed to predict
the odds that a patient will remain euthyroid off drugs, none has the requisite
sensitivity or specificity to be useful clinically; thyroid-stimulating antibody test-
ing is the most widely studied. If it is positive, the chances of remission are very
low, but even patients with negative titers have a 20% to 50% chance of relapse
(117,118). Practically speaking, it is probably more sensible to taper the antithy-
roid drug rather than to stop it abruptly. Patients should be monitored closely and
Hyperthyroidism 55

thyroid function tests checked monthly, and the drug should be gradually tapered
to discontinuation. Thyroid function tests are then performed every four to six
weeks, but patients are not necessarily seen for an office visit unless they become
hyperthyroid, or at three to four months if they remain normal. T3 thyrotoxicosis
frequently occurs during a relapse, so that serum T3 should be monitored along
with the serum T4 levels. Relapses are most likely to occur within the first six
months after drug discontinuation (119).
Some patients have persistent subclinical hyperthyroidism, with normal
serum T4 and T3 values but suppressed serum TSH concentrations. While the
chances that such patients will have a full-fledged relapse are greater (120), relapse
is not inevitable. Some experts treat patients with subclinical hyperthyroidism (i.e.,
suppressed TSH, normal FT4, and T3) as if they had relapsed and recommend
another trial of antithyroid drug therapy or radioiodine. Others simply observe
them expectantly and only recommend treatment if and when overt hyperthy-
roidism develops.
Remissions are not necessarily lifelong. Older data suggested that eventual
relapse was almost inevitable (121). However, more recent long-term follow-up
studies have shown that some patients have durable remissions that apparently
last for many years (122). A strategy for treatment of relapse should be discussed
with the patient in advance. Some patients will opt for another course of antithy-
roid drug, even though more than one prior relapse is associated with continued
relapse. Other patients will wish to move on to definitive radioiodine therapy, or,
more uncommonly, to surgery. Some patients end up on chronic antithyroid drug
treatment for decades without ill effect (123), but this is unusual unless the patient
is highly motivated and the hyperthyroidism relatively mild. As noted above,
some patients eventually develop spontaneous hypothyroidism, so that lifelong
follow-up is necessary.

Antithyroid Drug Side Effects


The side effects of antithyroid drugs are usually classified as “minor” or “major,”
depending on the level of potential harm to the patient (Table 3) (124). Overall, side
effects develop in 5% to 25% of patients and are among the most frequent reasons
for abandoning drug therapy. As noted above, methimazole-related drug reactions
are dose-related, but this does not appear to be the case for PTU. The commonest
minor reactions are fever, rash, pruritus, arthralgias, gastrointestinal distress, and
nausea. Rashes can be urticarial, macular, or morbilliform. A recent prospective
study reported that minor reactions occurred in 10% of PTU-treated individuals and
15% of those receiving methimazole, a difference that was not significant (125).
In another prospective trial, rash developed in approximately 20% of patients
treated with 30 mg of methimazole or 300 mg of PTU daily, versus 6.6% with
15 mg of methimazole daily (94). If a rash develops, it will sometimes resolve
spontaneously (even with continued use) with or without the use of antihistamines
to treat associated itching. Although switching to the alternative drug is another
56 Burman and Cooper

Table 3 Side Effects of Antithyroid Drugs


Overall frequencya Comments

Minor side effects


Skin reactions 4–6% Dose-related for MMI; possibly
more common with MMI
Arthralgias 1–5% Gastrointestinal 1–5%
Hair loss 4% Possibly related to change in
thyroid function
(hypothy-roidism)
Abnormal taste/smell only reported with MMI/CBZ
0.3%
Sialadenitis Very rare
major side effects
Severe polyarthritis 1–2%
Agranulocytosis 0.1–0.5%
Aplastic anemia Rare
Vasculitis Rare May be ANCA + drug-induced
SLE and other immune
syndromes also reported
Severe hepatitis 0.1–0.2%; 1% with Almost exclusively PTU; transient
high-dose PTU (14) increases in transaminases seen
in 30%
Cholestasis Rare Almost exclusively seen with MMI
or CBZ; no deaths reported
Hypoprothrombi Rare No case reports since 1982
nemia
Insulin-autoimmune Rare Seen almost exclusively in Asians
syndrome
a Rate of side effects (minor and major) is greater at high doses of MMI and may approach 30% at
high doses.
Abbreviations: MMI, methimazole; CBZ, carbimazole; PTU, propylthiouracil; ANCA, antineutrophil
cytoplasmic antibody.

possibility, the cross-reaction rate may be as high as 50%. Some patients may
simply elect to stop the offending drug and accept a definitive form of therapy.
Loss of sense of taste, sometimes associated with anosmia, is a rare minor
side effect reported only with methimazole (126). It develops suddenly after one to
two months of therapy and resolves after the drug is stopped. This side effect has
not been reported with PTU, although PTU may cause a metallic taste. A recent
report described the development of arthralgias and elevated serum CPK levels
in four patients treated with methimazole at a time when they were still clearly
hyperthyroid (127). The authors postulated that this side effect was the result of
rapid a decline in serum thyroid hormone levels.
Hyperthyroidism 57

Fever and arthralgias, while technically minor side effects, warrant drug dis-
continuation since they may be the harbinger of more serious problems. Similarly,
leukopenia, defined as a white blood cell (WBC) count ⬍4 × 109 /L, occurs in up
to 10% of patients. Leukopenia requires follow-up and prompt cessation of the
antithyroid drug if the WBC count falls below 3 × 109 /L, since leukopenia may
precede the development of full-blown agranulocytosis (see below). Antithyroid
drug-related leukopenia should be distinguished from the leukopenia that can be
seen in Graves’ disease and in healthy African-Americans by obtaining a baseline
WBC count.
The major side effects are quite rare, but the most frequent are agranu-
locytosis, vasculitis and drug-induced lupus, and hepatic damage (hepatitis and
cholestasis). Agranulocytosis develops in approximately 0.2% to 0.5% of patients.
In one case series, agranulocytosis developed in 12 of 2190 (0.55%) patients taking
PTU and 43 of 13,208 (0.31%) patients taking methimazole (128). Agranulocy-
tosis is usually defined as an absolute granulocyte count ⬍0.5 × 109 /L, but most
patients have granulocyte counts that are far lower, often close to zero. It should be
distinguished from the exceedingly rare cases of antithyroid drug–induced aplastic
anemia by a hematocrit ⬎30% and a platelet count 100 × 109 /L. Agranulocyto-
sis is thought to be autoimmune in origin, developing because of antigranulo-
cyte antibodies that are found in the serum of affected patients (129). Since the
development of agranulocytosis may be HLA-linked (130), it is probably wise
to avoid giving antithyroid drugs to close relatives of a patient who has had this
side effect.
Agranulocytosis typically develops in the first three months of therapy,
but there are notable exceptions (131). Older patients may be more susceptible,
and it can develop after one or more prior innocuous exposures to the drugs.
Routine monitoring of the WBC count has not been recommended because it is
not cost-effective, but a prospective study has cast doubt on the wisdom of this
policy (130). Antithyroid drug–treated patients had serial WBC counts performed
every two weeks for the first two months and monthly thereafter. Some patients
developed milder forms of agranulocytosis that resolved without progressing once
the antithyroid drug was stopped (130). In another study, it was shown that it
may be possible to differentiate those patients with modest depression of their
granulocyte count (granulocyte counts of 0.2–0.8 × 109 /L) who will go on to
develop full-blown agranulocytosis using a single dose of granulocyte-colony
stimulating factor (G-CSF) (128). Some clinicians monitor the complete blood
count and liver function tests in patients taking antithyroid agents, both prior to
and periodically during this treatment.
Patients with agranulocytosis may be afebrile until they develop an infection.
The typical victim has severe malaise, oropharyngitis and odynophagia, and high
fever. Immediate cessation of the antithyroid drug, hospitalization, and administra-
tion of broad-spectrum antibiotics is mandatory, with coverage for Pseudomonas
aeruginosa, which is frequently isolated from the blood in affected patients (132).
Although most patients recover, it should be recalled that agranulocytosis has
58 Burman and Cooper

been associated with a mortality rate as high as 16% (133). A bone marrow exam-
ination may provide helpful prognostic information; an extreme loss of myeloid
precursors suggests a longer time to recovery (133) as well as the potential for
a poor response to G-CSF therapy (134). The use of G-CSF has become stan-
dard in the management of drug-induced agranulocytosis. However, a randomized
controlled trial of G-CSF in antithyroid drug–induced agranulocytosis found no
statistically significant difference in the mean time to recovery from G-CSF com-
pared to conservative therapy (135). Other retrospective data supporting the use
of G-CSF, however, noted shortened recovery time. In a retrospective review of
109 patients with agranulocytosis related to antithyroid drugs, the mean time to
recovery was 5.5 days compared to 9.2 days in historic controls. There were no
deaths in either group (136). The doses of G-CSF used in the literature range
from 1 to 5 ␮g/kg/day subcutaneously. Common side effects include rashes, bone
pain, myalgias, and headache, all of which respond well to acetaminophen (137).
If thyrotoxicosis requires treatment during the acute episode of agranulocyto-
sis, beta-adrenergic blocking drugs, lithium, or iodinated contrast agents can be
used. Attempting to switch to the other antithyroid drug is not recommended, as
cross-sensitivity has been reported.
Some patients develop a condition that has been termed the “antithyroid
arthritis syndrome” (138). The frequency of this side effect is in the range of 1%
to 2%, and it usually develops within 60 days of initiating therapy. The syndrome
is characterized by hot, swollen, tender joints involving multiple sites. The spe-
cific laboratory studies are not typical of drug-induced lupus. Symptoms usually
resolve after one to two weeks of therapy with nonsteroidal anti-imflammatory
drugs; glucocorticoid therapy may be necessary in severe cases. A more uncom-
mon and potentially significant development is the appearance fever, rash, arthri-
tis, splenomegaly, glomerulonephritis, and other stigmata of drug-induced lupus,
almost always associated with PTU. Patients have elevated sedimentation rates,
positive antibodies for double-stranded DNA, and low serum complement levels.
The syndrome usually resolves in a few weeks after the drug is discontinued,
although it should be remembered that the simultaneous occurrence of other
autoimmune disorders is increased in patients with Graves’ disease.
Recently, an antithyroid drug–related syndrome that includes renal failure,
vasculitic skin changes, pulmonary and respiratory tract involvement, arthritis,
and positive circulating anticytoplasmic neutrophil antibodies (ANCA) has been
described, mainly in Asian patients (139). ANCA have typically been associ-
ated with Wegener’s granulomatosis and polyarteritis nodosa, but they may also
be present in drug allergy. In the antithyroid drug–related cases, the antibod-
ies are of the pericytoplasmic variety (so-called pANCA, with myeloperoxidase
being the putative antigen), and the vast majority have been patients exposed to
PTU (140,141). Although the syndrome usually resolves after a few weeks, some
patients with severe renal dysfunction or pulmonary involvement have required
high-dose glucocorticoid therapy or cyclophosphamide, and several patients have
Hyperthyroidism 59

needed short-term hemodialysis. In some antithyroid drug-treated patients, ANCA


are present, but patients remain asymptomatic (142).
Hepatic involvement with PTU typically presents as clinical hepatitis with
malaise, anorexia, jaundice, and tender hepatomegaly. Laboratory data and liver
biopsy histology are consistent with hepatocellular injury. The following criteria
for the diagnosis of PTU-induced hepatitis have been proposed: Clinical and lab-
oratory evidence of hepatocellular damage; temporal relationship to PTU therapy;
exclusion of known infectious agents, drugs, or toxins; and absence of shock or
sepsis (143). More than three dozen cases of PTU-related hepatitis have been
reported in the literature, with several fatalities and at some patients requiring
liver transplant (144). The mean duration of PTU therapy in reported cases is
three months, with a range of two days to one year; the average age of affected
individuals in one review was 28 years (144). Once the syndrome is recognized,
immediate cessation of the drug is mandatory. Expert management of potential
complications and hepatic failure is essential. Although glucocorticoid therapy has
been used, there is no evidence that it decreases the time to recovery or survival,
and glucocorticoids are not recommended (145). There have been several patients
whose ongoing hyperthyroidism has been managed successfully with methimazole
(146–148); other options would include beta-adrenergic blocking drugs, iodinated
contrast agents, and lithium.
Approximately one-third of PTU-treated patients develop asymptomatic
two-to sixfold elevations of serum tranaminases within two months of starting
the drug, which then resolve despite continued therapy (149). Also, up to 35%
of patients with Graves’ disease have elevations of liver function tests at baseline
(149,150). In one report, PTU therapy led to normalization of liver function tests
in two-thirds of patients, while the remaining one-third had a further elevation
before levels returned to baseline (149,150). These data suggest that abnormal
liver function tests are not an absolute contraindication to PTU therapy, although a
serious discussion of these issues must be held with the patient. All patients about
to embark on a course of PTU should be warned about the possibility of hepatitis
and told to discontinue the drug if malaise, jaundice, dark urine, or light-colored
stools develop. Monthly monitoring of liver function, at least for the first 6 months
of therapy, is a reasonable approach.
Methimazole therapy has not been associated with potentially lethal hep-
atic involvement. Rather, a cholestatic picture is characteristic, with severe hyper-
bilirubinemia, bile duct stasis, and preserved hepatocellular architecture on biopsy.
Although a recent review collected 30 cases in the literature (151), there are prob-
ably many cases that go unreported, as with PTU-induced hepatitis. The syndrome
usually resolves slowly over a period of several months after the drug is stopped.
In one case report of methimazole-related hepatotoxicity, PTU was substituted
for methimazole without sequelae (152). Extreme caution should be used when
employing one antithyroid agent in a situation in which the alternative agent has
caused hepatic abnormalities.
60 Burman and Cooper

Beta-Adrenergic Antagonist Drugs


Beta-adrenergic antagonist drugs play an important role in the management of
thyrotoxicosis (153). Blockade of adrenergic receptors provides patients with
considerable relief from adrenergic symptoms such as tremor, palpitation, anxiety,
and heat intolerance. Small decreases in serum T3 concentrations occur in patients
treated with large doses of selected beta-adrenergic antagonist drugs (propranolol)
because of inhibition of extrathyroidal conversion of T4 to T3 (154), but these are
probably clinically insignificant.
Although beta-blockers improve the negative nitrogen balance (155) and
decrease heart rate (156), cardiac output (157), and oxygen consumption (158)
in thyrotoxic patients, these measurements seldom become normal (159) except
in the mildest cases. Therefore, these drugs are used as primary therapy only in
patients with self-limited forms of thyrotoxicosis (e.g., the various forms of sub-
acute thyroiditis). They are most often used in Graves’ disease as an adjunct to
alleviate symptoms during the diagnostic evaluation, while awaiting the effects of
antithyroid drugs, the results of ablative therapy with radioiodine, or to prepare
patients for surgery.
Although propranolol was the drug originally used by most clinicians for
therapy of thyrotoxicosis, other beta-blockers have a longer duration of action (e.g.,
long-acting propranolol, atenolol, metoprolol, and nadolol) or are more cardiose-
lective (atenolol and metoprolol). The usual starting dose of propranolol is in the
range of 80 to 160 mg/day; similar effects are produced by 50 to 200 mg/day
of atenolol or metoprolol or 40 to 80 mg/day of nadolol. Large doses (e.g.,
360–480 mg/day of propranolol) are sometimes necessary for optimum clini-
cal effects, possibly because of accelerated drug clearance (160). Propranolol and
esmolol can be given intravenously to patients who are acutely ill (see discussion
of thyrotoxic storm).
Beta-adrenergic antagonist drugs are well tolerated. Common side effects
include nausea, headache, fatigue, insomnia, and depression. Rash, fever, agranu-
locytosis, and thrombocytopenia are rare. Undesirable effects related to the beta-
adrenergic antagonist effects are far more common. Patients with a clear history
of asthma should not receive these drugs; a cardioselective drug could be used
cautiously in patients with mild asthma. Patients with a history of congestive
heart failure should not receive a beta-adrenergic blocking drug except when the
heart failure is clearly rate-related or caused by atrial fibrillation (161). Even
then, the drug should be given cautiously, preferably with digoxin. Beta-blockers
are also relatively contraindicated in insulin-treated diabetic patients, in whom
hypoglycemic symptoms may be masked. They should not be given to patients
with bradyarrhythmias or Raynaud’s phenomenon or patients being treated with
a monoamine oxidase inhibitor; they should also probably not be given routinely
to pregnant patients.
The potential usefulness of diltiazem in thyrotoxicosis has been studied
(162). This calcium channel blocking agent reduced resting heart rate by 17%,
Hyperthyroidism 61

comparable to what can be achieved with a beta-adrenergic antagonist drug. Cal-


cium channel blockers should be considered in patients with severe tachycardia in
whom beta-blockers are contraindicated, for example, in patients with both asthma
and thyrotoxicosis.

Lithium
Although lithium has not been extensively used, it is a potential second-line
agent in the treatment of hyperthyroidism. Its mechanism of action is unknown
but is thought to relate to inhibition of thyroid hormone synthesis and secretion
(163,164). Lithium use should be reserved for the unusual patient who has not
responded to or is allergic to more standard antithyroid agents. Lithium is mainly
indicated for short-term use in hospitalized patients with thyroid storm or to
prepare them for surgery. The dose of lithium is usually 300 mg tid, with periodic
monitoring of serum levels to ensure that they are in a therapeutic range. Potential
complications relate to neuromuscular and CNS perturbations, such as ataxia,
tremor, and seizures.

Potassium Perchlorate
Potassium perchlorate is an antithyroid medication that inhibits thyroidal uptake
by the sodium/iodide symporter (165–167), resulting in decreased synthesis and
secretion of T4 and T3. This agent is also a second-line alternative to more standard
therapy. The typical dose is approximately 400 mg bid po. Potential adverse effects
include bone marrow suppression and skin rash. It is unknown if the risk of bone
marrow toxicity increases with concomitant PTU or methimazole use. For several
decades the use of perchlorate had fallen into disrepute due to its potential bone
marrow toxicity. However, interest in employing this medication has resurfaced,
as it was found to be effective for amiodarone-induced hyperthyroidism (168).
In selected circumstances, its use has now extended to all types of high-uptake
hyperthyroidism. Unfortunately, perchlorate is no longer available in the United
States.

Radioiodine (131 I) Therapy for Graves’ Disease


131
I therapy has been utilized for approximately 50 years for patients with Graves’
disease; it is considered safe and effective (169–171). The goal of therapy is to
render the patient permanently hypothyroid, a process that typically takes approx-
imately three months. Radioiodine therapy is considered first-line therapy in most
adults with thyrotoxic Graves’ disease (90). Patients and their families must be
counseled about the advantages and disadvantages of radioiodine therapy and
must participate in the decision process. Elderly or severely thyrotoxic patients
considered too ill to undergo the therapeutic manipulations inherent in the pro-
cess of giving 131 I are usually initially treated with antithyroid agents to render
them euthyroid, at which time a decision can be made with regard to definitive
radioiodine therapy.
62 Burman and Cooper

A 24-hour radioactive iodine uptake test is required before administering 131 I


therapy. Measurement of radioactive iodine uptake prior to giving a therapeutic
dose of 131 I is important since it documents that the uptake is sufficiently elevated to
administer 131 I. On rare occasion, a patient will have had exposure to radiocontrast
dye or compounds containing sufficient iodine to suppress the radioactive iodine
uptake. In this circumstance, giving a therapeutic dose of 131 I would fail to treat the
patient and exposure him or her unnecessarily to radiation. 131 I therapy should not
be given to a breast-feeding woman or a patient who may be pregnant. Therefore,
a careful medical history and a serum beta-hCG should be obtained within five to
seven days prior to giving the therapy. A woman of childbearing age should be
counseled not to become pregnant for 6 to 12 months after 131 I therapy. This time
period is chosen based on the biologic half-life of the radioiodine, as well as the
desire for the patient to be euthyroid prior to becoming pregnant. Careful radiation
safety procedures must be followed by the patient and his or her family for about
a week following 131 I therapy. For example, if a patient has young children in
the house, they should not share eating utensils or be kissed or held closely. The
individual instructions vary from institution to institution and with the family
situation.
There are two general approaches to deciding on the appropriate therapeutic
dose of 131 I for a patient with Graves’ disease. The first method attempts to
determine the most appropriate dose for an individual by estimating the size of
the thyroid gland and delivering approximately 100 to 200 ␮C of 131 I per gram
of thyroid tissue. The estimated thyroid gland size is multiplied by the desired
delivered dose per gram of tissue (100–200 ␮C of 131 I), and this number is divided
by the 24-hour uptake expressed as a decimal (e.g., 80% uptake is converted to
0.80). This method appears to be more quantitative than it actually is because
clinicians tend to underestimate the size of large thyroid glands and overestimate
the size of smaller ones, and it is impossible to predict radiation sensitivity of a
specific thyroid gland. Alternatively, a second method of treatment is to use “fixed
doses” of 131 I. With this approach, the physician arbitrarily picks a given 131 I
dose that is used in all patients with hyperthyroid Graves’ disease. A typical fixed
dose would be 10 to 15 mCi 131 I. This practice has the advantage of simplicity—
but, of course, it does not take into effect the size or activity of the thyroid
gland. Retrospective and prospective studies have not shown major differences
in outcomes for the two methods of dose determination (172,173). Some centers
perform dosimetric calculations after a tracer dose of radioiodine in an effort
to deliver a specific dose of radiation to the gland (e.g., 10,000 rads), but this
method is cumbersome and time-consuming and has not gained wide popularity.
The potential complications of radioiodine therapy are listed in Table 4.
The follow-up evaluation of a patient after 131 I therapy varies between
institutions and physicians, and, of course, depends upon the clinical circumstances
and goals of treatment. In the typical patient with Graves’ disease, 131 I is given
in order to induce permanent hypothyroidism. In this circumstance, following a
therapeutic dose of 131 I, serum FT4, and total T3 levels are determined periodically,
Hyperthyroidism 63

Table 4 Complications of Radioiodine Therapy and Surgery


131
I therapy
Exacerbation of hyperthyroidism, especially in elderly patients who are not euthyroid
Transient thyroid gland discomfort or tenderness
Sicca syndrome, taste distortion, sialadenitis
Fetal hypothyroidism if administered during pregnancy
Hypothyroidism (desirable goal)
Carcinogenic effects still being evaluated
Thyroidectomy
Hypothyroidism (desirable goal)
Anesthetic complications
Hypoparathyroidism with hypocalcemia
Transient postoperative hypocalcemia (“hungry bones”)
Recurrent or superior laryngeal nerve injury causing voice changes
Local hemorrhage
Persistent or recurrent hyperthyroidism
Exacerbation of hyperthyroidism
Jugular vein or carotid artery damage
Infection
Thoracic duct or lymph drainage damage

Source: From Ref. 2.

perhaps every three to six weeks, depending upon the clinical context. After several
months, when the thyroid function tests are decreased to the normal range and
the patient is being evaluated for possible hypothyroidism, serum TSH is also
determined. Most patients develop transient central hypothyroidism following
radioiodine therapy, with subnormal serum FT4 that usually, but not always,
evolves into permanent primary hypothyroidism (174). Approximately 5% to
10% of patients will require a second dose of 131 I, and approximately 1% of
patients require a third dose. It is prudent to wait 6 to 12 months for the full effects
of the initial dose to be manifest before another dose is considered. Overall,
approximately 50% of patients become permanently hypothyroid after one year,
with the rate being dependent on the radioiodine dose, with another 2% to 3%
developing hypothyroidism in the ensuing years. Thus, patients who do not become
hypothyroid early require lifelong follow up to monitor for the development of
hypothyroidism.
Prior to definitive therapy most patients will be given a beta-blocker when
the diagnosis of thyrotoxicosis is made, the goal being amelioration of symptoms
and a pulse rate ⬍100 beats per minute. Young patients who are mildly to moder-
ately thyrotoxic (e.g., minimal symptoms, otherwise healthy, minimal elevations
of T4 and T3) do not require methimazole or PTU therapy before or after 131 I treat-
ment, although they usually will continue their beta-blocker or be prescribed one.
Because of the possibility of worsening of thyroid function in the weeks following
radioiodine therapy (175,176), likely due to a transient increase in TSAb (177),
64 Burman and Cooper

older patients with moderate hyperthyroidism or patients with more severe disease
(e.g., presence of coexisting medical conditions, elevated FT4 and total T3 perhaps
two to three times above normal) are commonly given antithyroid agents before
or after radioiodine therapy to maintain normal thyroid hormone levels (178,179).
The drug is tapered as the thyroid function tests normalize, and then stopped when
the tests show hypothyroidism developing. When it is clear the patient is hypothy-
roid secondary to the 131 I rather than to the antithyroid agents, thyroxine is started.
Patients with very severe hyperthyroidism (e.g., marked symptoms, FT4 and T3
elevated more than three- to fourfold), elderly patients, or patients with known
or potential coexisting medical conditions such as cardiac disease will usually be
given antithyroid agents prior to 131 I therapy. It is hypothesized, but not proven,
that pretreatment with antithyroid drugs will decrease the likelihood of an exac-
erbation of thyroid function tests following 131 I therapy. Most endocrinologists
provide a window of about two to five days prior to and after 131 I therapy when
the patient does not receive antithyroid drugs.
In a survey of members of the American Thyroid Association, only about
one-third of respondents reported that they pretreat patients prior to administering
radioiodine (90). In meta-analysis of studies that examined outcomes of patients
either pretreated or not pretreated with antithyroid drugs prior to radioiodine (180),
new onset atrial fibrillation was reported in 1/660 (0.2%) patients pretreated with
ATDs versus 3/646 patients without ATDs (0.5%). Death after radioiodine was
reported in 1/660 (0.2%) pretreated with antithyroid agent and in 6/646 (0.9%)
without adjunctive drug therapy. Clearly, the routine use of antithyroid drugs
in this context is unnecessary and potentially exposes patients to drug toxicity.
Antithyroid drug pretreatment may interfere with the efficacy of radioiodine,
perhaps by acting as free radical scavengers within the irradiated gland (181–
184). Recent prospective, randomized studies have shown that PTU has a negative
effect on radioiodine outcome (185), but methimazole did not have such an effect
(186,187). Nevertheless, a meta-analysis of all studies of this question ascertained
that both antithyroid drugs lower the success rate whether if they are used before
or after radioiodine treatment (180). Therefore, if antithyroid drugs are to be used
before or after radioiodine therapy, it is recommended that the radioiodine dose
be increased by 25% to compensate for the potential radioprotective effect of
the drugs. One study has shown that thyroid hormone levels can be normalized
more quickly, and the cure rates with radioiodine-enhanced, with lithium therapy
(300 mg tid for six days beginning on the day of radioiodine therapy) (188).

Radioiodine and Graves’ Ophthalmopathy


131
I therapy is believed to exacerbate existing ophthalmopathy (189,190), at least
when it is more than minimally active on clinical grounds (191); thus it is impor-
tant to take a relevant history and perform a thorough ophthalmologic examination
(5). If there is moderately severe ophthalmopathy or if it seems to be progressive,
it is important to obtain an ophthalmology consultation, and an orbital computed
tomography (CT) scan (without contrast) or magnetic resonance imaging (MRI)
Hyperthyroidism 65

should be considered to evaluate the presence and extent of disease. These radi-
ologic studies may be required only in selected patients, but they do provide
quantitative, reliable information regarding proptosis, muscle size, and possible
optic nerve compression that can be used for comparative purposes later.
Tallstedt et al. (189) studied 168 patients with Graves’ hyperthyroidism
divided into age group 1 (20–34 years; n = 54 patients) and age group 2 (35–55
years; n = 114 patients). The patients in group 1 were randomly assigned to receive
either methimazole treatment for 18 months or subtotal thyroidectomy, while those
in group 2 were to receive either of these two treatments or, alternatively, 131 I
therapy. All patients were studied for at least 24 months. During the period of
evaluation, ophthalmopathy developed for the first time in 22 patients (13%) and
worsened in 8 patients (5%). The likelihood of the development or worsening of
ophthalmopathy was comparable among the patients in group 1 (medical therapy,
15%, and surgery, 11%). In group 2, ophthalmopathy developed or worsened in
10% of patients treated medically, in 16% treated surgically, and in 33% of patients
treated with 131 I (p = 0.02).
Bartalena (192) studied 443 patients with Graves’ hyperthyroidism who had
either no ophthalmopathy or minimal disease. Patients were randomly assigned to
receive radioiodine, radioiodine followed by a 3-month course of prednisone, or
methimazole for 18 months. The initial dose of prednisone was 0.4 to 0.5 mg/kg/
day starting two to three days after 131 I therapy and continuing for one month.
The dose was tapered and discontinued after two months. Patients were fol-
lowed closely, biochemically, and clinically for 12 months. In patients treated
with radioiodine alone (n = 150), ophthalmopathy developed or worsened in 15%
within six months after treatment. No patient in this group had an improvement
in ophthalmopathy. In marked contrast, in the group treated with radioiodine and
prednisone, 50 of the 75 (67%) with ophthalmopathy at baseline had improvement,
and no patient had progression. Of the 148 patients treated with methimazole, three
(2%) who had ophthalmopathy at baseline improved, four (3%) had worsening
of eye disease, and the remaining 141 had no change. These data clearly show
that prednisone therapy can help prevent radioiodine-associated deterioration in
ophthalmopathy. Nonetheless, it remains unclear which patients should receive
steroids. Ideally, steroid therapy should be considered as a team decision among
the patient, family, endocrinologist, and ophthalmologist.
Taking into account various regimens in the literature, one reasonable
approach in patients with moderate or severe ophthalmopathy is to administer
prednisone in doses of 40 to 60 mg prednisone daily starting several days prior to
131
I therapy and continuing for several weeks, trying to taper the drug completely
by six to eight weeks. Corticosteroid therapy is reserved for patients with moderate
or severe ophthalmopathy. Tallstedt et al. (189,193) noted that the initial serum
T3 level was an independent risk factor for the development of ophthalmopathy
and that postablative hypothyroidism should be avoided. Bartelena et al. (192)
also showed that cigarette smoking was a potent, independent risk factor for the
worsening of ophthalmopathy after radioiodine therapy, and patients with Graves’
66 Burman and Cooper

disease who smoke should be advised to discontinue. Some endocrinologists avoid


radioiodine therapy in patients with moderate to severe eye disease unless it has
been stable for at least one year.

Radioiodine and Cancer


A number of older studies have failed to show any causal relationship between
radioiodine therapy and the subsequent development of thyroid cancer, leukemia,
or other malignancies. Ron et al. (194) retrospectively analyzed 35,593 hyperthy-
roid patients (91% had Graves’ disease) who had been treated with radioiodine,
antithyroid drugs, or surgery between 1946 and 1954. Some 65% of these patients
were treated with radioactive iodine, thus allowing long-term comparison of
results between various therapeutic modalities. When studied in December 1990,
about half of the patients had died. The total number of cancer deaths in all
hyperthyroid patients was comparable in patients treated with radioactive iodine
and those treated with surgery or antithyroid agents, although there was a slight
excess of cancer-related mortality from lung, breast, kidney, and thyroid. Starting
at least one year after treatment, an enhanced risk of cancer mortality was also
seen in hyperthyroid patients treated with antithyroid drugs. After more than
five years following therapy, radioactive iodine therapy was associated with
an increased risk of thyroid cancer mortality, but only in patients with toxic
multinodular goiter. Overall, the risk of thyroid cancer in patients treated with
radioactive iodine treated resulted in a small, absolute excess in actual patient
deaths.
Franklyn et al. (195) also suggested that 131 I therapy was associated with
a higher incidence of thyroid cancer. They retrospectively studied 7417 patients
treated in Birmingham, England, with radioiodine for Graves’ disease. On ana-
lyzing 72,073 person-years of follow-up, 634 cancer diagnoses were found as
compared with an expected number of 761. The relative risk of cancer mortality
was also decreased. The incidence of cancers of the pancreas, bronchus, trachea,
bladder, and lymphatic and hematopoetic systems was decreased. However, there
were significant increases in incidence and mortality for cancers of the small
bowel and thyroid, although the absolute risk of these cancers was small (195).
In this English study, the goal of radioiodine therapy had been euthyroidism
rather than hypothyroidism. The destruction of all residual thyroid tissue would
be expected to lessen possibility of thyroid cancer. Finally, in one recent pop-
ulation based study of 3888 patients treated for hyperthyroidism, there was no
increase in cancer mortality over eight years of follow up, but the method of treat-
ment was not specified (196). In another similar study of 2973 patients treated
with radioiodine with a nine year follow up, there was an increase in overall and
cancer mortality (adjusted RR 1.36; 95% CI 1.12–1.65), mainly due to cancer
of the stomach and esophagus. However, the increase in total mortality was only
seen in patients with toxic multinodular goiter, but not in patients with Graves’
disease.
Hyperthyroidism 67

Thyroidectomy for Graves’ Disease


A total or near total thyroidectomy is also a reasonable therapeutic option for
selected patients with Graves’ disease (90,197–201). This therapy is reserved for
patients who are not well controlled on antithyroid agents; who have a particu-
lar reason for surgery, for example, a very large goiter or a thyroid nodule with
suspicious aspirate; and those who prefer this therapy after a careful considera-
tion of each option. Thyroidectomy must be performed by an experienced thyroid
surgeon, although there is still a risk of temporary or permanent hypocalcemia
secondary to injury of the parathyroid glands. Hoarseness may also occur as a
result of injury to the recurrent laryngeal nerve. These two serious complications
occur in approximately 1% of patients treated surgically (202). Other complica-
tions of surgery are rare and listed in Table 4. In most circumstances, a euthyroid
state should be established prior to surgery with antithyroid agents, although more
rapid preparation with beta-adrenergic blocking drugs and saturated solution of
potassium iodide (SSKI) has been used as well (203). If there is no immediate need
for surgery, antithyroid drug therapy is probably safer, as postoperative fever and
tachycardia more commonly develop after preparation with beta-blockers alone
(204). When surgery must be done quickly, a 5-day regimen of propranolol, glu-
cocorticoids, and oral cholecystographic agents sodium ipodate or iopanoate have
been used (205). Unfortunately, ipodate and iopanoate are not currently available in
the United States. It is preferable that patients’ thyroid function tests be normal for
several weeks prior to surgery, as this depletes intrathyroidal hormone stores and
make releases at the time of surgery less likely. SSKI (10 drops three times a day)
given daily for 10 days prior to surgery has been shown to decrease intraoperative
blood loss in a recent randomized trial (206). Despite optimal care, perioperative
thyroid storm may still occur, and patients should be treated aggressively if signs
or symptoms are present (207). Obviously, most patients become hypothyroid
immediately following surgery and require lifelong thyroxine therapy.
Occasionally, a patient who has had a portion of his or her thyroid gland
surgically removed in the past for Graves’ disease will present again with hyper-
thyroidism (208). The likelihood of this occurring depends upon the extent of
thyroidectomy and the titers of TSH receptor–stimulating antibodies. Relapses
may occur many years following surgery, often when the patient has been taking
L-thyroxine for previously diagnosed hypothyroidism. Patients typically have a
gradual development of symptoms and signs of hyperthyroidism in conjunction
with elevated serum FT4 and T3 levels. When the exogenous thyroxine is decreased
over time, the hyperthyroidism persists. Once the patient has discontinued thy-
roxine for at least four to six weeks and hyperthyroidism is still present, further
evaluation is required. Measurement of TSH receptor antibodies also may be help-
ful in this circumstance. In the absence of palpable thyroid tissue, determination of
serum thyroglobulin levels and performance of a 24-hour radioiodine uptake will
help distinguish iatrogenic or factitious hyperthyroidism from recurrent Graves’
disease (3).
68 Burman and Cooper

Choice of Therapy for Graves’ Disease: Summary


Antithyroid drugs are a reasonable choice for first-line therapy in patients with
small goiters and mild disease or in those whose TSI levels are normal. Also,
children and adolescents are traditionally treated initially with antithyroid drugs.
Because of concern that radioiodine can worsen underlying ophthalmopathy
(189,190), some clinicians recommend antithyroid drugs as the treatment of choice
in patients with significant eye disease even when the biochemical abnormalities
are more severe. In the future, costs of care may become more important in the
management of hyperthyroidism. In studies that examined costs of care, radioio-
dine therapy was the least expensive alternative compared to antithyroid drugs
and surgery (209,210). A computer-simulated, cost-effectiveness analysis reached
the same conclusion, but when the chances of remission with antithyroid drugs
was greater than 50%, they became as cost-effective as radioiodine (211). Patient
satisfaction is also an important outcome that has only recently been studied. In the
one report in which it was measured, antithyroid drugs, radioiodine, and surgery
were equal in terms of patient satisfaction, whether a particular therapy would be
recommended to friends or relatives, and in patients’ concern about possible side
effects (96).

TREATMENT OF GRAVES’ OPHTHALMOPATHY AND


PRETIBIAL MYXEDEMA
Although almost all patients with Graves’ disease have radiologic evidence of eye
muscle involvement, only approximately 30% of patients have obvious clinical
disease (212). Several different methods for assessing disease severity have been
described, but none of the classifications is perfect. The “NOSPECS” classification
is still in wide use today (Table 5) (213), although it has been criticized for
being overly subjective and not quantitative enough (214). In general, the most
common symptoms are related to soft tissue swelling due to orbital congestion,
with irritation, tearing, burning, and a gritty sensation in the eyes. Diplopia is
a more unusual and debilitating problem, and only rarely is vision threatened
because of corneal exposure or optic nerve involvement. Symptoms and cosmetic
concerns also significantly impact negatively on the quality of life of affected
patients (215).
The symptoms and signs of Graves’ ophthalmopathy are due to orbital
inflammation, with the extraocular muscles and/or retro-orbital fibroblasts being
the target of the autoimmune reaction (6,13,216–218). Glycosaminoglycans pro-
duced by fibroblasts responding to T-cell infiltration cause edema of the extraocular
muscles, and further expansion of the retro orbital tissues is due to increased orbital
fat (214). Ultimately, fibrosis of the extraocular muscles can lead to diplopia, and
severe enlargement of the muscles can cause an ischemic optic neuropathy due to
compression of the optic nerve as it exits the apex of the orbit.
Hyperthyroidism 69

Table 5 ‘NO SPECS’ Classification of Eye Changes of Graves’ Disease


Class Definition

O No physical signs or symptoms


I Only signs, no symptoms (e.g., upper lid retraction, stare, and eyelid lag)
II Soft tissue involvement (symptoms and signs)
III Proptosis
IV Extraocular muscle involvement
V Corneal involvement
VI Sight loss (optic nerve involvement)

The primary autoimmune target in Graves’ ophthalmopathy is unknown,


but probably is a cross-reacting antigen or antigens that are present in both in the
orbit and the thyroid gland (216). The TSH receptor has been hypothesized be
the putative common antigen. TSH receptor transcripts have been isolated from
extraocular muscle using PCR (219), TSH receptor protein has been identified in
orbital tissues by immunostaining (219), and the levels seem higher in Graves’
ophthalmopathy patients as compared to normal subjects (13). Further, autoanti-
bodies directed against TSHR or the insulin-like growth factor-1 (IGF-1) receptor
may play a role in the development or progression of Graves’ ophthalmopathy
(13).
Since the cause of Graves’ ophthalmopathy is unknown, treatment is directed
at symptoms. In most patients, the problem is self-limited, often resolving as the
hyperthyroidism is treated (220). Although most experts feel that it is best for the
patient to be euthyroid, there is no consistent relationship between a patient’s thy-
roid function and progression or regression of eye disease. There is good evidence
that smoking exacerbates Graves’ ophthalmopathy (221), and it is suggested that
smoking cessation has a beneficial effect (222). The mechanism by which smoking
affects Graves’ ophthalmopathy is unknown. There is solid evidence that radioac-
tive iodine therapy can exacerbate Graves’ eye disease when it is moderately severe
at baseline (189,190). When the condition is mild, symptoms such as irritation,
tearing, and photophobia are easily treated with artificial tears and lubricating eye
ointments. In more severe cases, high doses of glucocorticoids (e.g., 60–120 mg
of prednisone today) usually will result in prompt improvement in local symp-
toms and ocular motility (5). Unfortunately, as the glucocorticoid is tapered, the
ophthalmopathy often flares up, so that other measures are sometimes needed.
The use of long-acting octreotide in the treatment of Graves’ ophthalmopathy has
also been recently studied and two studies have noted no or minimal improvement
(217,223).
Orbital radiotherapy is usually the next step after glucocorticoid therapy
(5). Although somewhat controversial, orbital radiotherapy has been shown to
be effective in several studies, including a recent randomized prospective trial in
which half of the patients received sham irradiation (224). In this study, 60% of
70 Burman and Cooper

irradiated patients improved versus 31% of sham irradiated patients. However,


Gorman et al. (225) also performed a prospective, randomized, double-masked,
internally controlled, clinical trial of external beam radiotherapy for patients with
mild to moderate Graves’ ophthalmopathy. When analyzed 6 to 12 months after
orbital radiation, there was no apparent clinical benefit identified. As a result
of these recent studies, the role of external orbital radiation in the treatment of
Graves’ ophthalmopathy is uncertain.
Patients who fail radiation therapy may require surgical decompression of
the orbit if there is rapidly progressive visual loss due to optic neuropathy, steroid
dependence, or continuing ocular motility problems. Surgery is also indicated
to correct self-perceived cosmetic problems, particularly in patients with severe
proptosis or lid retraction. Unfortunately, orbital decompression often results in
more significant ocular motility problems that then require additional strabismus
surgery for correction (226–229).
The cause of pretibial myxedema remains obscure (230), but it probably
shares common features with Graves’ opthahlmopathy, including lymphocytic
infiltration and a response by fibroblasts to the subsequent inflammation. The usual
treatment is topical steroid cream with or without occlusion (231). Intralesional
steroids have also been used. Recently, octreotide has been reported to be of use
(232), but additional data are needed before this therapy can be recommended
outside of a clinical trial. Recent trials have suggested the efficacy of combined
pentoxifylline and intralesional triamcinolone acetonide (233,234).

SUBCLINICAL HYPERTHYROIDISM
Subclinical hyperthyroidism is a term generally utilized to describe patients with
normal serum total and FT4 and T3 levels and a decreased serum TSH level (235–
257). A recent consensus panel has published useful guidelines for approaching
and treating patients with subclinical hyperthyroidism (258). Patients most fre-
quently have undetectable TSH levels, although a TSH that is subnormal but still
detectable also meets the definition assuming that the values are consistent over
time. The serum TSH level should be measured in a third-generation assay that
is capable of discriminating degrees of low values, and the patient must not be
taking or receiving any medications known to alter the hypothalamic-pituitary axis
(such as corticosteroids and dopamine). TSH can be suppressed physiologically
in the first trimester of some pregnant patients, and the patient must have a normal
pituitary-thyroid axis (258). Further, the patient must be relatively healthy, with-
out serious systemic diseases, since the “euthyroid sick syndrome” will also affect
the serum TSH level. The patient with subclinical hyperthyroidism typically does
not have significant signs or symptoms of hyperthyroidism, such as weight loss,
nervousness, or palpitations. Some of the signs and symptoms of overt hyperthy-
roidism are vague and nonspecific and it may be difficult to determine if a patient
is really asymptomatic. In a young patient with subclinical hyperthyroidism due
to Graves’ disease, the thyroid is either not palpable or mildly enlarged, but a
Hyperthyroidism 71

multinodular goiter is often palpable in older patients with multinodular goiters.


Although the frequency of this disorder is not well defined, its recognition has
increased in part due to the advent of more sensitive TSH assays. When patients
with known thyroid disease are excluded, the incidence of subclinical hyperthy-
roidism is estimated to be approximately 2% (258). More details are available in
chapter 10 “Practical Management of Thyroid Disease in the Elderly.”
Sawin et al. (259) determined that low serum thyrotropin concentrations
are a risk factor for subsequent atrial fibrillation. They studied 2007 persons (814
men and 1193 women) older than 60 years during a 10-year follow-up period
to determine how frequently atrial fibrillation would develop. When analyzed
cumulatively over the entire study period, atrial fibrillation occurred in 28% of
the subjects with low serum thyrotropin values (≤0.1 ␮U/mL), as compared
with 11% among those with normal or slightly decreased serum TSH values. The
estimated relative risk for atrial fibrillation was 3.1 in Sawin’s study (259), and this
elevated risk has been confirmed in other studies (255). Although the most frequent
cause of the suppressed TSH levels in the patients in this study was exogenous
thyroid hormone therapy, there is no reason to think that the effects of endogenous
hormone would differ. Some studies have shown that postmenopausal women with
subclinical hyperthyroidism have lower bone mineral density as compared with
age-matched control women (260). However, others have not found this to be the
case (261).

Diagnosis
The evaluation of a patient with subclinical hyperthyroidism may vary among
physicians, but a reasonable approach in a typical asymptomatic patient is to
perform a complete history and physical examination, ensure the thyroid function
tests are measured in appropriately sensitive assays, and repeat the thyroid function
tests monthly for three months. Thyroid function tests should include FT4 and FT3,
since occasionally the free hormone levels may be increased disproportionately
compared to the total hormone levels (258,262). If atrial fibrillation, cardiovascular
disease, or other significant medical illnesses are present, earlier diagnosis and
treatment are appropriate, so it is recommended to repeat thyroid function tests
over a shorter period of time, such as two weeks (258). The suppressed TSH
level may represent the initial manifestations of hyperthyroidism that will evolve
into a more overt form over the ensuing months. Alternatively, the suppressed
TSH may be transient and return to the normal range, as a suppressed TSH
could represent an episode of transient thyroiditis. If this were correct, the TSH
would be expected to return to normal within three months. Therefore, to ensure
stability of the laboratory tests and help exclude the possibility of a laboratory
error, thyroid function tests are determined monthly for three months prior to
further evaluation. If stability in TSH levels is demonstrated, it seems reasonable
to perform a radioactive iodine uptake test and to consider measuring TSH receptor
antibody levels. Frequently, the radioactive iodine uptake is at the upper range of
72 Burman and Cooper

normal or slightly higher, and TSI (and TBII) levels are also minimally elevated
or normal, reflecting the minimal degree of hyperthyroidism. The radioactive
iodine uptake is measured not only to assess the level of thyroidal activity but
also to exclude painless thyroiditis. A thyroid sonogram may be useful in selected
patients to help quantitate thyroid gland size and to determine the presence and
characteristics of thyroid nodules.

Treatment
Once the tests are shown to be consistent with persistent subclinical hyperthy-
roidism, treatment options must be discussed. The first option is simply to continue
to monitor the patient and thyroid levels indefinitely. This is a reasonable option in
some patients, especially young premenopausal women and young men, but—as
noted above—subclinical hyperthyroidism may be associated with a higher risk
of accelerated bone loss and osteoporosis in postmenopausal women. It addition,
as noted above, it may also be associated with a higher risk of atrial fibrillation,
especially in older individuals (259,263,264). The second therapeutic option is to
administer radioactive iodine in an effort to induce permanent hypothyroidism.
Radioactive iodine is considered safe and effective and is the usual treatment
of choice for patients with overt hyperthyroidism. However, this treatment may
appear drastic for most patients with subclinical hyperthyroidism, in part because
they are asymptomatic and often do not desire definitive therapy. Similar comments
apply to the recommendation of a thyroidectomy. Given these considerations, a
6- to 12-month trial of antithyroid agents is a reasonable medical approach. It
normalizes the TSH, FT4, and FT3 levels, and yet also takes into account patients’
natural reluctance to have definitive therapy when they are asymptomatic and
when treatment directives revolve around a single laboratory test abnormality
(i.e., serum TSH). It is uncertain if most patients with subclinical hyperthyroidism
have multinodular goiters or Graves’ disease. Therefore, the remission rates with a
trial of antithyroid agents are unknown. One approach to a patient with subclinical
hyperthyroidism is to treat with low-dose antithyroid medication (e.g., methima-
zole 5–10 mg/day) for an arbitrary time period, such as 6 to 12 months (265). This
course of action might induce a long-lasting remission in a patient with Graves’
disease and, if the disorder subsequently recurs, the patient may be more likely to
accept definitive therapy. While a patient is being treated with antithyroid agents,
serum FT4, TSH, and possibly T3 or FT3 should be measured periodically. The risk
of adverse effects from the antithyroid agents, such as agranulocytosis, liver func-
tion abnormalities, and skin rash, is low given the small dose of medication used.
Mudde et al. (260) and Faber (266) have shown a benefit of antithyroid agent ther-
apy on bone density values in patients with subclinical hyperthyroidism. Buscemi
et al. (267) showed that restoration of euthyroidism with antithyroid agents had a
favorable effect on cardiac and bone paratmeters. However, more detailed, prospec-
tive studies are required regarding all aspects of subclinical hyperthyroidism to
allow a more evidence-based approach to these patients.
Hyperthyroidism 73

THYROID STORM
The term thyroid storm refers to severe and exaggerated symptoms and signs of
hyperthyroidism, usually in association with tachycardia, fever, diarrhea, vom-
iting, dehydration, disorientation, or mental confusion (207,268). Patients usu-
ally experience severe restlessness and anxiety and may be unable to reason.
There is a continuum between “routine” hyperthyroidism and thyroid storm,
and different observers may vary in their definition of thyroid storm. Although
there have been attempts to establish a uniform set of criteria for the diagnosis
of thyroid storm (207), these have not been generally accepted. Thyroid func-
tion tests, for example, FT4 and FT3, overlap between routine hyperthyroidism
and thyroid storm, and mean values are similar in most studies (269). Sher-
man and Ladenson (270) have noted lower socioeconomic status in patients
with thyroid storm compared to those with controlled hyperthyroidism, indi-
cating that lack of access to medical care may be an important predisposing
factor.
Thyroid storm is typically precipitated by a specific event, such as surgery
(especially patients having thyroid surgery without adequate preparation), severe
systemic illness (e.g., pneumonia, pharyngitis), or parturition. It is important to
prevent thyroid storm whenever possible by trying to predict circumstances in
which it may occur. It is preferable, in general, to treat patients as if they had
thyroid storm when it is suspected, rather than to delay therapy in the hope that
thyroid storm will not develop or become fully manifest.

Treatment
Patients should be treated with antithyroid agents to restore euthyroidism prior
to anticipated stressful events, such as surgery. As noted earlier, hyperthyroid
patients should be prepared with antithyroid medications for several weeks prior
to thyroidectomy. It is generally believed that when thyroid hormone stores are
decreased the risk of exacerbating thyroid storm is decreased, since there may be
less preformed T3 and T4 that may be released into the circulation at surgery.
However, Hermann et al. (271) studied thyroid hormone venous effluent from
the thyroid gland in patients with hyperthyroidism undergoing thyroidectomy
without antithyroid agent pretreatment. There were no significant differences in
T3 and T4 levels in the venous effluent compared to the peripheral levels. It is
unclear if, however, if this study applies to all hyperthyroid patients undergoing a
thyroidectomy and it is prudent to perform careful monitoring at the time of surgery
with an experienced surgical and anesthetic team. The precise pathophysiology
underlying a possible increase of thyroidal release of T4 and T3 around the
time of thyroidectomy is not known but likely relates to surgery and anesthesia
itself.
Treatment for a patient with severe thyrotoxicosis or thyroid storm is more
aggressive than that for a patient with less severe thyroid dysfunction (207). The
doses of medications are higher in patients considered to have thyroid storm.
74 Burman and Cooper

Propylthiouracil, 100 to 200 mg q4 h, or methimazole, 10 to 20 mg q4 h, is


recommended in conjunction with propranolol, 60 to 80 mg q8 h. In unusual
circumstances when patients cannot be administered oral medication effectively
PTU or methimazole can be given rectally or, in the case of methimazole, intra-
venously (272–275). Propanolol can also be given intravenously (2–5 mg every
four hours), and if there is a history of pulmonary disease, esmolol, at a dose of
50–100 ␮g/kg/min can be used. Hydrocortisone 100 mg q8 h is added to ensure
adequate adrenal function, and an iodine-containing agent may also be employed.
An intravenous preparation of sodium iodide is no longer available. SSKI, 5 drops
tid, or Lugol’s solution, 5 drops tid, contain sufficient iodine to reduce thyroid
production and secretion of T4 and T3, and their effects can be seen within several
days (3). These agents are very useful and effective but can have a significant
negative impact as well. Iodine-induced inhibition of thyroid gland synthesis (the
Wolff-Chaikoff effect) and glandular secretion persist for one to two weeks, espe-
cially in previously untreated patients with severe hyperthyroidism. Then there
may be “escape” from the inhibitory effects, with enhanced synthesis and release.
The mechanism of this effect relates to modulation of function of the sodium
iodide symporter (212). This biphasic effect of iodine is extremely important
and is the chief reason why hyperthyroid patients are not routinely treated with
iodine-containing drugs. Two oral cholecystographic agents, ipodate (Oragraffin)
and iopanoate (Telepaque), have been found to contain sufficient iodine to have
effects similar to Lugol’s solution and SSKI, but, in addition, both potently inhibit
T4-to-T3 conversion (276). Unfortunately, both ipodate and iopanoate have been
withdrawn from the commercial market and both are unavailable. Perchlorate
inhibits the sodium-iodide symporter and decreases intrathyroidal iodine which
is needed for sustained thyroid hormone synthesis (167). Perchlorate also is not
presently available commercially. Cholestyramine (or colestipol) administration is
mildy effective as adjunctive therapy as it decreases the enterohepatic circulation
of T4 and T3 (277). Care must be exercised to separate the interval between the
administration of cholestyramine (and colestipol) and other oral medications.

SOLITARY TOXIC NODULES


Introduction
A solitary autonomous toxic thyroid nodule produces sufficient thyroid hormone to
suppress TSH and cause overt or subclinical hyperthyroidism (278–282). Often,
a thyroid nodule may be autonomous and not yet be sufficiently functional to
suppress serum TSH. The capacity to secrete thyroid hormones varies with the
size of the thyroid nodule, and those that are ⬎3 cm in size are much more likely
to produce hyperthyroidism (283). The percentage of autonomous nodules that
secrete sufficient thyroid hormones to produce overt hyperthyroidism is relatively
low, in the range of 20%.
Hyperthyroidism 75

Pathology
Histologically, autonomous nodules are cellular, follicular adenomas with fre-
quent hemorrhage, fibrosis, calcification, and cysts. There is often a dense, fibrous
capsule. The vast majority of solitary autonomous thyroid nodules are benign. In
adults, as many as several percent of these nodules contain foci of papillary thy-
roid cancer, whereas in children and adolescents, the percentage of autonomous
nodules that contain thyroid cancer is likely higher (284–287). This reported
pathologic frequency is higher than found in clinical practice and may repre-
sent abnormalities in adjacent tissue rather than the adenoma itself. A fine-needle
aspiration biopsy of the nodule should be performed if there is any suspicion of
cancer based on clinical, historic, or laboratory studies. For example, the pres-
ence of associated cervical lymphadenopathy, recent growth (which could simply
represent hemorrhage), or a history of neck radiation may increase the suspi-
cion for associated cancer. The type of thyroid cancer found in this circumstance
is usually papillary thyroid cancer, although other types, such as follicular or
medullary thyroid cancer, may be identified. Earlier studies suggested that cytol-
ogists would read aspirates from autonomous nodules as suspicious for follicular
thyroid cancer, given the cellular nature of these nodules (288). However, subse-
quent analyses show that the fine-needle aspirations are rarely confusing and that
the majority of autonomous nodules yield benign cells in the presence of colloid
(289,290).

Pathogenesis
Autonomous function of the nodule is attributed to either a loss in suppression of
normal cell function due to a genetic defect in the TSH receptor or a downstream
pathway (e.g., G protein), causing excess stimulation of the thyroid cell and unreg-
ulated thyroid hormone production. The TSH receptor consists of seven transmem-
brane domains with three extracellular and three intracellular loops (291). The
TSH receptor is coupled to a G protein, which subsequently initiates cyclic AMP
production. Cyclic AMP is essential for production of thyroid cell products such
as thyroglobulin, thyroid peroxidase, and hormone production and secretion, as
well as cell growth and proliferation. The majority of solitary autonomous thyroid
nodule tissue contains mutations in the TSH receptor protein or, less often, in the
stimulatory G protein (292–295). Several different mutations in the TSH receptor
gene and the resultant mutated protein are associated with cyclic AMP–dependent
transcription independent of the presence of TSH. Mutated receptors retain a sig-
nificant response to TSH stimulation, although it is decreased when compared to
normal thyrocytes. The Gs protein regulates cell growth; when it becomes mutated,
this protein acts as an oncogene, leading to abnormal cellular regulation. Gs muta-
tions have been found in approximately 5% to 30% of toxic thyroid nodules
(293).
76 Burman and Cooper

Clinical Considerations
Approximately 1% of patients referred for thyroid disease and approximately
5% of hyperthyroid patients have an autonomous thyroid nodule (283). Patients
with autonomous functioning thyroid nodules present most frequently with a neck
mass but may also have subclinical hyperthyroidism (278,283). Overt symptoms
of hyperthyroidism are uncommon. The frequency of toxic adenomas increases
with age, and only about half of patients with autonomous nodules over the age of
60 manifest clinical signs or symptoms of hyperthyroidism. Autonomous nodules
are much more common in women. A solitary autonomous thyroid nodule is not
an autoimmune disease and is not associated with ophthalmopathy or dermopathy.
Patients with an autonomous nodule larger than 3 cm have approximately a 20%
chance of progressing to overt hyperthyroidism over several years, with smaller
lesions having a much lower progression rate (296). Toxic nodules may undergo
cystic or necrotic degeneration with return to euthyroidism (278). As many as 20%
to 30% patients with solitary autonomous nodules may have a restoration of normal
function secondary to hemorrhage (282,297). Iodine deficiency increases the risk
of iodine-induced hyperthyroidism in patients with autonomous thyroid nodules
(see Chapter 3, “Hyperthyroidism”). Even a small addition of iodine, perhaps
100 ␮g/day, to a low-iodine diet can initiate hyperthyroidism (298). Although
iodine deficiency is not a problem in the United States, hyperthyroidism has been
reported in patients with autonomous thyroid nodules within one to two months
after exposure to radiocontrast dyes as well as after exposure to iodine-containing
drugs such as amiodarone (299).

Diagnosis
Autonomous thyroid nodules are usually diagnosed by integration of the history
and physical examination with laboratory and nuclear medicine testing. The serum
TSH level is usually suppressed and is often unmeasurable. Free/total T4 and T3
are usually normal or slightly elevated but may vary depending upon nodule size
and iodine exposure. Preferential T3 secretion, that is, T3 toxicosis, appears more
common in patients with autonomous nodules than in Graves’ disease (262,296).
The identification of an autonomous nodule on technetium-99 m or (123 I) scinti-
graphic scanning is the sine qua non of the diagnosis. When a radionuclide scan
is performed, the solitary autonomous thyroid nodule represents the only tissue
that appears to be trapping radioactive iodine, with the remainder of the thyroid
tissue being suppressed. However, an abnormal thyroid scan alone does not prove
autonomous function, since there are technical factors that could make a dom-
inant nodule appear to be the sole area that traps radioactive iodine. When an
autonomous thyroid nodule is suspected, it may be appropriate to use radioiodine
rather than technetium as the diagnostic agent. Occasionally, a nodule may appear
to be “hot” with technetium, when it is actually a “cold” nodule with radioiodine.
This discordance is thought to relate to the ability of iodine to be trapped and
organified by thyroid tissue, whereas technetium can only be trapped (300).
Hyperthyroidism 77

Autonomous nodules lack an intact feedback mechanism to control iodine


uptake and hormone production. Therefore, several tests have been used in the
past to help diagnose the presence of autonomy. The T3 suppression test consists
of scintigraphic scanning of the thyroid before and after 10 days of exogenous
oral T3 administration at 50 to 100 ␮g/day. In addition, TSH, FT4, and FT3 are
drawn at baseline and on completion of the T3 dosing. A normal patient will have
suppression of endogenous TSH production; the second thyroid scan will show
no thyroid uptake and the FT4 level will have decreased. In contrast, a patient
with an autonomous thyroid nodule will show unchanged uptake over the nodule,
and the FT4 level will not have changed. T3 suppression testing in hyperthyroid
patients, especially elderly subjects, is fraught with potential danger by potentially
exacerbating already existing thyrotoxicosis. This test is rarely performed and its
use should be discouraged with the advent of third-generation TSH assays.

Treatment
Therapeutic decisions regarding autonomous nodules depend upon a variety of fac-
tors, including patient age, the severity of hyperthyroidism, and associated medical
conditions such as coronary artery disease and/or a history of cardiac arrhythmias.
Patients with overt hyperthyroidism in the presence of an undetectable TSH and
elevated FT4 and T3 should be treated. Older patients with undetectable TSH and
normal FT4 and T3 should be strongly considered for treatment because of possi-
ble deleterious effects on bone metabolism and the heart (see section “Subclinical
Hyperthyroidism” of this chapter). More controversial is whether young patients
with subclinical hyperthyroidism or people with nodules ⬎3 cm in diameter
should be treated prophylactically. Some patients desire treatment for cosmetic
reasons.
The most important therapeutic option is radioactive iodine ablation of the
toxic adenoma. Radioactive iodine concentrates in the autonomous nodule and may
accumulate in extranodular tissue, albeit to a much lesser extent. Although results
are variable, based upon size of the nodule, nodule uptake, dose administered,
and tissue radiation sensitivity, a typical toxic nodule will shrink approximately
40% within one year following an ablative dose. Radioactive iodine ablation of
an autonomous thyroid nodule may decrease the ability of that nodule to secrete
excess thyroid hormones; but the nodule itself, although functionally inactive,
may remain and must be monitored clinically over time (278,283). The beneficial
effects of radioiodine may not be maximal until 4 to 12 months after therapy.
Thyroid function may normalize or decrease to below the normal range in this
time period, but the nodule may still be palpable in perhaps half the patients.
Occasionally, a second or even a third dose of radioiodine is required to render
a patient euthyroid. It is recommended to wait at least six months after the ini-
tial dose before considering retreatment with radioiodine. Persistent or recurrent
hyperthyroidism can be expected in approximately 10% of patients, with studies
revealing a range of 0% to 41.3% (279,301,302).
78 Burman and Cooper

Following radioiodine therapy, hypothyroidism occurs at an average of


approximately 10% within the first year, with an annual rate thereafter of approx-
imately 3% (297,303,304). This occurrence rate may relate to the fact that the
contralateral normal tissue is exposed to significant amounts of radiation (279).
Underlying autoimmune thyroid disease may also predispose to the development
of hypothyroidism. Larger doses of 131 I appear to be a causative factor in stud-
ies finding a substantial rate of hypothyroidism. The dose per gram of thyroid
tissue, percent radioactive iodine uptake, pre- and/or posttreatment with antithy-
roid medications, and recent iodine exposure all may affect the outcome. In a
retrospective analysis of 45 patients with toxic adenomas receiving an average of
10.3 ± 3.5 mCi of 131 I, euthyroidism was achieved in 91% and 93% at 6 and 12
months, respectively (302), and no patients became hypothyroid. Three patients
had recurrent hyperthyroidism between 4.5 to 10 years and became hypothyroid
after a second therapy. However, two other studies found that 35% and 36% of
patients became hypothyroid following the first dose of 131 I, with an average
follow-up of 3.8 years and 8.5 years, respectively (303,305). Of note, these two
studies used a mean 131 I dose of 29.1 mCi and 23 ± 10 mCi. Higher doses are
associated with more prompt achievement of euthyroidism or hypothyroidism.
131
I is tolerated well by the majority of patients, although radiation thyroiditis
with or without exacerbation of the underlying thyrotoxicosis can occur. Repeat
scintigraphic scanning following radioactive iodine therapy may reveal relatively
normal thyroid distribution of tracer, the nodule may be cold, warm, or hot in
relation to the other tissue, or the autonomous nodule may remain evident with
continued suppression of extranodular tissue. In the latter case, the serum TSH
level will remain suppressed.
Surgery can also be performed to remove the affected lobe and isthmus,
leaving the remaining lobe intact (278,281,306). In experienced hands, surgery
has a low anesthetic risk and a low surgical complication rate of vocal cord
paralysis and hypoparathyroidism, especially if only a lobectomy rather than
a thyroidectomy is performed (306). Removal of the nodule alone or only a
portion of the affected lobe including the nodule is not recommended. Recurrence
following surgery is unusual. Multifocal autonomous nodules not appreciated on
the initial evaluation, and therefore not surgically removed, might be responsible
for some failures. Euthyroidism can be achieved relatively quickly with surgical
intervention, although hypothyroidism may occur. The hypothyroidism may be
secondary to inadequate residual thyroid tissue or possibly underlying autoimmune
thyroid disease. Although controversial, some experts recommend that patients
who have had a lobectomy and isthmusectomy be placed on lifelong thyroxine
therapy. The use of thyroxine therapy has not been proven to be effective in
decreasing growth of thyroid nodules on the contralateral side. However, following
a thyroid lobectomy, the risk of developing hypothyroidism is relatively high,
being 14% in one study (307). Hedman et al. (307) followed up 95 patients an
average of 15 years after thyroid lobectomy had been performed. Nine percent
of patients had an elevated TSH with normal serum thyroxine and an additional
Hyperthyroidism 79

5% of patients had hypothyroidism with elevated TSH and decreased T4. This
complication can develop in the immediate postoperative period or may develop
after many ensuing years. Periodic clinical examinations and thyroid function
tests are needed, regardless of whether or not an individual patient is receiving
L-thyroxine therapy.
Percutaneous ethanol injection (PEI) is an alternative therapeutic option in
patients with an autonomous thyroid nodule. Ethanol (95% to 98%) is injected
into the autonomous nodule using a 22-gauge needle under sonographic guidance
(308,309). Special attention is given to avoiding leakage of ethanol into the extra-
nodular tissue, which, should it occur, may cause serious complications. Other
potential complications of PEI include pain at the local injection site, dysphonia
(usually transient), exacerbation of thyrotoxicosis, fever, and hematoma. A case
of severe toxic necrosis of the larynx and associated necrotic dermatitis has been
reported following ethanol injection of a thyroid nodule (310). Nodules ⬎30 mL
in volume are more resistant to this therapy, having almost half the cure rate of
smaller nodules. Centers experienced in this modality note that approximately
four to eight treatments may be required to achieve success. The total volume of
ethanol delivered is usually one and a half times the nodule volume, and 1 to 8 mL
is administered in weekly sessions over a 2- to 12-week time frame. A multicenter
study (308) reported a success rate six months after PEI of 77.5% and 61.1% in
toxic adenomas and “pretoxic” adenomas, respectively. By 12 months, rates of
83.4% and 66.5% can be expected for toxic and pretoxic adenomas. Successful
resolution of hyperthyroidism and nodule autonomy can be expected between 3
and 12 months following PEI, with the majority achieving “cure” within the first
6-month period. Further studies assessing the utility of ethanol injection need to
be performed. PEI is operator-dependent, requires multiple sessions, and is less
effective in larger nodules. Also, there appears to be relatively little experience
with this therapy in the United States. Therefore, its use should be limited to
experienced clinicians and medical centers in selected circumstances.

TOXIC MULTINODULAR GOITER


Introduction
A toxic multinodular goiter is a thyroid gland that contains at least two autonomous
functioning thyroid nodules that secrete excessive amounts of thyroid hormone,
suppressing serum TSH and often causing typical symptoms and signs of hyper-
metabolism (311). These nodules may be more or less distinct on clinical examina-
tion and scan. Autonomous nodules require many years to develop and transition
through a phase when the TSH is normal and then subnormal with minimal clinical
evidence of hyperthyroidism (subclinical hyperthyroidism). Because of the time
required for this process to develop, most patients with toxic multinodular goiter
are over the age of 50. Many clinical aspects of patients with toxic multinodu-
lar goiters are similar to those found in those with solitary autonomous nodules
80 Burman and Cooper

A B

Figure 7 (A and B) An elderly patient with a toxic multinodular goiter. It is frequently


difficult to discern by clinical examination the extent of substernal extension of the thyroid
gland (330).

(Figs. 7 and 8). For example, exogenous iodine exposure can precipitate or aggra-
vate thyrotoxicosis (298). A recent analysis has suggested that the likelihood of
a toxic multinodular goiter harboring thyroid cancer was 9%, a value relatively
similar to the reported incidence of 10.6% in patients with a nontoxic multinodular
goiter (312) but lower (3.9%) than in Graves’ disease (6.5%) (313).

A B C

Figure 8 Thyroid radioisotope scans may be helpful in assessing certain patients with
hyperthyroidism. Panel A demonstrates symmetric isotope distribution (pertechnetate tech-
netium 99 m) typical of a patient with Graves’ disease. The right lobe appears larger than
the left because of rotation. Panel B shows an autonomous functioning thyroid nodule.
There is intense activity in the left-lobe nodule with absent activity in the right lobe because
of suppression of TSH by thyroid hormone secretion of the nodule. Panel C shows a toxic
multinodular goiter. Radioactive isotope activity is heterogenous, with areas of intense
activity interspersed with areas of reduced activity (330).
Hyperthyroidism 81

Pathogenesis
The pathogenesis of toxic multinodular goiter is not known, although it is believed
that individual thyroid follicles preferentially proliferate. Follicular size, colloid
content, and cellular characteristics vary widely in different parts of the multinodu-
lar goiter. Indeed, marked cellular variation is the hallmark of multinodular goiters.
It is believed that most of these nodules are clonal in origin. It is unknown how par-
ticular nodules grow and become autonomous. There are two theories concerning
the etiology of multinodular goiters. In the first, it is hypothesized that hyperplastic
nodules result from chronic activation by external factors such as TSH, iodide,
IGF-1, or thyroid growth–stimulating immunoglobulins (314,315). These factors
then cause the growth of polyclonal nodules. This theory does not have widespread
support, since trophic hormonal stimulation is not generally thought to result in
autonomous growth. The more widely held view is that several individual clones
develop, perhaps related to a genetic mutation, and that these cells gradually grow,
become autonomous, and finally result in excess hormonal secretion (316–319).
A rare entity, McCune-Albright syndrome, may be associated with the presence of
a toxic multinodular goiter due to a mutation in the gene encoding the Gs protein
alpha subunit which couples transmembrane-domain receptors to adenyl cyclase.
This results in constitutive activation of adenyl cyclases and overproduction of
cAMP. This syndrome is also associated with precocious puberty, fibrous dyspla-
sia of bone, abnormal gonadal function, and café-au-lait spots on the skin (320).
These patients may have a higher risk of thyroid cancer (320).

Diagnosis
Patients present with typical clinical and biochemical hyperthyroidism and small,
medium-sized, or large multinodular goiters. There is nothing particularly unusual
about their presentation, signs, or symptoms compared to patients with other
causes of hyperthyroidism, except the nodules and goiter may be sufficiently large
to cause local compressive symptoms. The radionuclide scan shows heterogenous
uptake with areas of hyper- and hypointensity (Fig. 8). There is a spectrum of
disease, ranging from minimal thyroid enlargement with small thyroid nodules
only detected on scan or sonogram to markedly enlarged thyroid glands and
large nodules. Color flow Doppler sonography may also be useful especially in
helping to differentiate nodular variants of Graves’ disease from toxic multinodular
goiters that are nonautoimmune mediated (321). Some multinodular goiters have
a significant substernal component.

Treatment
Radioiodine and surgery are the two major treatment modalities in patients with
toxic multinodular goiters, (311,322–324). Because multinodular goiters probably
are not responding to serum anti-TSH receptor immunoglobulins, it is not consid-
ered possible to induce a long-term remission with the chronic use of antithyroid
82 Burman and Cooper

agents. It may be appropriate to render a patient euthyroid with the use of antithy-
roid agents prior to radioiodine therapy, especially patients older than 60 years
or those with underlying heart disease. Selected patients can be maintained on
these agents for an indefinite period of time, for example, elderly patients with
accompanying serious medical disorders. However, this approach is applicable to
a minority of patients with toxic multinodular goiters.
Radioiodine therapy is most often used to restore euthyroidism or induce
hypothyroidism (322,323). 131 I therapy is generally thought to be less reliable
in controlling the hyperthyroidism, compared with Graves’ disease, and higher
doses are usually required (approximately 20–30 mCi 131 I). The explanation for
the difference in responses between Graves’ disease and toxic multinodular goiter
is unknown, but probably relates to the fact that there are different degrees of
autonomy and radiosensitivity among cells comprising the multinodular goiter.
The dose of 131 I varies, but it is reasonable to attempt to deliver approximately
160 to 200 ␮C 131 I/g thyroid tissue. In one recent study, following 131 I treatment
for a toxic multinodular goiter, 62% of patients were euthyroid, 19% hypothyroid
and 19% remained hyperthyroid. The mean size reduction overall was 32%
(325). It is often difficult to estimate the size of multinodular goiters, especially
because they may have a substernal component. A thyroid sonogram is helpful if
the gland is cervical in location, and CT scanning (without intravenous contrast)
is useful for estimating the size of a large multinodular goiter with substernal
extension.
It is difficult to predict the response to a given dose of 131 I; in one study,
the cure rate was relatively similar to that seen in Graves’ disease (325,326), but
others have not found similar degrees of success (327). Release of preformed and
stored thyroid hormones can occur after a dose of 131 I, so that these patients, who
are typically older, should be carefully monitored carefully. Not unexpectedly,
smaller thyroid glands seem to respond more consistently than larger glands,
although there is wide variation. Elderly patients or those with associated medical
disorders or heart disease should be treated with antithyroid agents prior to
administering 131 I, as this is thought but not proven to reduce the chance of
radioiodine-induced worsening of the hyperthyroidism. However, thionamide
exposure may cause a higher failure rate of radioiodine therapy (180), so higher
doses of radioiodine should be employed.
A near total thyroidectomy represents an alternative therapeutic option
(327,328). This procedure must be performed by an experienced thyroid sur-
geon, but even in this circumstance there is a risk of hypocalcemia and recurrent
laryngeal nerve paralysis as well as, in unusual circumstances, acute release of
stored thyroid hormone, with an exacerbation of hyperthyroidism. It is important
to establish a euthyroid state prior to surgery. The procedure of choice is a near
total thyroidectomy to remove as much thyroid tissue as the surgeon is comfortable
with, and to preserve the parathyroid glands and the recurrent laryngeal nerves.
Some clinicians prefer that patients be given thyroxine therapy postoperatively,
even if they have sufficient thyroid tissue remaining to prevent hypothyroidism.
Hyperthyroidism 83

This approach is controversial since, except for patients who have received exter-
nal radiation therapy in the past, thyroid hormone treatment to euthyroid patients
has not been proven to decrease the likelihood of recurrent nodule formation (329).
It does obviate the possibility of severe hypothyroidism developing, but periodic
monitoring is required whether or not thyroxine therapy is used.
Although therapy should be individualized and discussed with patients and
their families as appropriate, most patients with hyperthyroidism due to a multin-
odular goiter should be treated with 131 I therapy. Antithyroid agent therapy may
be useful for patients with smaller glands with less severe hyperthyroidism; long-
term antithyroid agent therapy should be reserved for selected patients. Surgery
addresses the problem expeditiously and can be used quite effectively; it should
be considered for patients with very large thyroid glands (above 150–200 g) since
the likelihood of such patients responding to 131 I therapy is lower. Compressive
symptoms such as hoarseness, superior vena caval syndrome, dysphagia, and/or
dyspnea are additional indications for surgery. However, in patients who are not
surgical candidates, radioiodine should be used.

REFERENCES
1. Burman KD, Baker JR Jr. Immune mechanisms in Graves’ disease. Endocr Rev
1985; 6:183–232.
2. Rapoport B, McLachlan SM. The thyrotropin receptor in Graves’ disease. Thyroid
2007; 17:911–922.
3. Burman K. Hyperthyroidism. In: Becker KA, ed. Principles and Practice of
Endocrinology and Metabolism, 2nd ed. Philadelphia, PA: J. B. Lippincott Co.,
1995:367–385.
4. Burman KD, Pandian R. Clinical utility of assays for TSH receptor antibodies. The
Endocrinologist 1998; 8:284.
5. Burch HB, Wartofsky L. Graves’ ophthalmopathy: Current concepts regarding patho-
genesis and management. Endocr Rev 1993; 14:747–793.
6. Bahn RS. Clinical review 157: Pathophysiology of Graves’ ophthalmopathy: The
cycle of disease. J Clin Endocrinol Metab 2003; 88:1939–1946.
7. Chen FQ, Okamura K, Sato K, et al. Reversible primary hypothyroidism with block-
ing or stimulating type TSH binding inhibitor immunoglobulin following recombi-
nant interferon-alpha therapy in patients with pre-existing thyroid disorders. Clin
Endocrinol (Oxf) 1996; 45:207–214.
8. Burman K, Becker. K, Cytryn A, Goodglick T. Thyroid Diseases. Philadelphia, PA:
Current Medicine, 1999.
9. Kasagi K, Hidaka A, Nakamura H, et al. Thyrotropin receptor antibodies in hypothy-
roid Graves’ disease. J Clin Endocrinol Metab 1993; 76:504–508.
10. Kraiem Z, Baron E, Kahana L, et al. Changes in stimulating and blocking TSH
receptor antibodies in a patient undergoing three cycles of transition from hypo to
hyper-thyroidism and back to hypothyroidism. Clin Endocrinol (Oxf) 1992; 36:211–
214.
11. Mengistu M, Lukes YG, Nagy EV, et al. TSH receptor gene expression in retroocular
fibroblasts. J Endocrinol Invest 1994; 17:437–441.
84 Burman and Cooper

12. Caturegli P, Kimura H, Rocchi R, et al. Autoimmune thyroid diseases. Curr Opin
Rheumatol 2007; 19:44–48.
13. Khoo TK, Bahn RS. Pathogenesis of Graves’ ophthalmopathy: The role of autoan-
tibodies. Thyroid 2007; 17:1013–1018.
14. Zakarija M, McKenzie JM. Pregnancy-associated changes in the thyroid-stimulating
antibody of Graves’ disease and the relationship to neonatal hyperthyroidism. J Clin
Endocrinol Metab 1983; 57:1036–1040.
15. Zakarija M, McKenzie JM, Munro D. Immunoglobulin G inhibitor of thyroid-
stimulating antibody is a cause of delay in the onset of neonatal Graves’ disease.
J Clin Invest 1983; 72:1352–1356.
16. Zakarija M, McKenzie JM, Hoffman WH. Prediction and therapy of intrauterine and
late-onset neonatal hyperthyroidism. J Clin Endocrinol Metab 1986; 62:368–371.
17. Polak M, Le Gac I, Vuillard E, et al. Fetal and neonatal thyroid function in relation
to maternal Graves’ disease. Best Pract Res Clin Endocrinol Metab 2004; 18:289–
302.
18. Burgi H, Kohler M, Morselli B. Thyrotoxicosis incidence in Switzerland and benefit
of improved iodine supply [letter]. Lancet 1998; 352:1034.
19. Mostbeck A, Galvan G, Bauer P, et al. The incidence of hyperthyroidism in Austria
from 1987 to 1995 before and after an increase in salt iodization in 1990. Eur J Nucl
Med 1998; 25:367–374.
20. Paunkovic N, Paunkovic J, Pavlovic O, et al. The significant increase in incidence
of Graves’ disease in eastern Serbia during the civil war in the former Yugoslavia
(1992 to 1995). Thyroid 1998; 8:37–41.
21. Jankovic SM, Radosavljevic VR, Marinkovic JM. Risk factors for Graves’ disease.
Eur J Epidemiol 1997; 13:15–18.
22. Bartley GB, Fatourechi V, Kadrmas EF, et al. The incidence of Graves’ ophthal-
mopathy in olmsted county, minnesota. Am J Ophthalmol 1995; 120:511–517.
23. Flynn RW, MacDonald TM, Morris AD, et al. The thyroid epidemiology, audit, and
research study: Thyroid dysfunction in the general population. J Clin Endocrinol
Metab 2004; 89:3879–3884.
24. Laurberg P, Bulow Pedersen I, Knudsen N, et al. Environmental iodine intake affects
the type of nonmalignant thyroid disease. Thyroid 2001; 11:457–469.
25. Delange F, de Benoist B, Alnwick D. Risks of iodine-induced hyperthyroidism after
correction of iodine deficiency by iodized salt. Thyroid 1999; 9:545–556.
26. Lind P, Langsteger W, Molnar M, et al. Epidemiology of thyroid diseases in iodine
sufficiency. Thyroid 1998; 8:1179–1183.
27. Sundick RS, Bagchi N, Brown TR. The role of iodine in thyroid autoimmunity:
From chickens to humans: A review. Autoimmunity 1992; 13:61–68.
28. Brown TR, Bagchi N. The role of iodine in the development of autoimmune thy-
roiditis. Int Rev Immunol 1992; 9:167–182.
29. Brown TR, Sundick RS, Dhar A, et al. Uptake and metabolism of iodine is crucial
for the development of thyroiditis in obese strain chickens. J Clin Invest 1991;
88:106–111.
30. Sundick RS, Herdegen DM, Brown TR, et al. The incorporation of dietary iodine into
thyroglobulin increases its immunogenicity. Endocrinology 1987; 120:2078–2084.
31. Sundick RS, Herdegen D, Brown TR, et al. Thyroidal iodine metabolism in obese-
strain chickens before immune- mediated damage. J Endocrinol 1991; 128:239–244.
Hyperthyroidism 85

32. Bagchi N, Brown TR, Urdanivia E, et al. Induction of autoimmune thyroiditis in


chickens by dietary iodine. Science 1985; 230:325–327.
33. Prummel MF, Wiersinga WM. Smoking and risk of Graves’ disease. JAMA 1993;
269:479–482.
34. Wong V, Fu AX, George J, et al. Thyrotoxicosis induced by alpha-interferon therapy
in chronic viral hepatitis. Clin Endocrinol (Oxf) 2002; 56:793–798.
35. Rapoport B, Chazenbalk GD, Jaume JC, et al. The thyrotropin (TSH) receptor:
Interaction with TSH and autoantibodies. Endocr Rev 1998; 19:673–716.
36. Weetman AP. Graves’ disease. N Engl J Med 2000; 343:1236–1248.
37. Siu CW, Yeung CY, Lau CP, et al. Incidence, clinical characteristics and outcome of
congestive heart failure as the initial presentation in patients with primary hyperthy-
roidism. Heart 2007; 93:483–487.
38. Siu CW, Zhang XH, Yung C, et al. Hemodynamic changes in hyperthyroidism-related
pulmonary hypertension: A prospective echocardiographic study. J Clin Endocrinol
Metab 2007; 92:1736–1742.
39. Burman KD, Monchik JM, Earll JM, et al. Ionized and total serum calcium and
parathyroid hormone in hyperthyroidism. Ann Intern Med 1976; 84:668–6671.
40. Uzzan B, Campos J, Cucherat M, et al. Effects on bone mass of long term treatment
with thyroid hormones: A meta-analysis. J Clin Endocrinol Metab 1996; 81:4278–
4289.
41. Faber J, Galloe AM. Changes in bone mass during prolonged subclinical hyper-
thyroidism due to L-thyroxine treatment: A meta-analysis. Eur J Endocrinol 1994;
130:350–356.
42. Hanna FW, Pettit RJ, Ammari F, et al. Effect of replacement doses of thyroxine on
bone mineral density. Clin Endocrinol (Oxf) 1998; 48:229–234.
43. Murphy E, Williams GR. The thyroid and the skeleton. Clin Endocrinol (Oxf) 2004;
61:285–928.
44. Wexler JA, Sharretts J. Thyroid and bone. Endocrinol Metab Clin North Am 2007;
36:673–705, vi.
45. Duncan WE, Wartofsky L, Burman KD. Influence of clinical characteristics and
parameters associated with thyroid hormone therapy on the bone mineral density of
women treated with thyroid hormone. Thyroid 1994; 4:183–190.
46. Affinito P, Sorrentino C, Farace MJ, et al. Effects of thyroxine therapy on bone
metabolism in postmenopausal women with hypothyroidism. Acta Obstet Gynecol
Scand 1996; 75:843–848.
47. Allain TJ, McGregor AM. Thyroid hormones and bone. J Endocrinol 1993; 139:9–
18.
48. Solomon BL, Wartofsky L, Burman KD. Prevalence of fractures in postmenopausal
women with thyroid disease. Thyroid 1993; 3:17–23.
49. Vestergaard P, Mosekilde L. Hyperthyroidism, bone mineral, and fracture risk–a
meta-analysis. Thyroid 2003; 13:585–593.
50. Papadopoulos KI, Diep T, Cleland B, et al. Thyrotoxic period paralysis: Report of
three cases and review of the literature. J Intern Med 1997; 241:521–524.
51. Manoukian MA, Foote JA, Crapo LM. Clinical and metabolic features of thyrotoxic
periodic paralysis in 24 episodes. Arch Intern Med 1999; 159:601–606.
52. Gonzalez-Trevino O, Rosas-Guzman J. Normokalemic thyrotoxic periodic paralysis:
A new therapeutic strategy. Thyroid 1999; 9:61–63.
86 Burman and Cooper

53. Kung AW, Lau KS, Cheung WM, et al. Thyrotoxic periodic paralysis and polymor-
phisms of sodium-potassium ATPase genes. Clin Endocrinol (Oxf) 2006; 64:158–
161.
54. Lin SH, Lin YF. Propranolol rapidly reverses paralysis, hypokalemia, and hypophos-
phatemia in thyrotoxic periodic paralysis. Am J Kidney Dis 2001; 37:620–623.
55. MacCrimmon DJ, Wallace JE, Goldberg WM, et al. Emotional disturbance and
cognitive deficits in hyperthyroidism. Psychosom Med 1979; 41:331–340.
56. Martin FI, Deam DR. Hyperthyroidism in elderly hospitalised patients. Clinical
features and treatment outcomes. Med J Aust 1996; 164:200–203.
57. Davis PJ, Rappeport JR, Lutz JH, et al. Three thyrotoxic criminals. Ann Intern Med
1971; 74:743–745.
58. Baba M, Terada A, Hishida R, et al. Persistent hemichorea associated with thyro-
toxicosis. Intern Med 1992; 31:1144–1146.
59. Krassas GE, Pontikides N, Kaltsas T, et al. Menstrual disturbances in thyrotoxicosis.
Clin Endocrinol (Oxf) 1994; 40:641–644.
60. Ridgway EC, Maloof F, Longcope C. Androgen and oestrogen dynamics in hyper-
thyroidism. J Endocrinol 1982; 95:105–115.
61. Carlson H. Current concepts-gynecomastia. N Engl J Med 1980; 305:795.
62. Carani C, Isidori AM, Granata A, et al. Multicenter study on the prevalence of sexual
symptoms in male hypo- and hyperthyroid patients. J Clin Endocrinol Metab 2005;
90:6472–6479.
63. Rotella CM, Zonefrati R, Toccafondi R, et al. Ability of monoclonal antibodies to the
thyrotropin receptor to increase collagen synthesis in human fibroblasts: An assay
which appears to measure exophthalmogenic immunoglobulins in Graves’ sera.
J Clin Endocrinol Metab 1986; 62:357–367.
64. Fatourechi V, Ahmed DD, Schwartz KM. Thyroid acropachy: Report of 40 patients
treated at a single institution in a 26-year period. J Clin Endocrinol Metab 2002;
87:5435–5441.
65. Fatourechi V. Pretibial myxedema: Pathophysiology and treatment options. Am J
Clin Dermatol 2005; 6:295–309.
66. Davis PJ, Davis FB. Hyperthyroidism in patients over the age of 60 years. Clinical
features in 85 patients. Medicine (Baltimore) 1974; 53:161–181.
67. Nordyke RA, Gilbert FI Jr, Harada AS. Graves’ disease. Influence of age on clinical
findings. Arch Intern Med 1988; 148:626–6231.
68. Trivalle C, Doucet J, Chassagne P, et al. Differences in the signs and symptoms
of hyperthyroidism in older and younger patients. J Am Geriatr Soc 1996; 44:50–
53.
69. Nuovo JA, Wartofsky L. Adverse Effects of Iodine. In: Becker KA, ed. Principles
and Practice of Endocrinology and Metabolism, 2nd ed. Philadelphia, PA: J. B.
Lippincott Co., 1995:332–338.
70. Endo T, Kogai T, Nakazato M, et al. Autoantibody against Na+/I- symporter in the
sera of patients with autoimmune thyroid disease. Biochem Biophys Res Commun
1996; 224:92–95.
71. Levy O, De la Vieja A, Carrasco N. The Na+/I- symporter (NIS): Recent advances.
J Bioenerg Biomembr 1998; 30:195–206.
72. Hollowell JG, Staehling NW, Hannon WH, et al. Iodine nutrition in the United
States. Trends and public health implications: Iodine excretion data from national
Hyperthyroidism 87

health and nutrition examination surveys I and III (1971–1974 and 1988–1994).
J Clin Endocrinol Metab 1998; 83:3401–3408.
73. Dallas JS, Cunningham SJ, Patibandla SA, et al. Thyrotropin (TSH) receptor anti-
bodies (TSHrAb) can inhibit TSH-mediated cyclic adenosine 3’,5’- monophosphate
production in thyroid cells by either blocking TSH binding or affecting a step sub-
sequent to TSH binding. Endocrinology 1996; 137:3329–3339.
74. Kohn LD, Suzuki K, Hoffman WH, et al. Characterization of monoclonal thyroid-
stimulating and thyrotropin binding-inhibiting autoantibodies from a Hashimoto’s
patient whose children had intrauterine and neonatal thyroid disease. J Clin
Endocrinol Metab 1997; 82:3998–4009.
75. Kung AW, Jones BM. A change from stimulatory to blocking antibody activ-
ity in Graves’ disease during pregnancy. J Clin Endocrinol Metab 1998; 83:514–
518.
76. Costagliola S, Morgenthaler NG, Hoermann R, et al. Second generation assay for
thyrotropin receptor antibodies has superior diagnostic sensitivity for Graves’ dis-
ease. J Clin Endocrinol Metab 1999; 84:90–907.
77. Grasso YZ, Kim MR, Faiman C, et al. Epitope heterogeneity of thyrotropin receptor-
blocking antibodies in Graves’ patients as detected with wild-type versus chimeric
thyrotropin receptors. Thyroid 1999; 9:531–537.
78. Di Cerbo A, Di Paola R, Menzaghi C, et al. Graves’ immunoglobulins activate
phospholipase A2 by recognizing specific epitopes on thyrotropin receptor. J Clin
Endocrinol Metab 1999; 84:3283–3292.
79. Reinhardt MJ, Moser E. An update on diagnostic methods in the investigation of
diseases of the thyroid. Eur J Nucl Med 1996; 23:587–594.
80. Cooper DS. Hyperthyroidism. Lancet 2003; 362:459–468.
81. Kamishlian A, Matthews N, Gupta A, et al. Different outcomes of neonatal thyroid
function after Graves’ disease in pregnancy: Patient reports and literature review.
J Pediatr Endocrinol Metab 2005; 18:1357–1363.
82. Luton D, Le Gac I, Vuillard E, et al. Management of Graves’ disease during preg-
nancy: The key role of fetal thyroid gland monitoring. J Clin Endocrinol Metab
2005; 90:6093–6098.
83. Weetman AP. New aspects of thyroid immunity. Horm Res 1997; 48:51–54.
84. Sakata S, Nakamura S, Miura K. Autoantibodies against thyroid hormones for
iodothyronine:Implications in diagnosis, thyroid function, and pathogenesis. Ann
Int Med 1985; 103:579–589.
85. Despres N, Grant AM. Antibody interference in thyroid assays: A potential for
clinical misinformation. Clin Chem 1998; 44:440–454.
86. McIver B, Gorman CA. Euthyroid sick syndrome: An overview. Thyroid 1997;
7:125–132.
87. Vexiau P, Perez-Castiglioni P, Socie G, et al. The ‘euthyroid sick syndrome’: Inci-
dence, risk factors and prognostic value soon after allogeneic bone marrow trans-
plantation. Br J Haematol 1993; 85:778–782.
88. Wartofsky L, Burman KD. Alterations in thyroid function in patients with systemic
illness: The euthyroid sick syndrome. Endocr Rev 1982; 3:164–217.
89. Cooper DS. Antithyroid drugs. N Engl J Med 2005; 352:905–917.
90. Solomon B, Glinoer D, Lagasse R, et al. Current trends in the management of Graves’
disease. J Clin Endocrinol Metab 1990; 70:1518–1524.
88 Burman and Cooper

91. Wartofsky L, et al. Differences and similarities in the diagnosis and treatment of
Graves’ disease in Europe, Japan, and the United States. Thyroid 1991; 1:129–135.
92. Okamura K, Ikenoue H, Shiroozu A, et al. Reevaluation of the effects of methylmer-
captoimidazole and propylthiouracil in patients with Graves’ hyperthyroidism. J Clin
Endocrinol Metab 1987; 65:719–723.
93. Nicholas W, Fischer R, Stevenson R, et al. Single daily dose of methimazole com-
pared to every 8 hours of propylthiouracil in the treatment of hyperthyroidism. South
Med J 1995; 88:973–976.
94. Nakamura H, Noh JY, Itoh K, et al. Comparison of methimazole and propylthiouracil
in patients with hyperthyroidism caused by Graves’ disease. J Clin Endocrinol Metab
2007; 92:2157–2162.
95. Hirota Y, Tamai H, Hayashi Y, et al. Thyroid function and histology in forty-five
patients with hyperthyroid Graves’ disease in clinical remission more than ten years
after thionamide drug treatment. J Clin Endocrinol Metab 1986; 62:165–169.
96. Torring O, Tallstedt L, Wallin G, et al. Graves’ hyperthyroidism: Treatment with
antithyroid drugs, surgery, or radioiodine–a prospective, randomized study. Thyroid
Study Group. J Clin Endocrinol Metab 1996; 81:2986–2993.
97. Young ET, Steel NR, Taylor JJ, et al. Prediction of remission after antithyroid drug
treatment in Graves’ disease. Q J Med 1988; 250:175–189.
98. Takamatsu JS, Sugawara M, Kuma K, et al. Ratio of serum triiodothyronine to
thyroxine and the prognosis of triiodothyronine-predominant Graves’ disease. Ann
Intern Med 1984; 100:372–375.
99. Kawai K, Tamai H, Matsubayashi S, et al. A study of untreated Graves’ patients with
undetectable TSH binding inhibitor immunoglobulins and the effect of antithyroid
drugs. Clin Endocrinol 1995; 43:551–556.
100. Amino N, Tanizawa O, Mori H, et al. Aggravation of thyrotoxicosis in early preg-
nancy and after delivery in Graves’ disease. J Clin Endocrinol Metab 1982; 55:108–
112.
101. Messina M, Milani P, Gentile L. Initial treatment of thyrotoxic Graves’ disease with
methimazole: A randomized trial comparing different dosages. J Endocrinol Invest
1987; 10:291–295.
102. Reinwein D, Benker G, Lazarus JH, et al. A prospective randomized trial of antithy-
roid drug dose in Graves’ disease therapy. J Clin Endocrinol Metab 1993; 76:1516–
1521.
103. Page SR, Sheard CE, Herbert M, et al. A comparison of 20 or 40 mg per day of
carbimazole in the initial treatment of hyperthyroidism. Clin Endocrinol (Oxf) 1996;
45:511–516.
104. Grebe SKG, Feek CM, Ford HC, et al. A randomized trial of short-term treatment
of Graves’ disease with high-dose carbimazole plus thyroxine versus low-dose car-
bimazole. Clin Endocrinol (Oxf) 1998; 48:585–592.
105. Jorde R, Ytre-Arne K, Stormer J, et al. Short-term treatment of Graves’ disease with
methimazole in high versus low doses. J Intern Med 1995; 238:161–165.
106. Chen JJ, Ladenson PW. Discordant hypothyroxinemia and hypertriiodothyroninemia
in treated patients with hyperthyroid Graves’ disease. J Clin Endocrinol Metab 1986;
63:102–106.
107. Hashizume K, Ichikawa I, Sakurai A, et al. Administration of thyroxine in treated
Graves’ Disease. N Engl J Med 1991; 324:947–953.
Hyperthyroidism 89

108. McIver B, Rae P, Beckett G, et al. Lack of effect of thyroxine in patients with
Graves’ hyperthyroidism who are treated with an antithyroid drug. N Engl J Med
1996; 334:220–224.
109. Rittmaster RS, Zwicker H, Abbott EC, et al. Effect of methimazole with or with-
out exogenous l-thyroxine on serum concentrations of thyrotropin (TSH) receptor
antibodies in patients with Graves’ disease. J Clin Endocrinol Metab 1996; 81:3283–
3288.
110. Pfeilschifter J, Zeigler R. Suppression of serum thyrotropin with thyroxine in patients
with Graves’ disease: Effects on recurrence hyperthyroidism and thyroid volume.
Eur J Endocrinol 1997; 136:81–86.
111. Lucas A, Salinas I, Rius F, et al. Medical therapy of Graves’ disease: Does thyroxine
prevent recurrence of hyperthyroidism? J Clin Endocrinol Metab 1997; 82:2410–
2413.
112. Pujol P, Osman A, Grabar S, et al. TSH suppression combined with carbimazole
for Graves’ disease: Effect on remission and relapse rates. Clin Endocrinol 1988;
48:635–640.
113. Tamai H, Nakagawa T, Fukino O, et al. Thionamide therapy in Graves’ disease:
Relation of relapse rate to duration of therapy. Ann Intern Med 1980; 92:488–
490.
114. Lippe BM, Landaw EM, Kaplan SA. Hyperthyroidism in children treated with
long term medical therapy: Twenty-five percent remission every two years. J Clin
Endocrinol Metab 1987; 64:1241–1245.
115. Maugendre D, Gatel A, Campion L, et al. Antithyroid drugs and Graves’ disease–
prospective randomized assessment of long-term treatment. Clin Endocrinol (Oxf)
1999; 50:127–132.
116. Garcia-Mayor RVG, Paramo C, Luna-Cano R, et al. Antithyroid drug and Graves’
hyperthyroidism. Significance of treatment duration and TRAb determination on
lasting remission. J Endocrinol Invest 1992; 15:815–820.
117. Teng CS, Yeung RTT. Changes in thyroid-stimulating antibody activity in Graves’
disease treated with antithyroid drug and its relationship to relapse: A prospective
study. J Clin Endocrinol Metab 1980; 50:144–147.
118. Cho BY, Shong MH, Yi KH, et al. Evaluation of serum basal thyrotrophin levels and
thyrotrophin receptor antibody activities as prognostic markers for discontinuation
of antithyroid drug treatment in patients with Graves’ disease. Clin Endocrinol (Oxf)
1992; 36:585–590.
119. Vitti P, Valente WA, Ambesi-Impiombato FS, et al. Graves’ IgG stimulation of
continuously cultured rat thyroid cells: A sensitive and potentially useful clinical
assay. J Endocrinol Invest 1982; 5:179–182.
120. Nagai K, Tamai H, Mukuta T, et al. A follow-up study of 85 patients with Graves’ dis-
ease in remission who developed undetectable serum thyroid-stimulating hormone
concentrations using sensitive TSH assays. Horm Res 1991; 35:185–189.
121. Sugrue D, McEvoy M, Feely J, et al. Hyperthyroidism in the land of Graves:
Results of treatment by surgery, radio-iodine and carbimazole in 837 cases. Q J
Med 1980;49:51–61.
122. Hedley AJ, Young RE, Jones SJ, et al. Antithyroid drugs in the treatment of hyper-
thyroidism of Graves’ disease: Long-term follow-up of 434 patients. Clin Endocrinol
1989; 31:209–218.
90 Burman and Cooper

123. Slingerland DW, Burrows BA. Long-term antithyroid treatment in hyperthyroidism.


JAMA 1979; 242:2408–2410.
124. Cooper D. The side-effects of antithyroid drugs. The Endocrinologist 1999; In Press.
125. Werner MC, Romaldini JH, Bromberg N, et al. Adverse effects related to thionamide
drugs and their dose regimen. Am J Med Sci 1989; 296:216–219.
126. Hallman BL, Hurst JW. Loss of taste as toxic effect of methimazole (tapazole)
therapy. JAMA 1953; 152:322.
127. Suzuki S, Ichihawa K, Nagai M, et al. Elevation of serum creatine kinase during
treatment with antithyroid drugs in patients with hyperthyroidism due to Graves’
disease. Arch Int Med 1997; 157:693–696.
128. Tajiri J, Noguchi S, Murakami N. Usefulness of granulocyte count measurement four
hours after injection of granulocyte colony-stimulating factor for detecting recovery
from antithyroid drug-induced granulocytopenia. Thyroid 1997; 4:575–578.
129. Salama A, Mueller-Eckhardt C. Immune-mediated blood cell dyscrasias related to
drugs. Sem Hematol 1992; 29:54–63.
130. Tamai H, Sudo T, Kimura A, et al. Association between the DRB1∗ 08032 histocom-
patibility antigen and methimazole-induced agranulocytosis in Japanese patients
with Graves’ disease. Ann Intern Med 1996; 124:490–494.
131. Pearce SH. Spontaneous reporting of adverse reactions to carbimazole and propy-
lthiouracil in the UK. Clin Endocrinol (Oxf) 2004; 61:589–594.
132. Sheng WH, Hung CC, Chen YC, et al. Antithyroid-drug-induced agranulocytosis
complicated by life-threatening infections. QJM 1999; 92:455–461.
133. Julia A, Olona M, Bueno J, et al. Drug-induced agranulocytosis: Prognostic factors
in a series of 168 episodes. Br J Haematol 1991; 79:366–371.
134. Tajiri J, Noguchi S, Okamura S, et al. Granulocyte colony-stimulating factor treat-
ment of antithyroid drug-induced granulocytopenia. Arch Int Med 1993; 153:509–
514.
135. Fukata S, Kuma K, Sugawara M. Granulocyte colony-stimulating factor (G-CSF)
does not improve recovery from antithyroid drug-induced agranulocytosis: A
prospective study. Thyroid 1999; 9:29–31.
136. Tajiri J, Noguchi S. Antithyroid drug-induced agranulocytosis: How has granulocyte
colony-stimulating factor changed therapy? Thyroid 2005; 15:292–297.
137. Editorial. Drug-induced agranulocytosis. Drug Ther Bull 1997; 35:49.
138. Shabtai R, Shapiro MS, Orenstein D, et al. The antithyroid arthritis syndrome
reviewed. Arth Rheum 1984; 27:227–229.
139. Gunton JE, Steil J, Caterson RJ, et al. Anti-thyroid drugs and antineutrophil cyto-
plasmic antibody positive vasculitis. A case report and review of the literature. J Clin
Endocrinol Metab 1999; 84:13–16.
140. Sato H, Hattori M, Fujieda M, et al. High prevalence of antineutrophil cytoplasmic
antibody positivity in childhood onset Graves’ disease treated with propylthiouracil.
J Clin Endocrinol Metab 2000; 85:4270–4273.
141. Harper L, Chin L, Daykin J, et al. Propylthiouracil and carbimazole associated-
antineutrophil cytoplasmic antibodies (ANCA) in patients with Graves’ disease.
Clin Endocrinol (Oxf) 2004; 60:671–675.
142. Guma M, Salinas I, Reverter JL, et al. Frequency of antineutrophil cytoplasmic
antibody in Graves’ disease patients treated with methimazole. J Clin Endocrinol
Metab 2003; 88:2141–2146.
Hyperthyroidism 91

143. Hanson J. Propylthiouracil and hepatitis. Arch Int Med 1984; 144:944–996.
144. Williams KV, Nayak S, Becker D, et al. Fifty years of experience with
propylthiouracil-associated hepatotoxicity: What have we learned? J Clin Endocrinol
Metab 1997; 82:1727–1733.
145. Kirsh BM, Lam N, Layden TJ, et al. Diagnosis and management of fulminant hepatic
failure. Comp Ther 1995; 21:166–171.
146. Weiss M, Hassin D, Bank H. Propylthiouracil-induced hepatic damage. Arch Intern
Med 1980; 140:1184–1185.
147. Parker W. Propylthiouracil-induced hepatotoxicity. Clin Pharm 1982; 1:471–474.
148. Waseem M, Seshadri KG, Kabadi UM. Successful outcome with methimazoleand
lithium combination therapy for propylthiouracil-induced hepatotoxicity. Endo Prac
1998; 4:197–200.
149. Liaw YF, Huang MJ, Fan KD, et al. Hepatic injury during propylthiouracil therapy
in patients with hyperthyroidism. Ann Inter Med 1993; 118:424–428.
150. Huang MJ, Li KL, Wei JS, et al. Sequential liver and bone biochemical changes in
hyperthyroidism: Prospective controlled follow-up study. Am J Gastroenterol 1994;
89:1071–1076.
151. Woeber KA. Methimazole-induced hepatotoxicity. Endocr Pract 2002; 8:222–224.
152. Arab DM, Malatjalian DA, Rittmaster RS. Severe cholestatic jaundice in uncom-
plicated hyperthyroidism treated with methimazole. J Clin Endocrinol Metab 1995;
80:1083–1085.
153. Geffner DL, Hershman JM. Beta-adrenergic blockade for the treatment of hyperthy-
roidism. Am J Med 1992; 93:61–68.
154. Cooper DS, Daniels GH, Ladenson PW, et al. Hyperthyroxinemia in patients treated
with high-dose propranolol. Am J Med 1982; 73:867–871.
155. Georges LP, Santangelo RP, Mackin JF, et al. Metabolic effects of propranolol in
thyrotoxicosis. I. Nitrogen, calcium and hydroxyproline. Metabolism 1975; 24:11–
21.
156. Valcavi R, Menozzi C, Roti E, et al. Sinus node function in hypethyroid patients.
J Clin Endocrinol Metab 1992; 75:239–242.
157. Grossman W, Robin NI, Johnson LW, et al. Effects of beta blcokade on the peripheral
manifestations of thyrotoxicosis. Ann Intern Med 1971; 74:875–879.
158. Saunders J, Hall SEH, Crowther A, et al. The effect of propanolol on thyroid hor-
mones and oxygen consumption in thyrotoxicosis. Clin Endocrinol 1978; 1978:67–
72.
159. O’Malley BP, Abbott RJ, Barnett DB, et al. Propranolol versus carbimazole as the
sole treatment for thyrotoxicosis. A consideration of circulating thyroid hormone
levels and tissue thyroid function. Clin Endocrinol 1982; 16:545–552.
160. Feely JS, Stevenson IH, Crooks J. Increased clearance of propranolol in thyrotoxi-
cosis. Ann Intern Med 1981; 94:472–474.
161. Klein I, Ojamaa K. Thyrotoxicosis and the heart. Endocrinol Clin North Am 1998;
27:51–62.
162. Roti E, Montermini M, Roti S, et al. The effect of diltiazem, a calcium channel-
blocking drug, on cardiac rate and rhythm in hyperthyroid patients. Arch Intern Med
1988; 148:1919–1921.
163. Wolff J. Effects of lithium on thyroid gland function. Neurosci Res Program Bull
1976; 14:178–180.
92 Burman and Cooper

164. Boehm TM, Burman KD, Barnes S, et al. Lithium and iodine combination therapy
for thyrotoxicosis. Acta Endocrinol (Copenh) 1980; 94:174–183.
165. Lawrence JE, Lamm SH, Braverman LE. The use of perchlorate for the prevention
of thyrotoxicosis in patients given iodine rich contrast agents. J Endocrinol Invest
1999; 22:405–407.
166. Spitzweg C, Joba W, Morris JC, et al. Regulation of sodium iodide symporter gene
expression in FRTL-5 rat thyroid cells. Thyroid 1999; 9:821–830.
167. Wolff J. Perchlorate and the thyroid gland. Pharmacol Rev 1998; 50:89–105.
168. Bartalena L, Brogioni S, Grasso L, et al. Treatment of amiodarone-induced thyrotox-
icosis, a difficult challenge: Results of a prospective study. J Clin Endocrinol Metab
1996; 81:2930–2933.
169. Graham GD, Burman KD. Radioiodine treatment of Graves’ disease. An assessment
of its potential risks. Ann Intern Med 1986; 105:900–905.
170. Levy EG. Treatment of Graves’ disease: The American way. Baillieres Clin
Endocrinol Metab 1997; 11:585–595.
171. Hennessey JV. Diagnosis and management of thyrotoxicosis. Am Fam Physician
1996; 54:1315–1324.
172. Peters H, Fischer C, Bogner U, et al. Radioiodine therapy of Graves’ hyperthy-
roidism: Standard vs. calculated 131iodine activity. Results from a prospective,
randomized, multicentre study. Eur J Clin Invest 1995; 25:186–193.
173. Leslie WD, Ward L, Salamon EA, et al. A randomized comparison of radioio-
dine doses in Graves’ hyperthyroidism. J Clin Endocrinol Metab 2003; 88:978–
983.
174. Uy HL, Reasner CA, Samuels MH. Pattern of recovery of the hypothalamic-pituitary-
thyroid axis following radioactive iodine therapy in patients with Graves’ disease.
Am J Med 1995; 99:173–179.
175. Tamagna E, Levine GA, Hershman JM. Thyroid hormone concentrations after
radioiodine therapy for hyperthyroidism. J Nucl Med 1979; 20:387–391.
176. Stensvold AD, Jorde R, Sundsfjord J. Late and transient increases in free T4
after radioiodine treatment for Graves’ disease. J Endocrinol Invest 1997; 20:580–
584.
177. Chiovato L, Fiore E, Vitti P, et al. Outcome of thyroid function in Graves’ patients
treated with radioiodine: Role of thyroid-stimulating and thyrotropin-blocking anti-
bodies and of radioiodine-induced thyroid damage. J Clin Endocrinol Metab 1998;
83:40–46.
178. Andrade VA, Gross JL, Maia AL. Effect of methimazole pretreatment on serum
thyroid hormone levels after radioactive treatment in Graves’ hyperthyroidism.
J Clin Endocrinol Metab 1999; 84:4012–4016.
179. Bonnema SJ, Bennedbaek FN, Veje A, et al. Continuous methimazole therapy and
its effect on the cure rate of hyperthyroidism using radioactive iodine: An evaluation
by a randomized trial. J Clin Endocrinol Metab 2006; 91:2946–2951.
180. Walter MA, Briel M, Christ-Crain M, et al. Effects of antithyroid drugs on radioiodine
treatment: Systematic review and meta-analysis of randomised controlled trials. BMJ
2007; 334:514.
181. Velkeniers B, Vanhaelst L, Cytryn R, et al. Treatment of hyperthyroidism with
radioiodine: Adjuctive therapy with antithyroid drugs reconsidered. Lancet 1988;
1:1127–1129.
Hyperthyroidism 93

182. Marcocci C, Gianchecchi D, Masini I, et al. A reappraisal of the role of methimazole


and other factors on the efficacy and outcome of radioiodine therapy of Graves’
hyperthyroidism. J Endocrinol Invest 1990; 13:513.
183. Tuttle RM, Patience T, Budd S. Treatment with propylthiouracil before radioactive
iodine therapy is associated with a higher treatment failure rate than radioiodine
therapy alone in Graves’ disease. Thyroid 1995; 5:243–247.
184. Hancock LD, Tuttle RM, LeMar H, et al. The effect of propylthiouracil on subsequent
radioactive iodine therapy in Graves’ disease. Clin Endocrinol (Oxf) 1997; 47:425–
430.
185. Bonnema SJ, Bennedbaek FN, Veje A, et al. Propylthiouracil before 131I therapy of
hyperthyroid diseases: Effect on cure rate evaluated by a randomized clinical trial.
J Clin Endocrinol Metab 2004; 89:4439–4444.
186. Andrade VA, Gross JL, Maia AL. The effect of methimazole pretreatment on the
efficacy of radioactive iodine therapy in Graves’ hyperthyroidism: One-year follow-
up of a prospective, randomized study. J Clin Endocrinol Metab 2001; 86:3488–3493.
187. Braga M, Walpert N, Burch HB, et al. The effect of methimazole on cure rates after
radioiodine treatment for Graves’ hyperthyroidism: A randomized clinical trial.
Thyroid 2002; 12:135–139.
188. Bogazzi F, Bartalena L, Brogioni S, et al. Comparison of radioiodine with radioiodine
plus lithium in the treatment of Graves’ hyperthyroidism. J Clin Endocrinol Metab
1999; 84:499–503.
189. Tallstedt L, Lundell G, Torring O, et al. Occurrence of ophthalmopathy after treat-
ment for Graves’ hyperthyroidism. The Thyroid Study Group. N Engl J Med 1992;
326:1733–1738.
190. Bartalena L, Marcocci C, Bogazzi F, et al. Relation between therapy or hyperthy-
roidism and the course of Graves’ ophthalmopathy. N Engl J Med 1998; 338:73–78.
191. Perros P, Kendall-Taylor P, Neoh C, et al. A prospective study of the effects of
radioiodine therapy for hyperthyroidism in patients with minimally active graves’
ophthalmopathy. J Clin Endocrinol Metab 2005; 90:5321–5323.
192. Bartalena L, Marcocci C, Tanda ML, et al. Cigarette smoking and treatment outcomes
in Graves ophthalmopathy. Ann Intern Med 1998; 129:632–635.
193. Tallstedt L, Lundell G, Blomgren H, et al. Does early administration of thyroxine
reduce the development of Graves’ ophthalmopathy after radioiodine treatment? Eur
J Endocrinol 1994; 130:494–497.
194. Ron E, Doody MM, Becker DV, et al. Cancer mortality following treatment for adult
hyperthyroidism. Cooperative Thyrotoxicosis Therapy Follow-up Study Group.
JAMA 1998; 280:347–355.
195. Franklyn JA, Maisonneuve P, Sheppard M, et al. Cancer incidence and mortality
after radioiodine treatment for hyperthyroidism: A population-based cohort study.
Lancet 1999; 353:2111–2115.
196. Flynn RW, Macdonald TM, Jung RT, et al. Mortality and vascular outcomes in
patients treated for thyroid dysfunction. J Clin Endocrinol Metab 2006; 91:2159–
2164.
197. Chou FF, Wang PW, Huang SC. Results of subtotal thyroidectomy for Graves’
disease. Thyroid 1999; 9:253–257.
198. Duh QY. Thyroidectomy for the treatment of Graves’ disease [comment]. Thyroid
1999; 9:259–261.
94 Burman and Cooper

199. Leech NJ, Dayan CM. Controversies in the management of Graves’ disease. Clin
Endocrinol (Oxf) 1998; 49:273–280.
200. Linos DA, Karakitsos D, Papademetriou J. Should the primary treatment of hyper-
thyroidism be surgical? Eur J Surg 1997; 163:651–657.
201. Witte J, Goretzki PE, Roher HD. Surgery for Graves disease in childhood and
adolescence. Exp Clin Endocrinol Diabetes 1997; 105:58–60.
202. Werga-Kjellman P, Zedenius J, Tallstedt L, et al. Surgical treatment of hyperthy-
roidism: A ten-year experience. Thyroid 2001; 11:187–192.
203. Lennquist S, Jortso E, Anderberg BO, et al. Beta blockers compared with antithyroid
drugs as preoperative treatment in hyperthyroidism: Drug tolerance, complications,
and postoperative thyroid function. Surgery 1985; 98:1141–1147.
204. Feely J, Crooks J, Forrest A, et al. Propranolol in the surgical treatment of hyperthy-
roidism, including severely thyrotoxic patients. Br J Surg 1981; 68:865–869.
205. Baeza A, Aguayo M, Barria M, et al. Rapid preoperative preparation in hyperthy-
roidism. Clin Endocrinol 1991; 35:439–442.
206. Erbil Y, Ozluk Y, Giris M, et al. Effect of lugol solution on thyroid gland blood
flow and microvessel density in the patients with Graves’ disease. J Clin Endocrinol
Metab 2007; 92:2182–2189.
207. Burch HB, Wartofsky L. Life-threatening thyrotoxicosis. Thyroid storm. Endocrinol
Metab Clin North Am 1993; 22:263–277.
208. Hermann M, Roka R, Richter B, et al. Early relapse after operation for Graves’
disease: Postoperative hormone kinetics and outcome after subtotal, near-total, and
total thyroidectomy. Surgery 1998; 124:894–900.
209. Levetan C, Wartofsky L. A clinical guide to the management of Graves’ disease with
radioactive iodine. Endocr Pract 1995; 1:205.
210. Patel NN, Abraham P, Buscombe J, et al. The cost effectiveness of treatment modal-
ities for thyrotoxicosis in a U. K. center. Thyroid 2006; 16:593–598.
211. Dietlein M, Moka D, Dederichs B, et al. Cost-effectiveness analysis: Radioiodine
or antithyroid medication in primary treatment of immune hyperthyroidism. Nuk-
learmedizin 1999; 38:7–14.
212. Eng PH, Cardona GR, Fang SL, et al. Escape from the acute Wolff-Chaikoff effect is
associated with a decrease in thyroid sodium/iodide symporter messenger ribonucleic
acid and protein. Endocrinology 1999; 140:3404–3410.
213. Werner SC. Classification of the eye changes of Graves’ disease. Am J Ophthalmol
1969; 68:646–648.
214. Major BJ, Busuttil BE, Frauman AG. Graves’ ophthalmopathy: Pathogenesis and
clinical implications. Aust N Z J Med 1998; 28:39–45.
215. Gerding MN, Terwee CB, Dekker FW, et al. Quality of life in patients with Graves’
ophthalmopathy is markedly decreased: Measurement by the medical outcomes
study instrument. Thyroid 1997; 7:885–889.
216. Bahn RS. Understanding the immunology of Graves’ ophthalmopathy. Is it an
autoimmune disease? Endocrinol Metab Clin North Am 2000; 29:287–296, vi.
217. Stan MN, Garrity JA, Bradley EA, et al. Randomized, double-blind, placebo-
controlled trial of long-acting release octreotide for treatment of Graves’ ophthal-
mopathy. J Clin Endocrinol Metab 2006; 91:4817–4824.
218. Bahn RS. TSH receptor expression in orbital tissue and its role in the pathogenesis
of Graves’ ophthalmopathy. J Endocrinol Invest 2004; 27:216–220.
Hyperthyroidism 95

219. Spitzweg C, Joba W, Hunt N, et al. Analysis of human thyrotropin receptor gene
expression and immunoreactivity in human orbital tissue. Eur J Endocrinol 1997;
136:599–607.
220. Perros P, Crombie AL, Kendall-Taylor P. Natural history of thyroid associated oph-
thalmopathy. Clin Endocrinol (Oxf) 1995; 42:45–50.
221. Tallstedt L, Lundell G, Taube A. Graves’ ophthalmopathy and tobacco smoking.
Acta Endocrinol (Copenh) 1993; 129:147–150.
222. Hegedius L, Brix TH, Vestergaard P. Relationship between cigarette smoking and
Graves’ ophthalmopathy. J Endocrinol Invest 2004; 27:265–271.
223. Wemeau JL, Caron P, Beckers A, et al. Octreotide (long-acting release formulation)
treatment in patients with graves’ orbitopathy: Clinical results of a four-month,
randomized, placebo-controlled, double-blind study. J Clin Endocrinol Metab 2005;
90:841–848.
224. Mourits MP, van Kempen-Harteveld ML, Garcia MB, et al. Radiotherapy for Graves’
orbitopathy: Randomised placebo-controlled study. Lancet 2000; 355:1505–1509.
225. Gorman CA. Radiotherapy for Graves’ ophthalmopathy: Results at one year. Thyroid
2002; 12:251–255.
226. Jernfors M, Valimaki MJ, Setala K, et al. Efficacy and safety of orbital decompres-
sion in treatment of thyroid-associated ophthalmopathy: Long-term follow-up of 78
patients. Clin Endocrinol (Oxf) 2007; 67:101–107.
227. Marcocci C, Pinchera A, Marino M. A treatment strategy for Graves’ orbitopathy.
Nat Clin Pract Endocrinol Metab 2007; 3:430–436.
228. Wiersinga WM. Management of Graves’ ophthalmopathy. Nat Clin Pract Endocrinol
Metab 2007; 3:396–404.
229. Meyer DR. Timing of surgery for Graves’ orbitopathy. Ophthalmology 2007;
114:393; [author reply1–2].
230. Rapoport B, Alsabeh R, Aftergood D, et al. Elephantiasic pretibial myxedema:
Insight into and a hypothesis regarding the pathogenesis of the extrathyroidal mani-
festations of Graves’ disease. Thyroid 2000; 10:685–692.
231. Fatourechi V, Pajouhi M, Fransway AF. Dermopathy of Graves disease (pretibial
myxedema). Review of 150 cases. Medicine (Baltimore) 1994; 73:1–7.
232. Kuyvenhoven JP, Van Der Pijl JW, Goslings BM, et al. Graves’ dermopathy: Does
octreotide scintigraphy predict the response to octreotide treatment? Thyroid 1996;
6:385–389.
233. Engin B, Gumusel M, Ozdemir M, et al. Successful combined pentoxifylline and
intralesional triamcinolone acetonide treatment of severe pretibial myxedema. Der-
matol Online J 2007; 13:16.
234. Pineda AM, Tianco EA, Tan JB, et al. Oral pentoxifylline and topical clobetasol
propionate ointment in the treatment of pretibial myxoedema, with concomitant
improvement of Graves’ ophthalmopathy. J Eur Acad Dermatol Venereol 2007;
21:1441–1443.
235. Gurlek A, Gedik O. Effect of endogenous subclinical hyperthyroidism on bone
metabolism and bone mineral density in premenopausal women. Thyroid 1999;
9:539–543.
236. Koutras DA. Subclinical hyperthyroidism. Thyroid 1999; 9:311–315.
237. Cooper DS. Subclinical thyroid disease: A clinician’s perspective [editorial]. Ann
Intern Med 1998; 129:135–138.
96 Burman and Cooper

238. Saadi H. Detecting and managing subclinical hyperthyroidism. Cleve Clin J Med
1998; 65:65–66.
239. Utiger RD. Subclinical hyperthyroidism–just a low serum thyrotropin concentration,
or something more? [editorial]. N Engl J Med 1994; 331:1302–1303.
240. Abe E, Sun L, Mechanick J, et al. Bone loss in thyroid disease: Role of low TSH
and high thyroid hormone. Ann N Y Acad Sci 2007; 1116:383–391.
241. Rosario PW. The natural history of subclinical hyperthyroidism in patients below
the age of 65 years. Clin Endocrinol (Oxf) 2008; 68:491–492.
242. Duggal J, Singh S, Barsano CP, et al. Cardiovascular risk with subclinical hyper-
thyroidism and hypothyroidism: Pathophysiology and management. J Cardiometab
Syndr 2007; 2:198–206.
243. Franklyn JA. Subclinical thyroid disorders–consequences and implications for treat-
ment. Ann Endocrinol (Paris) 2007; 68:229–230.
244. Iervasi G, Molinaro S, Landi P, et al. Association between increased mortality and
mild thyroid dysfunction in cardiac patients. Arch Intern Med 2007; 167:1526–
1532.
245. Singh S, Duggal J, Molnar J, et al. Impact of subclinical thyroid disorders on coronary
heart disease, cardiovascular and all-cause mortality: A meta-analysis. Int J Cardiol
2008; 125:41–48.
246. Belaya ZE, Melnichenko GA, Rozhinskaya LY, et al. Subclinical hyperthyroidism
of variable etiology and its influence on bone in postmenopausal women. Hormones
(Athens) 2007; 6:62–70.
247. Morris MS. The association between serum thyroid-stimulating hormone in its ref-
erence range and bone status in postmenopausal American women. Bone 2007;
40:1128–1134.
248. Cooper DS. Approach to the patient with subclinical hyperthyroidism. J Clin
Endocrinol Metab 2007; 92:3–9.
249. Pohlenz J, Pfarr N, Kruger S, et al. Subclinical hyperthyroidism due to a thyrotropin
receptor (TSHR) gene mutation (S505R). Acta Paediatr 2006; 95:1685–1687.
250. Walsh JP, Bremner AP, Bulsara MK, et al. Subclinical thyroid dysfunction and blood
pressure: A community-based study. Clin Endocrinol (Oxf) 2006; 65:486–491.
251. Volzke H, Alte D, Dorr M, et al. The association between subclinical hyperthyroidism
and blood pressure in a population-based study. J Hypertens 2006; 24:1947–1953.
252. Lee WY, Oh KW, Rhee EJ, et al. Relationship between subclinical thyroid dys-
function and femoral neck bone mineral density in women. Arch Med Res 2006;
37:511–516.
253. Casey BM. Subclinical hypothyroidism and pregnancy. Obstet Gynecol Surv 2006;
61:415–420.
254. Reverter JL, Holgado S, Alonso N, et al. Lack of deleterious effect on bone min-
eral density of long-term thyroxine suppressive therapy for differentiated thyroid
carcinoma. Endocr Relat Cancer 2005; 12:973–981.
255. Dorr M, Volzke H. Cardiovascular morbidity and mortality in thyroid dysfunction.
Minerva Endocrinol 2005; 30:199–216.
256. Walsh JP, Bremner AP, Bulsara MK, et al. Subclinical thyroid dysfunction as a risk
factor for cardiovascular disease. Arch Intern Med 2005; 165:2467–2472.
257. Gharib H, Tuttle RM, Baskin HJ, et al. Subclinical thyroid dysfunction: A joint state-
ment on management from the American Association of Clinical Endocrinologists,
Hyperthyroidism 97

the American Thyroid Association, and the Endocrine Society. J Clin Endocrinol
Metab 2005; 90:581–585; discussion 6–7.
258. Surks MI, Ortiz E, Daniels GH, et al. Subclinical thyroid disease: Scientific review
and guidelines for diagnosis and management. JAMA 2004; 291:228–238.
259. Sawin CT, Geller A, Wolf PA, et al. Low serum thyrotropin concentrations as a
risk factor for atrial fibrillation in older persons. N Engl J Med 1994; 331:1249–
1252.
260. Mudde AH, Houben AJ, Nieuwenhuijzen Kruseman AC. Bone metabolism dur-
ing anti-thyroid drug treatment of endogenous subclinical hyperthyroidism. Clin
Endocrinol (Oxf) 1994; 41:421–424.
261. Bauer D, Nevitt M, Ettinger B, et al. Low thyrotropin levels are not assoictaed with
bone loss in older women: A prospective study. J Clin Endocrinol Metab 1997;
82:2931–2936.
262. Figge J, Leinung M, Goodman AD, et al. The clinical evaluation of patients with
subclinical hyperthyroidism and free triiodothyronine (free T3) toxicosis. Am J Med
1994; 96:229–234.
263. Kahaly GJ, Nieswandt J, Mohr-Kahaly S. Cardiac risks of hyperthyroidism in the
elderly. Thyroid 1998; 8:1165–1169.
264. Cappola AR, Fried LP, Arnold AM, et al. Thyroid status, cardiovascular risk, and
mortality in older adults. JAMA 2006; 295:1033–1041.
265. Shrier DK, Burman KD. Subclinical hyperthyroidism: Controversies in management.
Am Fam Physician 2002; 65:431–438.
266. Faber J, Jensen I, Petersen L, et al. Normalization of serum thyrotrophin by means
of radioidoine treatment in subclinical hyperthyroidism: Effect on bone loss in
postmenopausal women. Clin Endocrinol 1998; 48:285–290.
267. Buscemi S, Verga S, Cottone S, et al. Favorable clinical heart and bone effects of
anti-thyroid drug therapy in endogenous subclinical hyperthyroidism. J Endocrinol
Invest 2007; 30:230–235.
268. Nayak B, Burman K. Thyrotoxicosis and thyroid storm. Endocrinol Metab Clin
North Am 2006; 35:663–686.
269. Brooks MH, Waldstein SS, Bronsky D, et al. Serum triiodothyronine concentration
in thyroid storm. J Clin Endocrinol Metab 1975; 40:339–341.
270. Sherman SI, Simonson L, Ladenson PW. Clinical and socioeconomic predispositions
to complicated thyrotoxicosis: A predictable and preventable syndrome? Am J Med
1996; 101:192–198.
271. Hermann M, Richter B, Roka R, et al. Thyroid surgery in untreated severe hyper-
thyroidism: Perioperative kinetics of free thyroid hormones in the glandular venous
effluent and peripheral blood. Surgery 1994; 115:240–245.
272. Hodak SP, Huang C, Clarke D, et al. Intravenous methimazole in the treatment of
refractory hyperthyroidism. Thyroid 2006; 16:691–695.
273. Walter RM Jr, Bartle WR. Rectal administration of propylthiouracil in the treatment
of Graves’ disease. Am J Med 1990; 88:69–70.
274. Jongjaroenprasert W, Akarawut W, Chantasart D, et al. Rectal administration of
propylthiouracil in hyperthyroid patients: Comparison of suspension enema and
suppository form. Thyroid 2002; 12:627–631.
275. Nabil N, Miner DJ, Amatruda JM. Methimazole: An alternative route of adminis-
tration. J Clin Endocrinol Metab 1982; 54:180–181.
98 Burman and Cooper

276. Roti E, Robuschi G, Gardini E, et al. Comparison of methimazole, methimazole


and sodium ipodate, and methimazole and saturated solution of potassium iodide in
the early treatment of hyperthyroid Graves’ disease. Clin Endocrinol 1988; 28:305–
314.
277. Solomon BL, Wartofsky L, Burman KD. Adjunctive cholestyramine therapy for
thyrotoxicosis. Clin Endocrinol (Oxf) 1993; 38:39–43.
278. Burch HB, Shakir F, Fitzsimmons TR, et al. Diagnosis and management of the
autonomously functioning thyroid nodule: The Walter Reed Army Medical Center
experience, 1975–1996. Thyroid 1998; 8:871–880.
279. Ferrari C, Reschini E, Paracchi A. Treatment of the autonomous thyroid nodule: A
review. Eur J Endocrinol 1996; 135:383–390.
280. Als C, Rosler H, Kinser JA. Radioiodine therapy of the autonomous thyroid nodule
in patients with or without visible extranodular activity letter; comment. J Nucl Med
1996; 37:401–403.
281. David E, Rosen IB, Bain J, et al. Management of the hot thyroid nodule. Am J Surg
1995; 170:481–483.
282. Burman KD. Solitary autonomous thyroid nodules. Postgrad Med 1974; 56:70–74.
283. Burman KD, Earll JM, Johnson MC, et al. Clinical observations on the solitary
autonomous thyroid nodule. Arch Intern Med 1974; 134:915–919.
284. Siddiqui AR, Karanauskas S. Hurthle cell carcinoma in an autonomous thyroid
nodule in an adolescent. Pediatr Radiol 1995; 25:568–569.
285. Sandrock D, Olbricht T, Emrich D, et al. Long-term follow-up in patients with
autonomous thyroid adenoma. Acta Endocrinol (Copenh) 1993; 128:51–55.
286. Caplan RH, Strutt PJ, Kisken WA, et al. Fine needle aspiration biopsy of thyroid
nodules. Wis Med J 1991; 90:285–288.
287. Smith M, McHenry C, Jarosz H, et al. Carcinoma of the thyroid in patients with
autonomous nodules. Am Surg 1988; 54:448–449.
288. Wool MS. Thyroid nodules. The place of fine-needle aspiration biopsy in manage-
ment. Postgrad Med 1993; 94:111–112, 115–122.
289. Zelmanovitz F, Gross JL. Cytopathological findings from fine-needle aspiration
biopsy are accurate predictors of thyroid pathology in patients with functioning
thyroid nodules. J Endocrinol Invest 1998; 21:98–101.
290. Burch HB, Burman KD, Reed HL, et al. Fine needle aspiration of thyroid nodules.
Determinants of insufficiency rate and malignancy yield at thyroidectomy. Acta
Cytol 1996; 40:1176–1183.
291. Paschke R, Ludgate M. The thyrotropin receptor in thyroid diseases. N Engl J Med
1997; 337:1675–1681.
292. Fuhrer D, Holzapfel HP, Wonerow P, et al. Somatic mutations in the thyrotropin
receptor gene and not in the Gs alpha protein gene in 31 toxic thyroid nodules.
J Clin Endocrinol Metab 1997; 82:3885–3891.
293. Derwahl M, Manole D, Sobke A, et al. Pathogenesis of toxic thyroid adenomas and
nodules: Relevance of activating mutations in the TSH-receptor and Gs-alpha gene,
the possible role of iodine deficiency and secondary and TSH-independent molecular
mechanisms. Exp Clin Endocrinol Diabetes 1998; 106:S6–S9.
294. Krohn K, Fuhrer D, Holzapfel HP, et al. Clonal origin of toxic thyroid nodules with
constitutively activating thyrotropin receptor mutations. J Clin Endocrinol Metab
1998; 83:130–134.
Hyperthyroidism 99

295. Paschke R. Constitutively activating TSH receptor mutations as the cause of toxic
thyroid adenoma, multinodular toxic goiter and autosomal dominant non autoim-
mune hyperthyroidism. Exp Clin Endocrinol Diabetes 1996; 104:129–132.
296. Blum M, Shenkman L, Hollander CS. The autonomous nodule of the thyroid: Cor-
relation of patient age, nodule size and functional status. Am J Med Sci 1975;
269:43–50.
297. Silverstein GE, Burke G, Cogan R. The natural history of the autonomous hyper-
functioning thyroid nodule. Ann Intern Med 1967; 67:539–548.
298. Woeber KA. Iodine and thyroid disease. Med Clin North Am 1991; 75:169–178.
299. Brian SR, Cheng DW, Goldberg PA. Unusual case of amiodarone-induced thyro-
toxicosis: “illicit” use of a technetium scan to diagnose a transiently toxic thyroid
nodule. Endocr Pract 2007; 13:413–416.
300. Thrall JH, Burman KD, Wartofsky L, et al. Discordant imaging of a thyroid nodule
with 131I and 99mTc: Concordance of 131I and fluorescent scans. Radiology 1978;
128:705–706.
301. Burman KD, Adler RA, Wartofsky L. Hemiagenesis of the thyroid gland. Am J Med
1975; 58:143–146.
302. Ross D, Ridgway E, Daniels G. Successful treatment of solitary toxic thyroid nodules
with relatively low-dose iodine-131 with low prevalence of hypothyroidism. Ann Int
Med 1984; 101:488–490.
303. Goldstein R, Hart M. Follow-up of solitary autonomous thyroid nodules treated with
131I. New Engl J Med 1983; 309:1473–1476.
304. Wiener JD, de Vries AA. On the natural history of Plummer’s disease. Clin Nucl
Med 1979; 4:181–190.
305. O’Brien T, Gharib H, Suman VJ, et al. Treatment of toxic solitary thyroid nodules:
Surgery versus radioactive iodine. Surgery 1992; 112:1166–1170.
306. Thomas CG Jr, Croom RD. Current management of the patient with autonomously
functioning nodular goiter. Surg Clin North Am 1987; 67:315–328.
307. Hedman I, Jansson S, Lindberg S. Need for thyroxine in patients lobectomised
for benign thyroid disease as assessed by follow-up on average fifteen years after
surgery. Acta Chir Scand 1986; 152:481–486.
308. Lippi F, Ferrari C, Manetti L, et al. Treatment of solitary autonomous thyroid nod-
ules by percutaneous ethanol injection: Results of an Italian multicenter study. The
Multicenter Study Group. J Clin Endocrinol Metab 1996; 81:3261–3264.
309. Monzani F, Caraccio N, Goletti O, et al. Five-year follow-up of percutaneous ethanol
injection for the treatment of hyperfunctioning thyroid nodules: A study of 117
patients. Clin Endocrinol (Oxf) 1997; 46:9–15.
310. Mauz PS, Maassen MM, Braun B, et al. How safe is percutaneous ethanol injection
for treatment of thyroid nodule? Report of a case of severe toxic necrosis of the
larynx and adjacent skin. Acta Otolaryngol 2004; 124:1226–1230.
311. Siegel RD, Lee SL. Toxic nodular goiter. Toxic adenoma and toxic multinodular
goiter. Endocrinol Metab Clin North Am 1998; 27:151–168.
312. Cerci C, Cerci SS, Eroglu E, et al. Thyroid cancer in toxic and non-toxic multinodular
goiter. J Postgrad Med 2007; 53:157–160.
313. Cappelli C, Braga M, De Martino E, et al. Outcome of patients surgically treated for
various forms of hyperthyroidism with differentiated thyroid cancer: Experience at
an endocrine center in Italy. Surg Today 2006; 36:125–130.
100 Burman and Cooper

314. Joba W, Spitzweg C, Schriever K, et al. Analysis of human sodium/iodide symporter,


thyroid transcription factor-1, and paired-box-protein-8 gene expression in benign
thyroid diseases. Thyroid 1999; 9:455–466.
315. Kimura ET, Kopp P, Zbaeren J, et al. Expression of transforming growth factor beta1,
beta2, and beta3 in multinodular goiters and differentiated thyroid carcinomas: A
comparative study. Thyroid 1999; 9:119–125.
316. Tonacchera M, Vitti P, Agretti P, et al. Activating thyrotropin receptor mutations
in histologically heterogeneous hyperfunctioning nodules of multinodular goiter.
Thyroid 1998; 8:559–564.
317. Tonacchera M, Chiovato L, Pinchera A, et al. Hyperfunctioning thyroid nodules
in toxic multinodular goiter share activating thyrotropin receptor mutations with
solitary toxic adenoma. J Clin Endocrinol Metab 1998; 83:492–498.
318. Gabriel EM, Bergert ER, Grant CS, et al. Germline polymorphism of codon 727 of
human thyroid-stimulating hormone receptor is associated with toxic multinodular
goiter. J Clin Endocrinol Metab 1999; 84:3328–3335.
319. Holzapfel HP, Wonerow P, von Petrykowski W, et al. Sporadic congenital hyperthy-
roidism due to a spontaneous germline mutation in the thyrotropin receptor gene.
J Clin Endocrinol Metab 1997; 82:3879–3884.
320. Chanson P, Salenave S, Orcel P. McCune-Albright syndrome in adulthood. Pediatr
Endocrinol Rev 2007; 4(suppl 4):453–462.
321. Boi F, Loy M, Piga M, et al. The usefulness of conventional and echo colour
Doppler sonography in the differential diagnosis of toxic multinodular goitres. Eur
J Endocrinol 2000; 143:339–346.
322. Huysmans DA, Buijs WC, van de Ven MT, et al. Dosimetry and risk estimates of
radioiodine therapy for large, multinodular goiters. J Nucl Med 1996; 37:2072–2079.
323. Hurley DL, Gharib H. Evaluation and management of multinodular goiter. Otolaryn-
gol Clin North Am 1996; 29:527–540.
324. Pradeep PV, Agarwal A, Baxi M, et al. Safety and efficacy of surgical management
of hyperthyroidism: 15-year experience from a tertiary care center in a developing
country. World J Surg 2007; 31:306–312; discussion 13.
325. Tarantini B, Ciuoli C, Di Cairano G, et al. Effectiveness of radioiodine (131-I)
as definitive therapy in patients with autoimmune and non-autoimmune hyperthy-
roidism. J Endocrinol Invest 2006; 29:594–598.
326. Franklyn JA, Daykin J, Holder R, et al. Radioiodine therapy compared in patients
with toxic nodular or Graves’ hyperthyroidism. QJM 1995; 88:175–180.
327. Erickson D, Gharib H, Li H, et al. Treatment of patients with toxic multinodular
goiter. Thyroid 1998; 8:277–282.
328. Jensen MD, Gharib H, Naessens JM, et al. Treatment of toxic multinodular goiter
(Plummer’s disease): Surgery or radioiodine? World J Surg 1986; 10:673–680.
329. Fogelfeld L, Wiviott MB, Shore-Freedman E, et al. Recurrence of thyroid nodules
after surgical removal in patients irradiated in childhood for benign conditions [see
comments]. N Engl J Med 1989; 320:835–840.
330. Morris JC, ed. Hyperthyroidism from Toxic nodules and other Causes. Philadelphia,
PA: Current Medicine, 1999.
3
Thyroiditis and Other More
Unusual Forms of
Hyperthyroidism

Shon E. Meek and Robert C. Smallridge


College of Medicine, Mayo Clinic, Jacksonville,
Florida, U.S.A.

THYROIDITIS AND HYPERTHYROIDISM


Subacute Thyroiditis
Introduction
Subacute thyroiditis is a painful, inflammatory thyroid condition associated with
thyrotoxicosis. In the past, it has also been called granulomatous thyroiditis, giant-
cell thyroiditis, noninfectious thyroiditis, acute nonsuppurative thyroiditis, and de
Quervain’s thyroiditis.
Epidemiology
Subacute thyroiditis is not nearly as common as Graves’ disease, but it is more
common than silent thyroiditis if postpartum thyroiditis is excluded. It has been
reported to occur at the rate of one in five to eight cases of Graves’ disease (1). In
Olmsted County, Minnesota, subacute thyroiditis occurred at a rate of 4.9 cases
per 100,000/yr (2). Subacute thyroiditis has been reported in North America,
Europe, Scandinavia, and Japan, but it is not often reported in the tropical and
subtropical areas of the world. In Hawaii, subacute thyroiditis is seen at the
same rate among Caucasians and Japanese indicating a similar prevalence among
these two races living in the same environment (3). It is not known whether

101
102 Meek and Smallridge

the lack of occurrence in the tropical and subtropical areas is due to a lower
actual frequency or ascertainment bias. However, despite the possible geographic
variation, subacute thyroiditis is recognized more frequently during the summer
months (4,5). Subacute thyroiditis has been reported in all age groups. It is most
common in the third to sixth decades of life, and it is rare in children. Female
patients outnumber male patients in a ratio of up to 6:1 (6–8).

Pathophysiology
The cause of subacute thyroiditis is not known; however, it tends to occur following
an upper respiratory tract infection. Mumps, measles, influenza, common colds,
adenovirus, Epstein-Barr virus, coxsackievirus, and cat-scratch disease have all
been associated with subacute thyroiditis (9–15). Subacute thyroiditis is associated
with HLA-B35, approximately 72% of the time (14,16). Subacute thyroiditis has
been reported in twins and family members (17,18). Infiltrative diseases, such
as amyloid, have also been reported to cause a subacute thyroiditis–like picture
(19), and anaplastic thyroid cancer can rarely begin with painful, rapid thyroid
enlargement known as “malignant pseudothyroiditis” (20,21). Thyroiditis induced
by amiodarone can also occasionally present with a similar clinical picture.
The thyroid gland in subacute thyroiditis is enlarged and firm; it may adhere
to adjacent tissues. Thyroid tissue obtained by fine-needle aspiration (FNA) biopsy
shows an inflammatory infiltrate of neutrophils, lymphocytes, histiocytes, and
multinucleated giant cells (22–24).

Diagnosis
Patients with subacute thyroiditis usually present with an acute onset of malaise,
feverishness, and pain in the region of the thyroid gland. The pain may radiate
from the thyroid to the jaw and to the ears, or down to the anterior chest wall.
Coughing, swallowing, turning the head, or wearing tight clothing around the
neck can aggravate the pain. Approximately one-third to one-half of patients
may present with unilateral thyroid pain. Approximately one-third of patients
can have migratory pain throughout the thyroid, so-called “creeping thyroiditis”.
Approximately one-third of cases may present with diffuse pain in the thyroid
(1,15,25). Some biopsy-proven cases of subacute thyroiditis have been reported
to be painless (23,26,27).
Many patients may have systemic symptoms of malaise, myalgia, fever, and
anorexia. As noted above, approximately 50% of patients may have a history of
an antecedent upper respiratory infection. Symptoms of thyrotoxicosis are also
present in 50% to 60% of patients, and these may include heat intolerance, pal-
pitations, tremor, and nervousness. Cases of subacute thyroiditis causing thyroid
storm have been reported (28).
Physical examination shows an uncomfortable patient with a tender,
enlarged, and firm thyroid gland. The process is often asymmetric, and lym-
phadenopathy is usually not present. Symptoms of thyrotoxicosis may last 4 to
Thyroiditis and Other Forms of Hyperthyroidism 103

Figure 1 The graph correlates the time from the beginning of pain to the disappearance of
pain and palpable abnormalities in 70 patients observed through the course of the disease.
In all instances, pain and tenderness ceased first. The mean duration of pain was 65 days
versus 84 days for palpable abnormalities. Source: From Ref. 3.

10 weeks, but the inflammation with a painful, tender thyroid often lasts for eight
weeks, and on rare occasions, up to one year. Pain and tenderness resolve first,
followed by resolution of the palpable thyroid abnormalities, as shown in Figure 1
(3). If the patient is not seen until late in the course of the disease and after pain
resolves, the discovery of a thyroid nodule as the residual of lobar enlargement
may lead to unnecessary surgery, unless an FNA is performed.
Laboratory evaluation shows increased serum levels of thyroxine (T4) and
triiodothyronine (T3) due to follicular disruption, with release of stored thyroid
hormones and thyroglobulin into the systemic circulation. The T4/T3 ratio is
typically higher than in Graves’ disease, reflecting glandular hormone stores.
The serum TSH level is suppressed (29). The white blood cell count is usu-
ally normal, but it may be moderately increased. The erythrocyte sedimentation
rate is virtually always increased, often to as high or higher than 100 mm/hr
(15,25). Thyroid autoantibodies are usually absent or present in low titer, and if
present, they are usually transient. The 24-hour radioactive iodine uptake is very
low.
As the course of subacute thyroiditis progresses, the serum concentration
of thyroid hormones returns to normal. In more severe cases, transient hypothy-
roidism develops (30). Thyroid function usually returns to normal, but perma-
nent hypothyroidism may occur in 15% of patients with extended follow-up.
Patients treated with corticosteroid therapy may develop hypothyroidism more
commonly than those not treated with corticosteroids (2). Recurrent bouts of
subacute thyroiditis may occur in 4%of patients, 6 to 21 years after the initial
episode (2). Figure 2 shows the typical phases of thyroid function during subacute
thyroiditis.
104 Meek and Smallridge

Figure 2 Natural history of subacute thyroiditis.

Treatment
Nonsteroidal anti-inflammatory drugs (NSAIDs) or salicylates (2 g/day) are used
initially to treat subacute thyroiditis (31,32). However, corticosteroids are used for
more severe cases or in patients not responding to NSAIDs, and result in rapid
clinical improvement (33). Corticosteroids produce partial or near complete relief
of pain and neck tenderness within 24 to 48 hours. If symptoms do not respond
promptly, an alternate diagnosis, such as acute infectious thyroiditis, should be
considered. Typically, prednisone in an initial dose of 40 mg/day is used for about
a week, followed by a tapering dose of 10 mg/wk and withdrawal by four weeks.
As the drug is tapered, exacerbation of pain may occur in approximately 20% of
patients (8,32). If this occurs, the dose can be increased and treatment continued for
another month. In extremely rare cases, neck pain and malaise may be prolonged. In
these cases, thyroidectomy may be needed (34). Beta-adrenergic antagonist drugs
may be helpful in controlling symptoms of thyrotoxicosis. However, they are not
usually needed because corticosteroids or NSAIDs usually alleviate thyrotoxicosis
as well as the thyroid pain.

Silent Thyroiditis
Introduction
Silent or “painless” thyroiditis is a painless inflammation of the thyroid that pro-
duces a transient hyperthyroid state (35). The terms silent thyroiditis and painless
thyroiditis are used most commonly to describe this condition. However, silent
thyroiditis has also been called transient painless thyroiditis, painless thyroiditis
with transient hyperthyroidism, painless subacute thyroiditis, atypical thyroidi-
tis, occult subacute thyroiditis, lymphocytic thyroiditis, spontaneously resolving
lymphocytic thyroiditis, and transient thyrotoxicosis with lymphocytic thyroidi-
tis. Silent thyroiditis often occurs in the postpartum period, and is then called
postpartum thyroiditis.
Thyroiditis and Other Forms of Hyperthyroidism 105

Epidemiology
The incidence of painless thyroiditis was reported with increasing frequency in
the late 1970s and early 1980 s in the Great Lakes region of the United States
and Canada. Silent thyroiditis has also been reported in South America, India,
and Japan. However, it has been reported to be less frequent on the east and west
coasts of the United States and in Europe and Argentina (36). Patients are usually
between 30 and 60 years of age, but silent thyroiditis can occur in all age groups.
There is a female-to-male predominance of approximately 1.5 to 1. There is an
11% chance that patients may have recurrent episodes of silent thyroiditis (37).

Pathophysiology
In most cases, silent thyroiditis is an autoimmune disease and likely a variant
of Hashimoto’s thyroiditis. Histologically, silent thyroiditis is characterized by a
lymphocytic infiltration of the thyroid, and it is sometimes associated with lym-
phoid follicles (38,39). It is associated with other autoimmune diseases, such
as autoimmune adrenal insufficiency, lupus erythematosus, idiopathic thrombo-
cytopenic purpura, and rheumatoid arthritis (40–43). Silent thyroiditis has been
associated with HLA DR3, which suggests a genetic component to the disease
(44). Thyroid autoantibodies are present in the serum in up to 50% of patients,
which suggests an autoimmune process (35). No association with a viral infection
has been found. However, the lack of antibodies in some patients and lack of
clear female predominance suggests that silent thyroiditis maybe a heterogeneous
disorder.

Diagnosis
Patients with silent thyroiditis present with symptoms and signs of thyrotoxicosis.
The most common symptoms include palpitations, weight loss, nervousness, heat
intolerance, and fatigue. The thyrotoxic phase may last from 1 to 12 months,
but it usually lasts about 3 months. Approximately one-half of patients have a
goiter, in which the thyroid is 1.5 to 3 times the normal size, diffusely enlarged,
symmetric, firm, and nontender (35). The course of the disease typically follows
three different phases. The first phase is characterized by hyperthyroidism, and
many, but not all patients will go on to develop hypothyroidism as a second stage
of silent thyroiditis. Most patients then become euthyroid in the third stage, but
permanent hypothyroidism may develop months to years later.
During the first phase of silent thyroiditis, the serum T4 and T3 levels
are increased and serum TSH is decreased. The T4/T3 ratio is higher in silent
thyroiditis than in Graves’ disease, reflecting glandular hormonal stores. The
radioactive iodine uptake is very low. Thyroglobulin levels are increased, which
may be useful in distinguishing silent thyroiditis from factitious thyrotoxicosis.
Serum thyroglobulin concentrations may remain slightly increased even one to two
years after recovery of normal thyroid function (45). Thyroid autoantibody levels
are increased approximately 30% to 50% of the time. However, approximately
106 Meek and Smallridge

50% of the positive antibody titers become negative within six months after thyroid
recovery (35,37). The white cell count is usually normal. The sedimentation rate
is normal in ⬎50% of cases, with only mild elevation in the remaining cases (46).
FNA of the thyroid shows lymphocytic infiltration, but aspiration is rarely needed
to make the diagnosis. Biopsies show that silent thyroiditis lacks some of the
features of chronic lymphocytic thyroiditis, such as no Hürthle cells or germinal
centers (38).
Patients with silent thyroiditis have a low radioactive iodine uptake, which
distinguishes it from states of high radioactive iodine uptake such as Graves’
disease or toxic nodular goiter. Silent thyroiditis, with its low radioactive iodine
uptake, must be distinguished from iodine-induced thyrotoxicosis, excess thyroid
hormone ingestion, and amiodarone-induced thyrotoxicosis (AIT). In addition,
struma ovarii can cause a low radioactive iodine uptake over the thyroid, but in
these cases, uptake over the ovarian tumor will also be increased.
Thyroid hormone levels decrease during the hypothyroid phase and then
return to normal during the recovery phase. TSH levels often rise transiently in
the recovery phase. The radioactive iodine uptake may also rise transiently above
the normal range during the recovery phase of silent thyroiditis.
Treatment
As silent thyroiditis usually presents with mild to moderate symptoms of hyper-
thyroidism, treatment to relieve symptoms may not be necessary. For patients
who are more than mildly symptomatic, beta-adrenergic blocking agents can be
administered. Antithyroid drugs are not useful because destroying thyroid cells
releases thyroid hormones, causing the thyrotoxic phase of silent thyroiditis. If
severe thyrotoxicosis is present, corticosteroids can be administered to decrease the
inflammatory process (46). Patients have rarely been treated with thyroidectomy
when they have had frequent debilitating episodes of silent thyroiditis (39,46).
Once normal thyroid function returns with a normal radioiodine uptake, patients
with recurrent episodes of silent thyroiditis may consider radioactive iodine abla-
tion of the thyroid. The hypothyroid phase of silent thyroiditis usually does not
need to be treated since it is usually quite mild, and most patients fully recover
normal thyroid function, at least initially. However, if the hypothyroid stage is
severe or prolonged, thyroxine can be administered for several months. Almost all
patients recover normal thyroid function after an episode of silent thyroiditis, but
since approximately 50% of patients with silent thyroiditis will ultimately develop
hypothyroidism, thyroid function should be monitored yearly (46).

Postpartum Thyroiditis
Introduction
Postpartum thyroiditis is a syndrome of thyroid dysfunction that occurs within
the first year following parturition. It is usually characterized by transient painless
Thyroiditis and Other Forms of Hyperthyroidism 107

thyrotoxicosis with a low radioactive iodine uptake, often followed by a hypothy-


roid phase that is then followed by thyroid recovery. However, many postpartum
thyroiditis patients ultimately develop permanent hypothyroidism within a few
years (47).
Epidemiology
Postpartum thyroiditis has been reported in North America, South America,
Europe, and Asia. An average prevalence figure of about 5% to 9% of postpar-
tum women has been generally accepted (48–55). The lower frequency of 1.1%
in Asia may be related to variations in regional, dietary iodine intake or genetic
differences in susceptibility (56). Approximately 10% of women in the general
population have positive antibodies and approximately one-half of these patients
develop postpartum thyroiditis. An increased incidence of postpartum thyroiditis
(10–25%) is found among patients with type 1 diabetes mellitus, reflecting the
underlying autoimmune diathesis (57–59). Postpartum thyroid dysfunction has
also been reported after a miscarriage, although, this case only had postmiscar-
riage hypothyroidism (60).
Pathophysiology
Women who are prone to developing postpartum thyroiditis most likely have pre-
existing, asymptomatic autoimmune thyroiditis. During pregnancy, the maternal
immune system is partially suppressed, with a subsequent rise in thyroid autoan-
tibodies after delivery. Studies have shown that higher thyroid antibody levels
are associated with a higher risk of thyroid dysfunction and clinical symptoms
(61–64). Postpartum thyroiditis has also been related to HLA type. HLA-DR3,
-DR4, and -DR5 are increased in patients with postpartum thyroiditis (65–68).
Biopsy specimens of thyroid tissue during postpartum thyroiditis have shown a
lymphocytic infiltration (52). Smoking was associated with postpartum thyroiditis
in two studies, (48,69) but it was not associated with smoking in three other studies
(53,70,71).
Several studies have shown postpartum thyroiditis to be associated with
the presence of goiter during pregnancy (50,61,72). One study using ultrasound
showed a significant increase in thyroid volume between 8- and 20-weeks’ ges-
tation in women who went on to develop postpartum thyroiditis (73). However,
a prospective study using ultrasound found that thyroid size, before, during, or
after pregnancy, was not a useful indicator for the development of postpartum
thyroiditis (70). Therefore, even though postpartum thyroiditis may be associated
with a goiter, the presence of a goiter is not a predictive indicator for postpartum
thyroiditis.
Diagnosis
Patients with postpartum thyroiditis may present with fatigue, palpitations, heat
intolerance, nervousness, emotional liability, and other hyperthyroid symptoms.
108 Meek and Smallridge

Hypothyroid Hyperthyroid-
40% Graves’
11%

Hyperthyroid-Thyroiditis
24%
Hyper/Hypothyroid
25%

Figure 3 Frequency of hyperthyroidism, hypothyroidism, or both in postpartum thyroid


dysfunction.

Many patients will have some enlargement of the thyroid. Postpartum thyroiditis is
almost universally painless, although one case of painful disease has been reported
(74). In postpartum thyroiditis, there is an absence of exophthalmos and, almost
always, an increase in antithyroid antibody titers. Patients may present at a time
when thyroid hormones levels are high, normal, or low. Since the hyperthyroid
phase is a destructive type of thyroiditis, there is a low 24-hour radioactive iodine
uptake.
Figure 3 shows the frequency of hyperthyroidism, hypothyroidism, or both
in postpartum thyroid dysfunction. The classical triphasic pattern of postpartum
thyroiditis is mild hyperthyroidism followed by transient hypothyroidism with
subsequent thyroid recovery. This pattern of postpartum thyroiditis occurs in 25%
of patients. The thyrotoxic phase usually presents one to six months postpartum.
Frequently, a period of hypothyroidism develops over the next three to four months,
followed by a return to normal thyroid function. Some patients with postpartum
thyroiditis only develop transient hyperthyroidism without subsequent hypothy-
roidism. This can either take the form of thyroiditis-induced hyperthyroidism in
approximately 24% of patients or hyperthyroidism caused by Graves’ disease in
approximately 11% of patients. Some patients (40%) with postpartum thyroiditis
present only with hypothyroidism that is followed by recovery of thyroid function
(75). Figure 4 shows the possible stages of thyroid function in the natural history
of postpartum thyroiditis.
All patients with postpartum thyroiditis should be monitored for the future
development of thyroid failure. Approximately 20% to 64% of patients with tran-
sient thyroid disease postpartum become hypothyroid with long-term follow-up
(62,65,66,76,77). Factors associated with the development of permanent hypothy-
roidism include higher titer of thyroid autoantibodies, greater severity of the
hypothyroid phase of postpartum thyroiditis, and a previous history of sponta-
neous abortion (74,76). Microsomal and thyroid peroxidase antibodies have been
reported to have a sensitivity range of 0.45 to 0.89 and a specificity range of 0.9
to 0.98 (75).
Thyroiditis and Other Forms of Hyperthyroidism 109

Figure 4 Natural history of postpartum thyroiditis.

Treatment
Treatment of the thyrotoxic phase of postpartum thyroiditis is often not needed,
since the symptoms are usually mild. Beta-adrenergic blocking drugs can be used
in symptomatic patients, but should be used with caution in lactating women. The
beta-adrenergic blocking drug can be tapered as the thyrotoxic phase resolves. If
symptoms are mild and transient, then the hypothyroid phase can also be observed
without treatment. If the hypothyroid stage is severe or prolonged, thyroxine should
be administered for 6 to 12 months. After several months, the thyroid hormone
can be withdrawn and the serum TSH measured to see if the patient is euthy-
roid. Even if full thyroid recovery occurs, patients with a history of postpartum
thyroiditis should be followed long-term for possible development of permanent
hypothyroidism. Some authors prefer to keep women on thyroid hormone therapy
until they are finished having children. Since postpartum thyroiditis can recur in
up to 80% of subsequent pregnancies, future pregnancies should also be moni-
tored (75). Negro et al. treated 85 euthyroid antibody positive women in the first
trimester of pregnancy with selenium, 200 ␮g daily starting at 12-weeks gestation,
versus placebo. Postpartum thyroiditis developed significantly less frequently in
the women administered selenium than in women given placebo (28.6 vs. 48.6%
p ⬍ 0.01) (78). This study will require confirmation and determination of possible
adverse effects before selenium can be recommended.

Acute Infectious Thyroiditis


Introduction
Infectious thyroiditis is an inflammatory process caused by invasion of the thyroid
by bacteria, mycobacteria, fungi, protozoa, or flatworms. Infectious thyroiditis
may rarely cause thyrotoxicosis.
110 Meek and Smallridge

Epidemiology
Infectious thyroiditis is a rare disorder. The thyroid is felt to be relatively resistant
to infection because of its vascularity, its large concentration of iodine, the presence
of hydrogen peroxide, and its encapsulation. Infectious thyroiditis may be more
prevalent in the pediatric age group (79).

Pathophysiology
Many different bacteria can infect the thyroid including Streptococcus, Staphylo-
coccus, Pneumococcus, Salmonella, Bacteroides, Pasteurella, (80) and Treponema
pallidum (81). Mycobacterium tuberculosis (82) and several fungi, including Coc-
cidioides immitis, Aspergillus, and Candida albicans (83), have been associated
with thyroiditis. Pneumocystis carinii may also cause infectious thyroiditis (84).
Patients who are immunocompromised or have acquired immunodeficiency syn-
drome are at particular risk for infectious thyroiditis.
Most often, this infection is caused by a direct extension of an internal
fistulous tract between the pyriform sinus and the thyroid (85,86). This tract
is more common in children and may represent the course of migration of the
ultimobranchial body from its embryonic origin in the fifth pharyngeal pouch. This
extension tends to develop more commonly in the left thyroid lobe than in the right.
However, infection in the thyroid may occur in a normal thyroid, multinodular
goiter, or in a degenerating thyroid nodule as well. Immunocompromised patients
may have a higher risk of infectious thyroiditis (87). Infectious thyroiditis has
also been reported to occur after FNA by staphylococcus aureus in a patient with
atopic dermatitis (88).

Diagnosis
Patients with infectious thyroiditis usually present with pain and may have a
swollen, hot, and tender thyroid (94%) (81). As a result, affected individuals may
avoid extension of their neck due to pain, swallowing may be painful, and dyspha-
gia may be present (91%) (81). They may also present with signs of infection in
adjacent tissues, cervical lymphadenopathy, and systemic signs of fever and chills
(92%) (81).
Laboratory data include an increased white blood cell count and increased
sedimentation rate. The patient may have increased thyroid hormone levels and
present with symptoms of thyrotoxicosis, due to hormonal release from the thy-
roid (80,89). In one review, 12 of 56 cases had laboratory data suggesting hyper-
thyroidism (81). However, most patients are biochemically euthyroid, and the
radioactive iodine uptake will usually be normal. Thyroid ultrasound or computed
tomography (CT) scan of the neck may show a local abscess that can be aspirated
and cultured to make the diagnosis (90). A barium swallow can be obtained to
evaluate for possible predisposing factors such as a fistulous tract between the
pyriform sinus and the thyroid.
Thyroiditis and Other Forms of Hyperthyroidism 111

Treatment
Treatment depends on the identification of the organism causing the infection.
Aspiration of the thyroid should be obtained with appropriate Gram stain and
culture of the material. Systemic antibiotics, which are tailored to the specific
infectious agent, are administered (79). An abscess will require surgical explo-
ration and drainage, and fistulae also require surgery to prevent recurrent infection
(91).

Radiation Thyroiditis
Introduction
Radioactive 131 I and external beam radiation are used to treat thyroid disease.
Radiation thyroiditis with a thyrotoxic phase has been reported following radiation
treatment with both forms of radiation therapy.
Epidemiology
Radiation thyroiditis from 131 I occurs in approximately 20% of patients receiving
≥50,000 rads (50 Gy) to ablate residual normal thyroid tissue (92). It is more
common with larger thyroid remnants. Transient increases in thyroid hormone
are commonly seen in hyperthyroid patients treated with radioactive iodine. How-
ever, clinically significant exacerbation of hyperthyroidism was not observed (93).
Radiation thyroiditis causing transient thyrotoxicosis has also been reported with
external beam radiation. In a prospective study of external beam radiation directed
to the neck for metastatic cancer treatment, eight of 22 patients developed a sub-
normal TSH after receiving 40 Gy of external beam radiation over two weeks.
Levels of T4 and T3 tended to rise after 40 Gy of radiation, but the levels were
not statistically different from baseline (94). Several case reports of external beam
radiation–induced overt thyrotoxicosis have been reported (95,96).
Pathophysiology
Radiation presumably causes a destructive thyroiditis with release of preformed
thyroid hormone into the bloodstream. The thyrotoxicosis is transient. The radi-
ation dose that has been reported to cause external beam radiation–induced thy-
rotoxicosis varies between 37 and 50 Gy (95,97). It is likely that the greater
the external beam radiation dose, the more frequently thyrotoxic thyroiditis and
subsequent hypothyroidism occur.
Diagnosis
In radioactive iodine–induced thyroiditis, manifestations usually occur about four
days after radioactive iodine is administered. Symptoms, if present, consist of neck
and ear pain, dysphagia, thyroid tenderness, and/or transient symptoms of thyro-
toxicosis. External beam radiation–induced thyrotoxic thyroiditis usually occurs
within a few weeks of radiation exposure. It is characterized by increased serum
112 Meek and Smallridge

levels of thyroid hormones and suppressed serum levels of TSH. Serum thyroid
autoantibodies are typically negative. The 24-hour radioactive iodine uptake is
low.
Treatment
Since the thyrotoxicosis from radiation thyroiditis is transient, observation may be
all that is needed. Treatment with beta-adrenergic blocking agents can be used to
control tachycardia and tremor. Patients with 131 I-induced thyrotoxic thyroiditis
may have significant neck pain requiring treatment with corticosteroids, especially
in thyroid cancer patients treated with large doses to ablate remnant thyroid tissue.
After patients are treated with external beam radiation therapy, they should be
monitored long-term for the development of hypothyroidism.
Trauma-Induced Thyroiditis
Several reports of trauma-induced thyroiditis have been described, and this condi-
tion may be associated with thyrotoxicosis. Thyroid biopsy, parathyroid surgery,
surgical trauma, and trauma induced by a seat belt have all been reported to
cause thyrotoxicosis (98–100). The thyroid may be tender due to the trauma. The
thyrotoxicosis is transient and associated with a low uptake of radioactive iodine.

DRUG-INDUCED HYPERTHYROIDISM
Iodine-Induced Hyperthyroidism
Introduction
Iodine-induced hyperthyroidism was first described in 1821 (101). However,
iodine-induced hyperthyroidism is not a single entity, but rather an end result
that can occur from a variety of underlying thyroid problems. The common pre-
cipitant of the thyrotoxicosis is exposure to high doses of iodine.
Epidemiology
The incidence of iodine-induced hyperthyroidism has varied according to the
underlying thyroid abnormality, the geographic region, and the particular time in
history. The highest incidence of iodine-induced hyperthyroidism occurs in areas
of iodine deficiency and in individuals with multinodular goiters, when iodine
supplementation is introduced into their diet.
In four different iodine-deficient European regions, the incidence of iodine-
induced hyperthyroidism was determined before and after iodinization of salt. In
a region of Holland, the average yearly incidence of iodine-induced hyperthy-
roidism increased from 0.001% to 0.02% after introduction of dietary iodine. In
Yugoslavia, the incidence of newly diagnosed hyperthyroidism increased from
0.0025% to 0.01% in Serbia and from 0.02% to 0.07% in Belgrade. In Tyrol,
the number of toxic adenomas doubled to an average incidence of iodide-induced
hyperthyroidism of 0.03%. In general, the incidence of hyperthyroidism rose
Thyroiditis and Other Forms of Hyperthyroidism 113

140

120

100

80
<=39Yr
>=40Yr
60

40

20

0
60
62
64
66
68
70
72
74
76
78
80
82

84
86
88
90
92
19

Figure 5 Incidence of hyperthyroidism in Tasmania. Ordinate is number of cases, abscissa


is year. Separate lines for patients younger and older than 40 years. (Supplied by G. Vidor.)
Source: From Ref. 101.

within about six months after introduction of a salt iodinization program, and
reached a peak after one to three years. The incidence of hyperthyroidism returned
to baseline about three to 10 years after iodinization began (102).
The best-documented epidemic of iodine-induced hyperthyroidism occurred
in Tasmania and is shown in Figure 5. The incidence of hyperthyroidism increased
after the introduction of iodophors for sanitation in the dairy industry in 1963.
Potassium iodate was also introduced as a bread dough conditioner. Patients with
endemic goiters who were older than 40 years were most likely to develop thyro-
toxicosis (101).
Pathophysiology
Patients with underlying abnormal thyroids, particularly those with nodular goi-
ter, are predisposed to iodine-induced thyrotoxicosis. Iodine-induced thyrotoxi-
cosis can occur from iodine supplementation or pharmacologic doses of iodine
in iodine-sufficient or deficient regions. However, iodine-induced thyrotoxicosis
occurs much more commonly in iodine-deficient areas. The term Jod-Basedow
phenomenon has been used to describe the condition of thyrotoxicosis produced by
iodine exposure: Jod is German for iodine, and Basedow is often credited with the
first description of thyrotoxicosis on the European continent. Some studies have
suggested that the dose of iodine may influence the development of iodine-induced
hyperthyroidism. Dietary iodine supplements providing ⬍50 ␮g/day are gener-
ally considered safe, whereas doses of 200 to 500 ␮g/day are capable of causing
hyperthyroidism in patients with abnormal thyroid physiology (103). However,
114 Meek and Smallridge

Table 1 Iodine-Containing Drugs and Their Brand Names


Radiologic contrast agents
Ipodate (Oragrafin)
Iopanoic acid (Telepaque)
Iothalamate (Angio-Conray)
Metrizamide (Amipaque)
Topical agents
Povidone iodine (Betadine)
Iodoform gauze (Nu gauze)
Iodochlorhydroxyquin cream (Vioform)
Diiodohydroxyquin cream (Vytone)
Solutions
Saturated potassium iodide
Lugol’s
lodinated glycerol (Organidin, Tuss Organidin, lophen)a
Calcium iodide (Calcidrine syrup)
Echothiophate iodide ophthalmic (Phospholine)
Hydriodic acid syrup
Drugs
Amiodarone (Cordarone)
Vitamins containing iodine
Kelp
Iodochlorohydroxyquinolone (Entero-Vioform)
Food coloring containing iodine
a Iodine was removed from Organidin and Tuss Organidin in 1995.

even large doses of iodine (1 g/day) do not usually cause hyperthyroidism. Many
different iodine sources have been reported to cause iodine-induced hyperthy-
roidism. Compounds such as potassium iodide and iodoquinolones, which release
iodine rapidly, seem to cause hyperthyroidism less frequently than do other com-
pounds, such as amiodarone, which can release high levels of iodine for long
periods of time (104).
A large number of different iodine-containing substances and drugs have
been reported to cause hyperthyroidism. These iodinated substances include sea-
weed (105), iodinated glycerol (106,107), and topical povidone iodine (108,109).
Many cases of hyperthyroidism induced by iodinated contrast agents have also
been described (110). In general, it is prudent to avoid administering large doses
of iodine to patients with known nontoxic multinodular goiters, since doing so can
induce hyperthyroidism (111). Table 1 lists various drugs that contain iodine.

Diagnosis
Iodine-induced hyperthyroidism occurs in older patients more commonly than in
children, and it can lead to serious morbidity in the elderly. Men can be affected as
often as women can be affected. Exophthalmos is usually absent. The thyroid may
Thyroiditis and Other Forms of Hyperthyroidism 115

be nodular, diffusely enlarged, or normal. Serum TSH levels are suppressed and
serum thyroid hormone levels are increased. Thyroid autoantibodies are usually
absent. The 24-hour radioactive iodine uptake is usually low at the time of diag-
nosis of thyrotoxicosis in iodine-sufficient regions. However, radioactive iodine
uptake may be normal or high, particularly in iodine-deficient regions, which may
represent the onset of Graves’ disease or toxic, multinodular goiter after iodine
exposure or supplementation. Thyroid scans are often poorly visualized due to low
iodine uptake in iodine-sufficient regions, although they may show patchy areas
of iodine uptake in iodine-deficient regions. Urinary iodine levels are increased,
confirming the exposure to excess iodine.
Therapy
Iodine-induced hyperthyroidism may resolve spontaneously over time, usually
within a few weeks to a few months. Observation or beta-blocker therapy may be
all that is necessary in mild cases. Antithyroid drugs have been used, but they are
not uniformly effective (102,104). Corticosteroids can be effective in promptly
lowering thyroid hormone levels (102,104). Radioactive iodine has been used, but
high doses may be needed since the 24-hour uptake is usually low. Thyroidectomy
has occasionally been required to treat iodine-induced hyperthyroidism.

Amiodarone-Induced Hyperthyroidism
Introduction
Amiodarone is an iodine-rich benzofuran derivative that is used to treat supraven-
tricular and ventricular arrhythmias. Approximately 37% of amiodarone, by
weight, is organic iodine. Treatment with typical doses of amiodarone leads to a
major expansion of the total-body iodine pool (112). Amiodarone, therefore, can
have dramatic effects on thyroid function and its use has been associated with
hypothyroidism as well as hyperthyroidism.
Epidemiology
The prevalence of amiodarone-induced hyperthyroidism and hypothyroidism
varies geographically and seems to correlate with dietary intake of iodine. The
prevalence of AIT has been reported to be 1% to 23% (113). In West Tuscany, Italy,
where iodine intake is low, the prevalence of amiodarone-induced hyperthyroidism
was reported to be 9.6%, while hypothyroidism was reported in 5% of exposed
patients. On the other hand, in Worcester, Massachusetts, where iodine intake
is sufficient, amiodarone-induced hyperthyroidism was reported in only 2% of
patients, while hypothyroidism occurred in 22% (114). In Los Angeles, Califor-
nia, hyperthyroidism associated with amiodarone was found in 3% of patients,
while hypothyroidism was found in 8% (115) of patients. In a retrospective
review, amiodarone-induced hyperthyroidism was found in 4.2% of patients seen
at the Cleveland Clinic (116). However, in an area of moderately sufficient iodine
intake, the incidence of amiodarone-induced hyperthyroidism was 12.1% and the
116 Meek and Smallridge

Table 2 Amiodarone-Induced Thyrotoxicosis


Feature Type 1 Type 2

Thyroid abnormality Graves’ disease, multinodular Destructive thyroiditis


goiter, autonomous nodule
Pathogenesis Thyroid hormone production Thyroid hormone release
Thyroid examination Diffuse or nodular goiter Normal or small goiter
Radioactive iodine Low, normal, or increased Very low
uptake
Thyroid antibodies Increased or negative Negative
Interleukin-6 Normal to high (⬍200 fmol/L) Normal or high (⬎250 fmol/L)
Thyroid ultrasound Increased Decreased
Doppler flow
Therapy options Stop amiodarone Corticosteroids
ATD, KClO4 , thyroidectomy Stopping amiodarone may not
be necessary

Abbreviations: ATD, antithyroid drugs; KClO4 , potassium perchlorate.

incidence of amiodarone-induced hypothyroidism was 6.9% (117). A United


States prospective study of atrial fibrillation patients treated with amiodarone,
reported 30.8% developed hypothyroidism, primarily subclinical with TSH 4.5 to
10, compared to 6.9% of controls. Hyperthyroidism occurred in 5.3% of patients,
all but one was subclinical, compared to 2.4% of controls (p = 0.07) (118).
Pathophysiology
Two major forms of AIT have been described. The characteristics of each form are
outlined in Table 2. Type 1 occurs in patients who have an underlying abnormal
thyroid, such as a nodular goiter or latent Graves’ disease. Type 2 AIT occurs in
patients who have a normal thyroid gland prior to amiodarone treatment; it is most
likely a destructive thyroiditis (119–121). This has also been suggested by FNA
biopsy (122).
In patients with type 1 AIT, the radioactive iodine uptake is inappropriately
normal or even increased in the presence of high levels of iodine, at least in some
European studies. For example, in 12 patients with type 1 AIT, 24-hour radioactive
iodine uptake ranged from 6% to 50% (mean 17%) (123). In another study, nine
of 11 patients with diffuse goiter and eight of 12 patients with nodular goiters and
AIT had 24-hour radioactive iodine uptake of greater than 8% (124). In patients
with type 2 AIT, 24-hour radioactive iodine uptakes were very low (123–125).
Twelve patients with type 2 AIT had 24-hour radioactive iodine uptakes of 0.5%
to 2%, with a mean of 1% (123).
Interleukin-6 (IL-6) levels have also been used to distinguish type 1 from
type 2 AIT, since IL-6 is a marker of inflammation. In theory, IL-6 levels are only
mildly increased in patients with type 1 AIT, similar to what is seen in patients
with traditional forms of spontaneous hyperthyroidism (119). In contrast, IL-6
Thyroiditis and Other Forms of Hyperthyroidism 117

levels in patients with type 2 AIT have been markedly increased in some studies,
due to the release of IL-6 produced by destroyed thyrocytes (119). However, the
utility of IL-6 levels in the differential diagnosis of AIT has not been confirmed
by other authors (126).
Continuous-flow Doppler ultrasound has also been used to help distinguish
type 1 from type 2 AIT. Patients with type 1 AIT show normal to increased
parenchymal blood flow in the thyroid, while patients with type 2 AIT show
decreased blood flow consistent with thyroid inflammation (127,128). Although
reliable data are not available from the United States, most authorities agree that
type 2 AIT is far more common than type 1 AIT.

Diagnosis
The presentation of AIT may be subtle, with relatively few clinical signs (129).
However, patients typically present with tachycardia, tremor, weight loss, nervous-
ness, or irritability. One review found weight loss as the most common presenting
symptom, with goiter and tremor being the most common presenting signs (127).
Some patients may present with recurrence of the arrhythmia that was once con-
trolled with the use of amiodarone (115,130). However, tachycardia in AIT may
not always be present, due to the beta-blocking properties of amiodarone.
Amiodarone treatment itself leads to increases in serum T4 and free T4 levels
and decreases in serum T3 levels by inhibiting type 1 iodothyronine deiodinase.
Soon after therapy commences, serum TSH levels increase due to inhibition of T4
to T3 deiodination in the pituitary. Subsequently, within two to six months, serum
TSH levels return to normal. When thyrotoxicosis develops, there is a further
increase in T4 levels and an increase in T3 levels in most patients (120). Serum T3
levels may only be in the upper range of normal, but they are higher than before
the onset of hyperthyroidism. TSH levels are decreased (116).

Treatment
The treatment of AIT is often difficult and protracted because of the long 100-day
half-life of amiodarone and the high intrathyroidal and tissue concentrations of
iodine. If amiodarone is discontinued, it may take up to eight months for thy-
rotoxicosis to subside (104). Nevertheless, stopping the drug in type 1 AIT is
recommended. The appropriate medical treatment of AIT depends on making the
distinction between type 1 and type 2 forms of thyrotoxicosis. Type 1 amiodarone-
induced disease has traditionally been treated with large doses of antithyroid drugs
and potassium perchlorate. However, potassium perchlorate is no longer available
in the United States. Type 2 disease, in contrast, is treated with corticosteroids.
Sometimes, when the diagnosis is uncertain or when one treatment fails, the thera-
pies are used in combination. The high concentration of circulating iodine from the
amiodarone suppresses the uptake of radioactive iodine, which means it generally
cannot be used as a treatment option. In patients with type 1 disease, amiodarone
therapy should be discontinued if at all possible, although in many cases this
118 Meek and Smallridge

may not be feasible. In contrast, type 2 disease resolves even when amiodarone is
continued.
Patients with type 1 AIT usually respond to an antithyroid drug to decrease
thyroid hormone synthesis. Large doses, for example 40 to 80 mg/day of methi-
mazole or 400 to 800 mg/day of propylthiouracil, are often required because the
high, intrathyroidal iodine content renders the hyperthyroidism less responsive to
thionamide drugs. In areas where potassium perchlorate is available, doses of 200
to 1000 mg/day have been used to decrease intrathyroidal iodine content (123). In
general, it may take one to three months before thyroid function has normalized.
Since the side-effects of antithyroid drugs are dose-related, and since perchlorate
therapy can cause aplastic anemia in higher doses than recommended here, it
seems prudent to monitor white blood cell counts in patients receiving both drugs
together.
When it is very mild, patients with type 2 AIT may not require therapy at
all, since it usually resolves spontaneously. All but the mildest cases are treated
with prednisone 40 to 60 mg/day for one to two months with subsequent taper-
ing of the prednisone over three months. In classic type 2 amiodarone-induced
thyrotoxicosis, elevated levels of free T4 and T3 normalize within 7 to 10 days.
Care must be taken not to taper the corticosteroids too quickly, since there can be
a rapid recrudescence of hyperthyroidism (123). In a recent study, patients with
serum free T4 concentrations ⬍50 pg/mL and lower thyroid volumes achieved
more rapid control of hyperthyroidism when treated with prednisone compared to
patients with higher serum free T4 levels and larger thyroid glands (131–133). If
patients do not respond to corticosteroids alone or if the pathogenesis is unclear, a
combination of corticosteroids and antithyroid drugs would be reasonable. In one
small study, lithium in combination with antithyroid drugs was reported to help
control thyrotoxicosis faster than treatment with antithyroid drugs alone (132). The
use of the oral cholecystographic agents, sodium ipodate or sodium iopanoate, in
combination with antithyroid drugs have also been reported to be effective in type
2 AIT (133), but these drugs are not available in the United States.
Some patients have continuing hyperthyroidism despite combination therapy
with prednisone and antithyroid drugs. If drug therapy is unsuccessful after several
months, total thyroidectomy should be considered, although these patients are
often poor operative candidates because of their underlying cardiovascular disease.
Of course, the surgical risks and benefits must be weighed against the risks of
continuing hyperthyroidism (134–136).
Patients with type 2 AIT have a risk of developing transient or permanent
hypothyroidism after resolution of the thyrotoxicosis, similar to other forms of
thyroiditis, and periodic surveillance of thyroid status is recommended (137).
Although no guidelines exist regarding screening for amiodarone-induced thyroid
dysfunction, some authors recommend a baseline thyroid examination, TSH, free
T4 and T3, anti-TPO antibodies, along with monitoring of TSH and free T4 every
three months (138).
Thyroiditis and Other Forms of Hyperthyroidism 119

Cytokine-Induced Hyperthyroidism

Introduction
Interferon-alpha has been used to treat chronic viral hepatitis and certain neoplasms
with promising results. However, the immune-mediated effects of interferon have
been shown to increase the frequency of autoimmune thyroid diseases and other
immune-mediated disorders. Interferon-alpha has been most often shown to cause
hypothyroidism, but hyperthyroidism can also be induced by interferon-alpha
therapy.

Epidemiology
Interferon-alpha has been shown to induce thyroid dysfunction in approximately
6% of patients. The majority (approximately 4%) develop hypothyroidism, but
2% develop hyperthyroidism (139). Twenty percent of patients with multiple scle-
rosis treated with interferon-beta developed hypothyroidism and 4.9% developed
hyperthyroidism, which was transient (140). Females seem to be more suscepti-
ble to interferon-induced hyperthyroidism than males, likely because of the higher
background prevalence of thyroid autoantibodies in females. Hyperthyroidism has
been reported to develop as early as six weeks after the onset of interferon ther-
apy (141), to as late as six months after interferon therapy was completed (142).
Some reports indicate hyperthyroidism is more often transient, most likely due to
thyroiditis (139,143), while other reports indicate hyperthyroidism is more likely
permanent due to Graves’ disease (144–146). Interleukin-2 (IL-2) therapy has also
been reported to cause hypothyroidism and transient hyperthyroidism (147).

Pathophysiology
Thyroid dysfunction in association with interferon is more common among
patients who have circulating thyroid antibodies prior to treatment. In a meta-
analysis of the literature, thyroid dysfunction occurred in 46% of patients with
baseline thyroid antibody positivity, but only in 5% of thyroid antibody–negative
patients (139). Approximately 9% of thyroid antibody–negative patients develop
thyroid autoantibodies during interferon therapy, and 42% of these patients develop
thyroid dysfunction (139) suggesting that the thyroid dysfunction is mediated
through an immune mechanism.
However, some patients with thyroid dysfunction have no evidence of thy-
roid antibodies, raising the possibility of a direct toxic effect of interferon on the
thyroid (148). Patients with either hepatitis C or malignancy have been found
to have a higher frequency of thyroid abnormalities in the absence of interferon
therapy, which could also partly explain the higher rate of thyroid dysfunction
when interferon is used to treat these conditions (149).
120 Meek and Smallridge

Diagnosis
Hyperthyroidism induced by interferon therapy presents with the usual signs
and symptoms. The thyroid may be normal in size or mildly enlarged, serum
thyroid hormone levels are increased, with a suppressed serum TSH concentration.
In addition to thyroid peroxidase and thyroglobulin antibodies, some patients
develop circulating anti-TSH receptor antibodies (TSAb) and have normal to
increased radioactive iodine uptakes. These patients may have Graves’ disease
precipitated by cytokine therapy. However, most hyperthyroid patients develop
transient hyperthyroidism associated with low radioactive iodine uptakes, which
may be followed by hypothyroidism.

Treatment
Interferon therapy–induced hyperthyroidism, when associated with a normal or
increased radioactive iodine uptake, may have to be treated with the usual forms
of therapy, such as antithyroid drugs or radioactive iodine. Discontinuation of the
interferon therapy will not reliably lead to resolution of the hyperthyroidism.
Although patients with interferon-induced destructive thyroiditis with low
radioactive iodine uptake have been treated with corticosteroids (150), a recent
trial showed corticosteroids were not more effective than simple withdrawal of
interferon therapy in restoring euthyroidism (151). However, transient thyrotox-
icosis may give rise to hypothyroidism, which may require thyroid hormone
replacement. Finally, withdrawal of interferon therapy may be associated with
resolution of the thyroid dysfunction, and symptomatic therapy may be all that
is needed (152). Some authors have suggested screening patients with TSH, free
T4, thyroid antibodies, and perhaps thyroid ultrasound before interferon therapy
is started, in addition to monitoring thyroid levels periodically during treatment
(148,150).

Lithium-Associated Thyrotoxicosis
Lithium, commonly used to treat manic depressive disorders, is a recognized
cause of hypothyroidism. However, some reports have suggested that lithium may
also be associated with thyrotoxicosis. One study reported 14 cases of thyro-
toxicosis associated with lithium use (153). The majority of these patients had
diffuse toxic goiters. A few had toxic multinodular goiters and a few had pain-
less thyroiditis associated with a low uptake of radioactive iodine on scan. These
authors suggested that the incidence of lithium-associated thyrotoxicosis was three
times higher than predicted based on regional thyrotoxicosis rates. Others have
reviewed the cases of lithium-associated thyrotoxicosis, and they have also con-
cluded that the association was not coincidental (154,155). The majority of cases
of lithium-associated thyrotoxicosis and increased radioactive iodine uptakes have
been successfully treated with antithyroid drugs, even during continued lithium
therapy.
Thyroiditis and Other Forms of Hyperthyroidism 121

Hyperthyroidism Due to Exogenous Thyroid Hormone


Introduction
Thyrotoxicosis factitia or factitious thyrotoxicosis describes a condition due to
the excess use of exogenous thyroid hormone, leading to symptoms and signs of
thyrotoxicosis. The term is used most commonly when the use of thyroid hormone
is surreptitious, but may also be applied in cases where the exposure to thyroid
hormone is inadvertent.

Epidemiology
Thyroid hormone has been used for a variety of nonthyroid conditions in the
past, ranging from obesity to depression to infertility. Thyrotoxicosis occurs when
the thyroid hormone dose is escalated to supraphysiologic doses. The secretive
use of thyroid hormone by psychiatrically disturbed patients is another common
cause of thyrotoxicosis factitia. Thyrotoxicosis factitia has been noted in young
or older women, with a middle-aged predominance (156). Occasionally, patients
such as children will present with thyrotoxicosis due to accidental ingestion of
thyroid hormone (157). Thyrotoxicosis has been reported in a patient taking herbal
supplements that contained thyroid hormone (158).

Pathophysiology
Thyrotoxicosis factitia–induced thyrotoxicosis is due to the excessive use of a
variety of thyroid preparations. These preparations can include L-Thyroxine, tri-
iodothyronine, or desiccated thyroid tablets. Thyrotoxicosis induced by excessive
thyroid hormone has also been caused by “hamburger thyrotoxicosis.” Two epi-
demics of thyrotoxicosis in the United States were caused by bovine thyroid gland,
which was included in hamburger made by a meat processor (159,160).

Diagnosis
Patients with thyrotoxicosis factitia present with the usual symptoms and signs
of thyrotoxicosis. However, the thyroid gland is normal to small in size and the
patient lacks the eye signs of Graves’ ophthalmopathy. The thyroid is not tender.
Serum T4 and T3 levels are increased if the patient is consuming thyroxine or
desiccated thyroid. T3 levels are increased, while T4 levels are low if the patient is
consuming triiodothyronine. The radioactive iodine uptake is low and scans show
no evidence of functioning thyroid tissue elsewhere in the body. Low thyroglob-
ulin measurements are useful in distinguishing thyrotoxicosis factitia from other
forms of hyperthyroidism with low uptake of radioactive iodine (161). However,
occasionally the test may not be reliable because of circulating antithyroglobulin
antibodies. The finding of increased levels of thyroxine in the stool has also been
used to confirm the diagnosis of thyrotoxicosis factitia (162).
122 Meek and Smallridge

Treatment
The treatment of thyrotoxicosis factitia is discontinuation of the exogenous thyroid
hormone. This can sometimes be difficult in the psychiatrically disturbed patient
who is taking thyroid hormone secretly; psychiatric consultation may be helpful.
Beta-blocking agents may be needed temporarily while thyroid hormone levels
are decreasing.

THYROTROPHIN-INDUCED HYPERTHYROIDISM
TSH-Secreting Pituitary Adenomas
Introduction
Since the advent of the TSH radioimmunoassay, TSH-producing pituitary tumors
have been recognized as a cause of hyperthyroidism. Earlier reports described
large pituitary tumors, but microadenomas are being recognized more frequently
in recent case series.
Epidemiology
TSH-producing pituitary tumors (TSHomas) are rare, occurring in about one
per one million people (163). However, the number of cases being reported has
increased, along with the introduction of second- and third-generation TSH assays.
TSH-producing pituitary tumors may account for 0.5% to 3% of pituitary tumors
(164–166). TSH-producing pituitary tumors may occur at any age, with males and
females affected about equally (163,167).
Pathophysiology
TSHomas are macroadenomas approximately 90% of the time in older series, but
up to 25% are microadenomas in more recent series (168). Approximately 30%
of TSH-producing pituitary tumors are mixed tumors secreting other pituitary
hormones. Growth hormone and prolactin are the most common hormones that
are cosecreted, but tumors that cosecrete LH and FSH have also been described
(163,167). The alpha subunit of the pituitary glycoprotein hormones may also be
produced in TSH-producing pituitary adenomas. Two cases of an ectopic TSH-
secreting pituitary tumor have also been reported (169,170). Tumor-derived TSH
may have varying degrees of glycosylation along with varying molecular weight,
which may give rise to variable biologic activity of the TSH (171). In contrast to
some growth hormone-secreting tumors, no activating mutation in the gene coding
for G proteins has been found in TSH-producing pituitary tumors (163).
Diagnosis
Patients with TSH-producing pituitary adenomas present with the usual signs
and symptoms of thyrotoxicosis, but these may be mild. Most patients have a
goiter that is diffusely enlarged. Exophthalmos has been reported rarely. In one
Thyroiditis and Other Forms of Hyperthyroidism 123

Table 3 Characteristics of TSH-Producing Pituitary Tumors (TSHoma) Versus Thyroid


Hormone Resistance

Feature TSHoma (%) Thyroid resistance (%)

TRH stimulation of TSH 8 96


T3 suppression of TSH 12 100
Elevated sex hormone–binding globulin 94 2
Elevated alpha subunit/TSH ratio 81 2
Family history 0 82
MRI pituitary adenoma 98 2

Source: Percentages taken from Ref. 144 (163).

case, it was due to orbital invasion by tumor, while Graves’ disease has been
present in several patients with coexisting TSHomas. Other manifestations of
Graves’ disease, such as dermopathy and acropachy are not found. If the tumor is
large, headache (23%) and visual field disturbance (40%) may be present (163).
Features of cosecreted hormones may be present, such as acromegaly (15%)
or the galactorrhea/amenorrhea (10%) syndrome (163). Hypopituitarism, most
commonly hypogonadism, may also be present.
Laboratory evaluation shows increased levels of thyroid hormones in the
presence of an inappropriately normal or increased serum level of TSH. A pituitary
adenoma should be distinguished from pituitary thyroid hormone resistance; both
conditions give rise to increased thyroid hormone levels with inappropriately nor-
mal or increased serum TSH levels. Table 3 compares the characteristics of TSH-
producing pituitary adenomas and thyroid hormone resistance. Thyroid autoan-
tibodies are usually not present, and the radioactive iodine uptake is increased.
Circulating levels of alpha subunit are increased in macroadenomas, but are usu-
ally normal in microadenomas (172). The alpha subunit/TSH molar ratio, which is
the molar concentration of serum alpha subunit divided by the molar concentration
of serum TSH, is usually ⬎1 [molar ratio = (alpha subunit ng/mL/TSH mU/L) ×
10]. However, the alpha subunit/TSH molar ratio is not reliable in postmenopausal
women and in men with primary hypogonadism since increased serum levels of
alpha subunit, due to increased serum levels of gonadotrophins, are present in
these conditions (163,167,173–175). An increased level of sex hormone–binding
globulin and bone carboxyterminal cross-linked telopeptide of type 1 collagen
have been proposed to help differentiate patients with TSH-producing adenomas
from those with thyroid hormone resistance (172,176).
Dynamic testing has also been used to differentiate TSH-producing ade-
nomas from thyroid hormone resistance. In adenoma patients, TRH stimulation
testing usually fails to stimulate TSH secretion, and the administration of T3
also fails to suppress TSH- or TRH-stimulated TSH. More recently, treatment
with long-acting somatostatin analogues for at least two months has been used
to distinguish between cases of TSH-producing pituitary tumors versus thyroid
124 Meek and Smallridge

hormone resistance syndrome, when the diagnosis was uncertain. Long-acting


somatostatin analogues caused a greater than 30% reduction in free T4 and free
T3 levels in seven out of eight patients with TSH-producing adenomas, compared
to four patients with thyroid hormone resistance of the pituitary, in which thyroid
hormone levels did not change (177).
Imaging of the sella turcica with CT scan or magnetic resonance imaging
(MRI) will usually reveal a pituitary tumor. For patients previously treated erro-
neously with thyroid ablation due to the mistaken diagnosis of Graves’ disease, the
pituitary tumor may enlarge and become invasive. In general, there is no correla-
tion between the serum TSH level and tumor size. In several case reports, routine
contrast–enhanced MRI has been normal. In one such case, inferior petrosal sinus
sampling disclosed a gradient consistent with a TSH-producing pituitary adenoma
(178). In another case, dynamic MRI imaging identified the tumor (179).

Treatment
Transsphenoidal surgical resection of the pituitary tumor is the treatment of choice
for TSH producing pituitary tumors. Approximately 35% to 50% of patients
can be cured with surgery alone, and earlier diagnosis improves the prognosis
(163,167,168). As is true for all pituitary tumors, the most important prognostic
factors for cure are smaller size of the pituitary tumor and the absence of cav-
ernous sinus invasion. Criteria for cure following surgery include biochemical
euthyroidism with a normalized TRH test and absence of residual tumor on MRI
(164,180). Unfortunately, TRH is not available in the United States. Antithyroid
drugs and beta-blockers or octreotide should be used to restore the euthyroid state
before surgery.
Radiation therapy is used for incompletely resected pituitary tumor.
Octreotide, 50 to 100 ␮g subcutaneously two to three times daily, has been suc-
cessfully used to treat patients with TSH-producing pituitary tumors (181,182).
The long-acting somatostatin analogue lanreotide has also been used successfully
to treat patients with TSH-producing tumors (183). Octreotide produces thera-
peutic success in up to 95% of patients, and in about one-half of cases there is
tumor regression (172). In addition, octreotide alters the glycosylation pattern, and
presumably the bioactivity, of serum TSH (184). Tachyphylaxis may develop in
approximately one-fourth of patients, necessitating an increased octreotide dose
(181). Approximately 10% of patients may escape from octreotide’s inhibitory
effects on TSH suppression. Octreotide has also been used to restore euthyroidism
in pregnant women with this disorder without apparent effects on fetal develop-
ment and thyroid function (185). Dopamine agonists have also been reported to
decrease TSH and thyroxine levels in a limited number of cases, but do not cause
tumor shrinkage (186). Iopanoic acid has been used to improve hyperthyroidism
acutely in the preoperative state in two cases (187), but is not available in the
United States. Ablation of the thyroid with radioactive iodine or surgery should
be avoided. However, patients who have refused surgery for their TSH producing
Thyroiditis and Other Forms of Hyperthyroidism 125

pituitary tumors have been successfully treated with either radioactive iodine or
thyroidectomy (172,188).

Thyroid Hormone Resistance


Introduction
Thyroid hormone resistance is a heterogeneous syndrome in which tissues have a
reduced response to thyroid hormone. Generalized thyroid hormone resistance is
at one end of the spectrum of thyroid hormone resistance. These patients have a
normal metabolism because the TSH stimulation of thyroid hormone production
maintains increased thyroid hormone levels. However, some patients may exhibit
clinical signs and symptoms of hypothyroidism or hyperthyroidism in some organ
systems. At the other end of the spectrum of thyroid hormone resistance are patients
with pituitary resistance, who have near normal peripheral tissue responsiveness
to thyroid hormone. In these patients, there are clinical signs and symptoms of
thyrotoxicosis.
Epidemiology
Thyroid hormone resistance was first described in 1967 (189). Pituitary resistance
to thyroid hormone was first described in 1975 (190). Since that time, more
than 1000 cases of thyroid hormone resistance have been reported. The exact
prevalence of resistance to thyroid hormone is unknown since the condition is not
detected by routine neonatal screening for hypothyroidism. However, a screening
for high blood T4 found one case per 40,000 births (191). Thyroid hormone
resistance occurs in males and females, and it has been reported in all races. The
majority of cases of thyroid hormone resistance have generalized tissue resistance
to thyroid hormone. Thyroid hormone resistance is usually inherited, and it is
usually autosomal dominant.
Pathophysiology
Thyroid hormone resistance is most often due to a mutation in the thyroid hor-
mone receptor-beta gene (TR-␤), found on chromosome 3 (192–194). However,
15% of patients with thyroid hormone resistance do not have a TR-␤ mutation
(195). Mosaicism of the TR-␤ has been reported (196). More recently, mutations
in the cell membrane transporter of thyroid hormone, MCT8, and mutations in the
SECISBP2 gene, which is required for thyroid hormone deiodinases, have been
reported (195). Patients with thyroid hormone resistance are usually heterozy-
gous for mutations that cluster within three areas of the thyroid hormone–binding
domain (197,198). The ability of the mutant receptor proteins to bind thyroid hor-
mone is reduced, and therefore, the ability to effect gene transcription is reduced.
Analysis of patients with pituitary-only resistance to thyroid hormone with conse-
quent clinical hyperthyroidism indicates that these individuals also are heterozy-
gous for mutations in the hormone-binding region of the TR-␤ receptor (198,199).
The same mutations in the TR-␤ gene in patients with pituitary resistance have
126 Meek and Smallridge

also been identified in patients from unrelated families with generalized thyroid
hormone resistance. In addition, within a family, the same receptor mutation may
result in generalized thyroid resistance in one family member, but in another,
thyrotoxicosis suggestive of pituitary thyroid hormone resistance.
Diagnosis
Patients with thyroid hormone resistance may have clinical symptoms and signs
that vary from hypothyroidism to hyperthyroidism. There is considerable overlap
of these findings between generalized resistance patients versus those thought to
have pituitary thyroid hormone resistance.
Patients with thyroid hormone resistance have a higher frequency of attention
deficit disorder, delayed speech development, lower IQ, shorter stature and lower
weight, delayed bone age, and hearing loss. There may also be a higher frequency
of ear, nose, and throat infections (197,200,201).
A goiter is found commonly (65–90%) in patients with thyroid hormone
resistance (197,198). Tachycardia and an increased frequency of arrhythmia have
been found in some thyroid hormone resistance patients. Increased levels of thy-
roid hormones, including T4 and T3, are found with inappropriately normal or
increased levels of TSH. The 24-hour radioactive iodine uptake is often increased.
Failure of serum TSH levels to increase in response to TRH, and failure of
serum TSH levels to decrease in response to supraphysiologic doses of thyroid
hormone suggest a TSH-producing pituitary adenoma. An increased molar ratio of
alpha subunit to TSH and a pituitary adenoma on MRI of the brain are diagnostic
of a TSH-producing pituitary adenoma (202).
Treatment
Most patients with generalized thyroid hormone resistance do not require treat-
ment. Patients who have been mistakenly treated with thyroidectomy or radioac-
tive iodine ablation of the thyroid typically require higher than normal doses of
thyroxine replacement to suppress TSH back to normal. One case of pituitary
enlargement was demonstrated in a patient previously treated with radioactive
iodine ablation. Supraphysiologic doses of thyroid hormone caused regression of
the pituitary gland back to normal size (203).
Patients with pituitary resistance to thyroid hormone who have mildly symp-
tomatic hyperthyroidism may be treated with beta-blocker therapy. Antithyroid
drugs are not ideal because they result in further increases in thyroid size, but rarely
may be needed in patients with severe hypermetabolism (195). Radioactive iodine
is not recommended as it may cause dramatic increases in TSH (197). Moderate
doses of T3 (25–50 ␮g daily) over a period of several months can decrease TSH
secretion, thyroid hormone levels, and clinical thyrotoxicosis (204), but they are
not often successful. D-thyroxine has also been reported to be beneficial (205,206).
Triiodothyroacetic acid (Triac) appears to be able to suppress TSH with minimal
peripheral thyromimetic actions (207–209). Bromocriptine and octreotide have
been used in some cases to suppress TSH production (193,195,208,210,211). In
Thyroiditis and Other Forms of Hyperthyroidism 127

general, it is appropriate to treat patients with pituitary thyroid hormone resistance


with conservative measures, if possible, including beta-blockers. If patients do not
respond to these measures, antithyroid drugs may be necessary.

HYPERTHYROIDISM OF EXTRATHYROID ORIGIN


Struma Ovarii Tumor
Introduction
Struma ovarii, a very rare cause of hyperthyroidism, is due to the presence of
an ovarian teratoma that contains hyperfunctioning autonomous thyroid tissue.
An ovarian teratoma that contains greater than 50% thyroid tissue or functioning
thyroid tissue causing thyrotoxicosis is called a struma ovarii (212).
Epidemiology
Struma ovarii represents less than 2% of ovarian teratomas, with peak frequency
during the fifth decade of life. One review found struma ovarii in five of 1390
(0.4%) ovarian tumors (213). Struma ovarii tumors usually do not cause hyper-
thyroidism. One review reported that preoperatively, three out of 41 patients (7%)
with struma ovarii had clinical symptoms and laboratory signs of hyperthyroidism
(212). Hyperthyroidism was found in eight of 25 patients with struma ovarii tumor
in another review (214).
The frequency of papillary or follicular carcinoma arising in a struma ovarii
is unknown but is quite rare. A review in Colorado showed a frequency of 0.3%
(215). Pardo-Mindan and Vazquez reviewed the literature on malignant struma
ovarii and found only 45 cases of malignant struma ovarii; 17% of these cases
were associated with hyperthyroidism (216).
Pathophysiology
Struma ovarii and associated thyrotoxicosis is due to the presence of autonomous
hyperfunctioning thyroid tissue within the teratoma. Struma ovarii tumors are
unilateral in 90% of cases, with the left ovary more frequently involved (216).
Most struma ovarii tumors are benign (214). However, it is sometimes difficult to
determine if the thyroid tissue in the tumor is benign or malignant (212). Papillary
carcinomas are more commonly reported than follicular or insular carcinoma
(217). Metastatic struma ovarii may spread to the peritoneum, intraabdominal
nodes, bone, liver, lung, mediastinum, and brain (214,218). Struma ovarii can be
mixed with a carcinoid tumor and has been reported to occur in association with
multiple endocrine neoplasia type IIA (219). One case of hyperthyroidism has
been reported, which was due to a thyrotoxic adenoma and struma ovarii (220).
Diagnosis
The diagnosis of struma ovarii–causing hyperthyroidism should be suspected in a
female without thyroid enlargement, and a very low thyroid uptake of radioactive
128 Meek and Smallridge

iodine. However, the thyroid has been reported to be enlarged in several reports
(214,221,222). Some patients may present with a pelvic mass, and ascites may
be present even in the absence of malignant struma ovarii (214). The TSH is sup-
pressed and thyroid hormone levels are increased. The diagnosis is established by
finding radioactive iodine uptake over the pelvis or an ovarian teratoma containing
thyroid tissue (223). However, radioactive iodine uptake has been reported in a
hemorrhagic ovarian cyst that did not contain thyroid tissue (224).

Treatment
Treatment of struma ovarii, with or without thyrotoxicosis, consists of surgery to
remove the tumor. If thyrotoxicosis is present, beta-blockers and/or antithyroid
drugs should be used before surgery. Radioactive iodine should not be used to
ablate the thyroid tissue, since the thyroid tissue is neoplastic and potentially
malignant. It also often contains nonthyroid tissue (225), and the effect of radiation
on the other tissues is unknown.
Malignant struma ovarii is treated with hysterectomy, bilateral salpingo-
oophorectomy, and thyroidectomy. Metastatic struma ovarii has been treated with
radioactive iodine after thyroidectomy. Thyroid hormone treatment with TSH
suppression is also recommended for metastatic struma ovarii.

Trophoblastic Tumors
Introduction
Molar pregnancy and trophoblastic tumors can cause hyperthyroidism. Human
chorionic gonadotrophin (hCG), secreted in large amounts by these tumors, has
TSH receptor–binding activity. This hCG cross-reactivity can lead to thyrotoxico-
sis (226).

Epidemiology
The prevalence of thyrotoxicosis in patients with trophoblastic tumors is unknown.
One study evaluated 20 patients from a referral center over one year, and the
researchers found that five of the patients had thyrotoxicosis (227). Another study
found that 30 of 52 patients with gestational trophoblastic tumors had thyrotox-
icosis (228). It has been estimated that 20% of women with hydatidiform moles
have hyperthyroidism (229).
Hydatiform mole occurs in about one out of 1500 pregnancies in the United
States and is about 10 times more common in Asian and Latin American countries
(230). Choriocarcinoma occurs in one of 50,000 pregnancies (226). Thyrotoxicosis
is reported more frequently in women with hydatiform mole than in those with
choriocarcinoma. In addition, a few men with testicular tumors that produce hCG
have been reported with hyperthyroidism (231–233).
Thyroiditis and Other Forms of Hyperthyroidism 129

Pathophysiology
hCG is composed of an alpha subunit and a beta subunit. The alpha subunit is
identical to the alpha subunit of LH, FSH, and TSH. The beta subunit of hCG
is larger than the beta subunit of TSH but is similar in structure. Beta-hCG has
an additional 33–amino acid peptide at the carboxyl terminal. The thyrotropin
effect of hCG is weak. However, when hCG is secreted in large amounts, it can
stimulate the TSH receptor in thyroid tissue enough to cause thyrotoxicosis. It is
probable that some molecular variants of hCG secreted by trophoblastic tumors
have greater thyrotrophic activity than hCG secreted by normal placental tissue
(229,234). Removal of the tumor, and therefore the hCG, is associated with rapid
resolution of the hyperthyroidism.

Diagnosis
Women with trophoblastic tumors may or may not have clinical evidence of
hyperthyroidism. The nausea, vomiting, and toxemia that occur in molar pregnancy
may obscure hyperthyroidism. The thyroid gland is either normal in size or slightly
enlarged.
Chorionic gonadotrophin, secreted in large amounts by trophoblastic tissue,
serves as a marker for the tumor. The hCG levels exceed 100 U/mL in patients
with hyperthyroidism and often exceed 300 U/mL (235–238). TSH is suppressed
and levels of free T4 and free T3 may be minimally or markedly increased.
Hyperthyroid patients with trophoblastic tumors usually have higher T4/T3 ratios
than patients with hyperthyroidism due to Graves’ disease (157). The uptake of
radioactive iodine is increased (239).

Treatment
Surgical removal of the hydatidiform mole or choriocarcinoma in a patient with
hyperthyroidism rapidly cures the thyrotoxicosis. However, patients with chori-
ocarcinoma who have hyperthyroidism usually have a larger tumor mass that
may be metastatic. Chemotherapy is the principle therapy used to achieve remis-
sion of metastatic choriocarcinoma with associated hyperthyroidism. Effective
chemotherapy provides long-term survival ranges from 86% to 100% (240). The
prognosis for men with testicular choriocarcinoma and hyperthyroidism is usually
poor.
Medical therapy for hyperthyroidism due to trophoblastic disease may
include potassium iodide, beta-blockers, and antithyroid drugs. Preoperative
iodine can help lower thyroid hormone levels rapidly in patients who require
urgent surgery. Beta-adrenergic blockers are given to control tachycardia and
tremor. Antithyroid drugs are given to help control hyperthyroidism periop-
eratively or in patients with metastatic disease. Surgical thyroidectomy is not
recommended.
130 Meek and Smallridge

Metastatic Thyroid Cancer and Hyperthyroidism


Introduction
Differentiated thyroid cancer usually does not produce thyroid hormone efficiently.
However, thyroid cancers do produce thyroglobulin, which can become iodinated
and form thyroid hormone. Rarely thyroid hormone production from thyroid can-
cer can become excessive, giving rise to thyrotoxicosis.

Epidemiology
Thyrotoxicosis due to thyroid cancer is quite rare. However, the age and sex
distribution of patients with hyperthyroidism caused by thyroid cancer is similar
to that of patients with thyroid cancer without thyrotoxicosis (241,242). Eighty-
five percent of patients with hyperfunctioning thyroid cancer are older than 40
years. The female-to-male ratio is about 3:1 (242).

Pathophysiology
Malignant thyroid tissue is functionally less efficient than normal thyroid tissue
(243). The estimated efficacy of the iodine-concentrating ability of functioning
metastases is approximately 10% of normal thyroid tissue (244,245). The inef-
ficient thyroid hormone production is due in part to lower iodine trapping by
tumor tissue and in part to abnormal thyroglobulin synthesis. Further, there is
evidence that expression of the TSH receptor in malignant thyroid tissue may be
absent or low (246). Therefore, many of the cases of thyrotoxicosis caused by
thyroid cancer are due to large, bulky metastatic tumors, often weighing 2 to 3 kg
(242).
Follicular thyroid cancer is the most common thyroid malignancy reported
to cause hyperthyroidism (241,242,247), but papillary thyroid cancer may also
cause hyperthyroidism. Patients with functioning metastases more commonly
come from areas of low iodine intake (241). Thyroglobulin levels have been
reported to be higher in patients with functioning metastases, but this finding did
not reach statistical significance (241). Finally, the time to metastasis and the
10-year survival rate appear to be equal for metastatic follicular carcinoma with
or without thyrotoxicosis (241,242,247). The discovery of metastases precedes
or occurs simultaneously with the onset of hyperthyroidism (247). Anaplastic
thyroid cancer and thyroid lymphoma have been reported to cause thyrotoxicosis
with a low uptake of radioactive iodine, so-called “malignant pseudothyroiditis”
(20,21,248,249).

Diagnosis
In most instances, the diagnosis of thyroid malignancy has been made and thy-
roidectomy has been accomplished. Since the treatment of thyroid cancer includes
suppressive doses of thyroid hormone, the fact that metastatic disease is causing
hyperthyroidism may not be recognized if the clinician erroneously believes that
Thyroiditis and Other Forms of Hyperthyroidism 131

the thyrotoxicosis is due to overzealous treatment with thyroid hormone. There-


fore, thyroid hormone treatment should be stopped to see if the signs and symptoms
of thyrotoxicosis resolve, and if the levels of T4 and T3 in the serum decrease. If the
signs, symptoms, and levels of thyroid hormone do not decrease, hyperthyroidism
may be due to functional thyroid metastases. In many cases, the thyrotoxicosis
may be due to T3 toxicosis, with suppressed TSH, and normal or even low serum
T4 levels (242,250,251). Hyperfunctioning metastatic cancer may be confirmed
by whole-body radioactive iodine scanning. However, uptake of radioactive iodine
in metastatic tissue may be low in the presence of the normal thyroid gland. The
distribution of hyperfunctioning metastatic thyroid cancer is typically the same as
in nonhyperfunctioning thyroid cancer, that is, in bone, lung, and mediastinum.

Treatment
Treatment of metastatic functioning thyroid cancer usually consists of radioac-
tive 131 I therapy. The usual dose of radioactive iodine ranges from 100 to
200 mCi. Treatment with radioactive iodine may exacerbate the thyrotoxico-
sis (252). Therefore, radioactive iodine should be administered with caution,
and patients are often treated prophylactically with beta-adrenergic blocking
agents. Some authors recommend treating the patient with antithyroid drugs
to control the hyperthyroid state prior to administering radioactive iodine to
prevent exacerbation of the hyperthyroid state (242). If normal thyroid tissue
is present, it must be removed prior to administering a therapeutic dose of
radioactive iodine to functioning thyroid metastases. If a small number of large,
accessible, and isolated metastases are producing the thyrotoxicosis, surgical
resection may be the best treatment option, following therapy with antithyroid
drugs.

REFERENCES
1. Woolner JB, McConahey WM, Beahrs OH. Granulomatous thyroiditis (de Quer-
vain’s thyroiditis). J Clin Endocrinol Metab 1957; 17:1202–1221.
2. Fatourechi V, Aniszewski JP, Fatourechi GZ, et al. Clinical features and outcome
of subacute thyroiditis in an incidence cohort: Olmsted County, Minnesota, study.
J Clin Endocrinol Metab 2003; 88(5):2100–2105.
3. Nordyke RA, Gilbert FI Jr, Lew C. Painful subacute thyroiditis in Hawaii. West J
Med 1991; 155(1):61–63.
4. Martino E, Buratti L, Bartalena L, et al. High prevalence of subacute thyroiditis
during summer season in Italy. J Endocrinol Invest 1987; 10 (3):321–323.
5. Saito S, Sakurada T, Yamamoto M, et al. Subacute thyroiditis: Observations on 98
cases for the last 14 years. Tohoku J Exp Med 1974; 113(2):141–147.
6. Hay ID. Thyroiditis: Aclinical update. Mayo Clin Proc 1985; 60(12):836–843.
7. Hamburger JI. The various presentations of thyroiditis. Diagnostic considerations.
Ann Intern Med 1986; 104(2):219–224.
8. Singer PA. Thyroiditis. Acute, subacute, and chronic. Med Clin North Am 1991;
75(1):61–77.
132 Meek and Smallridge

9. Hung W. Mumps thyroiditis and hypothyroidism. J Pediatr 1969; 74 (4):611–613.


10. Robertson WS. Acute inflammation of the thyroid gland. Lancet 1911; 1:930–931.
11. Hintze F, Fortelius P, Railo J. Epidemic thyroiditis. Acta Endocrinol 1964; 45:381–
401.
12. Swann N. Acute thyroiditis: Five cases associated with adenovirus infection.
Metabolism 1964; 13:908–910.
13. Shumway M, Davis P. Cat-scratch thyroiditis treated with thyrotrophin hormone.
J Clin Endocrinol Metab 1954; 14:742–743.
14. Bech K, Nerup J, Thomsen M, et al. Subacute thyroiditis de Quervain: A disease
associated with HLA-B antigen. Acta Endocrinol 1977; 8:504–509.
15. Volpe R, Johnston MW. Subacute thyroiditis: A disease commonly mistaken for
pharyngitis. Can Med Assoc J 1957; 77:297–307.
16. Nyulassy S, Hnilica P, Buc M, et al. Subacute (de Quervain’s) thyroiditis: Association
with HLA-Bw35 antigen and abnormalities of the complement system, immunoglob-
ulins and other serum proteins. J Clin Endocrinol Metab 1977; 45 (2):270–274.
17. Kramer AB, Roozendaal C, Dullaart RP. Familial occurrence of subacute thyroiditis
associated with human leukocyte antigen-B35. Thyroid 2004; 14(7):544–547.
18. Hamaguchi E, Nishimura Y, Kaneko S, et al. Subacute thyroiditis developed in
identical twins two years apart. Endocr J 2005; 52(5):559–562.
19. Ikenoue H, Okamura K, Kuroda T, et al. Thyroid amyloidosis with recurrent subacute
thyroiditis-like syndrome. J Clin Endocrinol Metab 1988; 67(1):41–45.
20. Oppenheim A, Miller M, Anderson GH Jr, et al. Anaplastic thyroid cancer presenting
with hyperthyroidism. Am J Med 1983; 75(4):702–704.
21. Alagol F, Tanakol R, Boztepe H, et al. Anaplastic thyroid cancer with transient
thyrotoxicosis: Case report and literature review. Thyroid 1999; 9 (10):1029–1032.
22. Mizukami Y, Michigishi T, Kawato M, et al. Immunohistochemical and ultrastruc-
tural study of subacute thyroiditis, with special reference to multinucleated giant
cells. Hum Pathol 1987; 18(9):929–935.
23. Sanders LR, Moreno AJ, Pittman DL, et al. Painless giant cell thyroiditis diagnosed
by fine needle aspiration and associated with intense thyroidal uptake of gallium.
Am J Med 1986; 80(5):971–975.
24. Shabb NS, Salti I. Subacute thyroiditis: Fine-needle aspiration cytology of 14 cases
presenting with thyroid nodules. Diagn Cytopathol 2006; 34(1):18–23.
25. Greene JN. Subacute thyroiditis. Am J Med 1971; 51(1):97–108.
26. Rotenberg Z, Weinberger I, Fuchs J, et al. Euthyroid atypical subacute thyroiditis
simulating systemic or malignant disease. Arch Intern Med 1986; 146(1):105–107.
27. de Bruin TW, Riekhoff FP, de Boer JJ. An outbreak of thyrotoxicosis due to atypical
subacute thyroiditis. J Clin Endocrinol Metab 1990; 70(2):396–402.
28. Swinburne JL, Kreisman SH. A rare case of subacute thyroiditis causing thyroid
storm. Thyroid 2007; 17(1):73–76.
29. Weihl AC, Daniels GH, Ridgway EC, et al. Thyroid function tests during the early
phase of subacute thyroiditis. J Clin Endocrinol Metab 1977; 44(6):1107–1114.
30. Larsen PR. Serum triiodothyronine, thyroxine, and thyrotropin during hyperthy-
roid, hypothyroid, and recovery phases of subacute nonsuppurative thyroiditis.
Metabolism 1974; 23(5):467–471.
31. Van Herle AJ, Vassart G, Dumont JE. Control of thyroglobulin synthesis and secre-
tion. (First of two parts). N Engl J Med 1979; 301(5):239–249.
Thyroiditis and Other Forms of Hyperthyroidism 133

32. Volpe R. The management of subacute (de Quervain’s) thyroiditis. Thyroid 1993;
3(3):253–525.
33. Yamamoto M, Saito S, Sakurada T, et al. Effect of prednisolone and salicylate on
serum thyroglobulin level in patients with subacute thyroiditis. Clin Endocrinol (Oxf)
1987; 27(3):339–344.
34. Duininck TM, van Heerden JA, Fatourechi V, et al. de Quervain’s thyroiditis: Surgical
experience. Endocr Pract 2002; 8(4):255–258.
35. Woolf PD. Transient painless thyroiditis with hyperthyroidism: A variant of lym-
phocytic thyroiditis? Endocr Rev 1980; 1(4):411–420.
36. Schneeberg NG. Silent thyroiditis [letter]. Arch Intern Med 1983; 143(11):2214.
37. Nikolai TF, Coombs GJ, McKenzie AK. Lymphocytic thyroiditis with spontaneously
resolving hyperthyroidism and subacute thyroiditis. Long-term follow-up. Arch
Intern Med 1981; 141(11):1455–1458.
38. Nikolai TF, Brosseau J, Kettrick MA, et al. Lymphocytic thyroiditis with spon-
taneously resolving hyperthyroidism (silent thyroiditis). Arch Intern Med 1980;
140(4):478–482.
39. Gorman CA, Duick DS, Woolner LB, et al. Transient hyperthyroidism in patients
with lymphocytic thyroiditis. Mayo Clin Proc 1978; 53(6):359–365.
40. Parker M, Klein I, Fishman LM, et al. Silent thyrotoxic thyroiditis in association
with chronic adrenocortical insufficiency. Arch Intern Med 1980; 140(8):1108–
1109.
41. Nagai K, Sakata S, Takuno H, et al. A case of silent thyroiditis associated with
idiopathic thrombocytopenic purpura. Endocrinol Jpn 1988; 35(6):791–794.
42. Sakata S, Nagai K, Shibata T, et al. A case of rheumatoid arthritis associated with
silent thyroiditis. J Endocrinol Invest 1992; 15(5):377–380.
43. Magaro M, Zoli A, Altomonte L, et al. The association of silent thyroiditis with
active systemic lupus erythematosus. Clin Exp Rheumatol 1992; 10(1):67–70.
44. Farid NR, Hawe BS, Walfish PG. Increased frequency of HLA-DR3 and 5 in the
syndromes of painless thyroiditis with transient thyrotoxicosis: Evidence for an
autoimmune aetiology. Clin Endocrinol (Oxf) 1983; 19(6):699–704.
45. Smallridge RC, De Keyser FM, Van Herle AJ, et al. Thyroid iodine content and
serum thyroglobulin: Cues to the natural history of destruction-induced thyroiditis.
J Clin Endocrinol Metab 1986; 62(6):1213–1219.
46. Nikolai TF, Coombs GJ, McKenzie AK, et al. Treatment of lymphocytic thyroiditis
with spontaneously resolving hyperthyroidism (silent thyroiditis). Arch Intern Med
1982; 142(13):2281–2283.
47. Smallridge RC. Postpartum thyroid dysfunction: A frequently undiagnosed
endocrine disorder. Endocrinologist 1996; 1:44–50.
48. Fung HY, Kologlu M, Collison K, et al. Postpartum thyroid dysfunction in Mid
Glamorgan. Br Med J (Clin Res Ed) 1988; 296(6617):241–244.
49. Walfish PG, Meyerson J, Provias JP, et al. Prevalence and characteristics of post-
partum thyroid dysfunction: Results of a survey from Toronto, Canada. J Endocrinol
Invest 1992; 15(4):265–272.
50. Amino N, Mori H, Iwatani Y, et al. High prevalence of transient post-partum thyro-
toxicosis and hypothyroidism. N Engl J Med 1982; 306(14):849–852.
51. Gerstein HC. How common is postpartum thyroiditis? A methodologic overview of
the literature. Arch Intern Med 1990; 150(7):1397–1400.
134 Meek and Smallridge

52. Nikolai TF, Turney SL, Roberts RC. Postpartum lymphocytic thyroiditis. Preva-
lence, clinical course, and long-term follow-up. Arch Intern Med 1987; 147(2):221–
224.
53. Jansson R, Bernander S, Karlsson A, et al. Autoimmune thyroid dysfunction in the
postpartum period. J Clin Endocrinol Metab 1984; 58(4):681–687.
54. Freeman R, Rosen H, Thysen B. Incidence of thyroid dysfunction in an unselected
postpartum population. Arch Intern Med 1986; 146(7):1361–1364.
55. Stagnaro-Green A. Clinical review 152: Postpartum thyroiditis. J Clin Endocrinol
Metab 2002; 87(9):4042–4047.
56. Rajatanavin R, Chailurkit LO, Tirarungsikul K, et al. Postpartum thyroid dysfunction
in Bangkok: A geographical variation in the prevalence. Acta Endocrinol (Copenh)
1990; 122(2):283–287.
57. Gerstein HC. Incidence of postpartum thyroid dysfunction in patients with type I
diabetes mellitus. Ann Intern Med 1993; 118(6):419–423.
58. Alvarez-Marfany M, Roman SH, Drexler AJ, et al. Long-term prospective study of
postpartum thyroid dysfunction in women with insulin dependent diabetes mellitus.
J Clin Endocrinol Metab 1994; 79(1):10–16.
59. Bech K, Hoier-Madsen M, Feldt-Rasmussen U, et al. Thyroid function and autoim-
mune manifestations in insulin-dependent diabetes mellitus during and after preg-
nancy. Acta Endocrinol (Copenh) 1991; 124(5):534–539.
60. Stagnaro-Green A. Post-miscarriage thyroid dysfunction. Obstet Gynecol 1992; 80(3
Pt 2):490–492.
61. Hayslip CC, Fein HG, O’Donnell VM, et al. The value of serum antimicrosomal
antibody testing in screening for symptomatic postpartum thyroid dysfunction. Am
J Obstet Gynecol 1988; 159(1):203–209.
62. Solomon BL, Fein HG, Smallridge RC. Usefulness of antimicrosomal antibody
titers in the diagnosis and treatment of postpartum thyroiditis. J Fam Pract 1993;
36(2):177–182.
63. Harris B, Othman S, Davies JA, et al. Association between postpartum thyroid
dysfunction and thyroid antibodies and depression. BMJ 1992; 305(6846):152–156.
64. Stagnaro-Green A, Roman SH, Cobin RH, et al. A prospective study of lymphocyte-
initiated immunosuppression in normal pregnancy: Evidence of a T-cell etiology for
postpartum thyroid dysfunction. J Clin Endocrinol Metab 1992; 74(3):645–653.
65. Tachi J, Amino N, Tamaki H, et al. Long term follow-up and HLA association in
patients with postpartum hypothyroidism. J Clin Endocrinol Metab 1988; 66(3):480–
484.
66. Vargas MT, Briones-Urbina R, Gladman D, et al. Antithyroid microsomal autoanti-
bodies and HLA-DR5 are associated with postpartum thyroid dysfunction: Evidence
supporting an autoimmune pathogenesis. J Clin Endocrinol Metab 1988; 67(2):327–
333.
67. Kologlu M, Fung H, Darke C, et al. Postpartum thyroid dysfunction and HLA status.
Eur J Clin Invest 1990; 20(1):56–60.
68. Jansson R, Safwenberg J, Dahlberg PA. Influence of the HLA-DR4 antigen and
iodine status on the development of autoimmune postpartum thyroiditis. J Clin
Endocrinol Metab 1985; 60(1):168–173.
69. Kuijpens JL, Pop VJ, Vader HL, et al. Prediction of post partum thyroid dysfunction:
Can it be improved? Eur J Endocrinol 1998; 139(1):36–43.
Thyroiditis and Other Forms of Hyperthyroidism 135

70. Rasmussen NG, Hornnes PJ, Hoier-Madsen M, et al. Thyroid size and function in
healthy pregnant women with thyroid autoantibodies. Relation to development of
postpartum thyroiditis. Acta Endocrinol (Copenh) 1990; 123(4):395–401.
71. Lazarus JH, Hall R, Othman S, et al. The clinical spectrum of postpartum thyroid
disease. Q J Med 1996; 89(6):429–435.
72. Lervang HH, Pryds O, Ostergaard Kristensen HP. Thyroid dysfunction after delivery:
Incidence and clinical course. Acta Med Scand 1987; 222(4):369–374.
73. Adams H, Jones MC, Othman S, et al. The sonographic appearances in postpartum
thyroiditis. Clin Radiol 1992; 45(5):311–315.
74. Lazarus JH, Othman S. Thyroid disease in relation to pregnancy. Clin Endocrinol
(Oxf) 1991; 34(1):91–98.
75. Smallridge R. Postpartum thyroid disease A model of immunologic dysfunction.
Clin Appl Immunol Rev 2000; 1(2):89–103.
76. Othman S, Phillips DI, Parkes AB, et al. A long-term follow-up of postpartum
thyroiditis. Clin Endocrinol (Oxf) 1990; 32(5):559–564.
77. Azizi F. The occurrence of permanent thyroid failure in patients with subclinical
postpartum thyroiditis. Eur J Endocrinol 2005; 153(3):367–371.
78. Negro R, Greco G, Mangieri T, et al. The influence of selenium supplementation on
postpartum thyroid status in pregnant women with thyroid peroxidase autoantibodies.
J Clin Endocrinol Metab 2007; 92(4):1263–1268.
79. Brook I. Microbiology and management of acute suppurative thyroiditis in children.
Int J Pediatr Otorhinolaryngol 2003; 67(5):447–451.
80. McLaughlin SA, Smith SL, Meek SE. Acute suppurative thyroiditis caused by
Pasteurella multocida and associated with thyrotoxicosis. Thyroid 2006; 16(3):307–
310.
81. Berger SA, Zonszein J, Villamena P, et al. Infectious diseases of the thyroid gland.
Rev Infect Dis 1983; 5(1):108–122.
82. Das DK, Pant CS, Chachra KL, et al. Fine needle aspiration cytology diagnosis
of tuberculous thyroiditis. A report of eight cases. Acta Cytol 1992; 36(4):517–
522.
83. Gandhi RT, Tollin SR, Seely EW. Diagnosis of Candida thyroiditis by fine needle
aspiration. J Infect 1994; 28(1):77–81.
84. Guttler R, Singer PA, Axline SG, et al. Pneumocystis carinii thyroiditis. Report of
three cases and review of the literature. Arch Intern Med 1993; 153(3):393–396.
85. Hatabu H, Kasagi K, Yamamoto K, et al. Acute suppurative thyroiditis associ-
ated with piriform sinus fistula: Sonographic findings. Am J Roentgenol 1990;
155(4):845–847.
86. Miyauchi A, Matsuzuka F, Kuma K, et al. Piriform sinus fistula: An underlying
abnormality common in patients with acute suppurative thyroiditis. World J Surg
1990; 14(3):400–405.
87. Yu EH, Ko WC, Chuang YC, et al. Suppurative Acinetobacter baumanii thyroidi-
tis with bacteremic pneumonia: Case report and review. Clin Infect Dis 1998;
27(5):1286–1290.
88. Nishihara E, Miyauchi A, Matsuzuka F, et al. Acute suppurative thyroiditis after
fine-needle aspiration causing thyrotoxicosis. Thyroid 2005; 15(10):1183–1187.
89. Nieuwland Y, Tan KY, Elte JW. Miliary tuberculosis presenting with thyrotoxicosis.
Postgrad Med J 1992; 68(802):677–679.
136 Meek and Smallridge

90. Bernard PJ, Som PM, Urken ML, et al. The CT findings of acute thyroiditis and
acute suppurative thyroiditis. Otolaryngol Head Neck Surg 1988; 99(5):489–493.
91. Cigliano B, Cipolletta L, Baltogiannis N, et al. Endoscopic fibrin sealing of congen-
ital pyriform sinus fistula. Surg Endosc 2004; 18(3):554–556.
92. Maxon HR, Thomas SR, Saenger EL, et al. Ionizing irradiation and the induction of
clinically significant disease in the human thyroid gland. Am J Med 1977; 63(6):967–
978.
93. Koornstra JJ, Kerstens MN, Hoving J, et al. Clinical and biochemical changes
following 131I therapy for hyperthyroidism in patients not pretreated with antithyroid
drugs. Neth J Med 1999; 55(5):215–221.
94. Nishiyama K, Kozuka T, Higashihara T, et al. Acute radiation thyroiditis. Int J Radiat
Oncol Biol Phys 1996; 36(5):1221–1224.
95. Blitzer JB, Paolozzi FP, Gottlieb AJ, et al. Thyrotoxic thyroiditis after radiotherapy
for Hodgkin’s disease. Arch Intern Med 1985; 145(9):1734–1735.
96. Aizawa T, Watanabe T, Suzuki N, et al. Radiation-induced painless thyrotoxic thy-
roiditis followed by hypothyroidism: A case report and literature review. Thyroid
1998; 8(3):273–275.
97. Petersen M, Keeling CA, McDougall IR. Hyperthyroidism with low radioiodine
uptake after head and neck irradiation for Hodgkin’s disease. J Nucl Med 1989;
30(2):255–257.
98. Kobayashi A, Kuma K, Matsuzuka F, et al. Thyrotoxicosis after needle aspiration of
thyroid cyst. J Clin Endocrinol Metab 1992; 75(1):21–24.
99. Calle RA, Cohen KL. Transient thyroiditis due to surgical trauma. Am J Med 1993;
95(5):546–548.
100. Leckie RG, Buckner AB, Bornemann M. Seat belt-related thyroiditis documented
with thyroid Tc-99 m pertechnetate scans. Clin Nucl Med 1992; 17(11):859–860.
101. Stanbury JB, Ermans AE, Bourdoux P, et al. Iodine-induced hyperthyroidism: Occur-
rence and epidemiology. Thyroid 1998; 8(1):83–100.
102. Fradkin JE, Wolff J. Iodide-induced thyrotoxicosis. Medicine (Baltimore) 1983;
62(1):1–20.
103. Livadas DP, Koutras DA, Souvatzoglou A, et al. The toxic effect of small iodine
supplements in patients with autonomous thyroid nodules. Clin Endocrinol (Oxf)
1977; 7(2):121–127.
104. Leger AF, Massin JP, Laurent MF, et al. Iodine-induced thyrotoxicosis: Analysis of
eighty-five consecutive cases. Eur J Clin Invest 1984; 14(6):449–455.
105. Shilo S, Hirsch HJ. Iodine-induced hyperthyroidism in a patient with a normal
thyroid gland. Postgrad Med J 1986; 62(729):661–662.
106. Becker CB, Gordon JM. Iodinated glycerol and thyroid dysfunction. Four cases and
a review of the literature. Chest 1993; 103(1):188–192.
107. Huseby JS, Bennett SW, Hagensee ME. Hyperthyroidism induced by iodinated
glycerol. Am Rev Respir Dis 1991; 144(6):1403.
108. Bryant WP, Zimmerman D. Iodine-induced hyperthyroidism in a newborn. Pediatrics
1995; 95(3):434–436.
109. Jacobson JM, Hankins GV, Murray JM, et al. Self-limited hyperthyroidism following
intravaginal iodine administration. Am J Obstet Gynecol 1981; 140(4):472–473.
110. Martin FI, Tress BW, Colman PG, et al. Iodine-induced hyperthyroidism due to
nonionic contrast radiography in the elderly. Am J Med 1993; 95(1):78–82.
Thyroiditis and Other Forms of Hyperthyroidism 137

111. Vagenakis AG, Wang CA, Burger A, et al. Iodide-induced thyrotoxicosis in Boston.
N Engl J Med 1972; 287(11):523–527.
112. Rao RH, McCready VR, Spathis GS. Iodine kinetic studies during amiodarone
treatment. J Clin Endocrinol Metab 1986; 62(3):563–568.
113. Harjai KJ, Licata AA. Effects of amiodarone on thyroid function. Ann Intern Med
1997; 126(1):63–73.
114. Martino E, Safran M, Aghini-Lombardi F, et al. Environmental iodine intake and
thyroid dysfunction during chronic amiodarone therapy. Ann Intern Med 1984;
101(1):28–34.
115. Nademanee K, Singh BN, Callahan B, et al. Amiodarone, thyroid hormone indexes,
and altered thyroid function: Long-term serial effects in patients with cardiac arrhyth-
mias. Am J Cardiol 1986; 58(10):981–986.
116. Harjai KJ, Licata AA. Amiodarone induced hyperthyroidism: A case series and brief
review of literature. Pacing Clin Electrophysiol 1996; 19(11 Pt 1):1548–1554.
117. Trip MD, Wiersinga W, Plomp TA. Incidence, predictability, and pathogene-
sis of amiodarone-induced thyrotoxicosis and hypothyroidism. Am J Med 1991;
91(5):507–511.
118. Batcher EL, Tang XC, Singh BN, et al. Thyroid function abnormalities during
amiodarone therapy for persistent atrial fibrillation. Am J Med 2007; 120(10):880–
885.
119. Bartalena L, Grasso L, Brogioni S, et al. Serum interleukin-6 in amiodarone-induced
thyrotoxicosis. J Clin Endocrinol Metab 1994; 78(2):423–427.
120. Seminara S, Daniels G. Amiodarone and the thyroid. Endocr Pract 1998; 4:48–
57.
121. Chiovato L, Martino E, Tonacchera M, et al. Studies on the in vitro cytotoxic effect
of amiodarone. Endocrinology 1994; 134(5):2277–2282.
122. Roti E, Minelli R, Gardini E, et al. Thyrotoxicosis followed by hypothyroidism in
patients treated with amiodarone. A possible consequence of a destructive process
in the thyroid. Arch Intern Med 1993; 153(7):886–892.
123. Bartalena L, Brogioni S, Grasso L, et al. Treatment of amiodarone-induced thyrotox-
icosis, a difficult challenge: Results of a prospective study. J Clin Endocrinol Metab
1996; 81(8):2930–2933.
124. Martino E, Aghini-Lombardi F, Lippi F, et al. Twenty-four hour radioactive iodine
uptake in 35 patients with amiodarone associated thyrotoxicosis. J Nucl Med 1985;
26(12):1402–1407.
125. Martino E, Bartalena L, Mariotti S, et al. Radioactive iodine thyroid uptake in patients
with amiodarone-iodine-induced thyroid dysfunction. Acta Endocrinol (Copenh)
1988; 119(2):167–173.
126. Bauters C, D’Herbomez M, Nocaudie M, et al. Interleukin-6 and thyroid radioactive
iodine uptake in patients with amiodarone-induced thyrotoxicosis. J Endocrinol
Invest 1999; 22:94.
127. Bogazzi F, Bartalena L, Brogioni S, et al. Color flow doppler sonography rapidly
differentiates type I and type II amiodarone-induced thyrotoxicosis. Thyroid 1997;
7(4):541–545.
128. Loy M, Perra E, Melis A, et al. Color-flow Doppler sonography in the differen-
tial diagnosis and management of amiodarone-induced thyrotoxicosis. Acta Radiol
2007; 48(6):628–634.
138 Meek and Smallridge

129. Jonckheer MH, Blockx P, Broeckaert I, et al. ‘Low T3 syndrome’ in patients chron-
ically treated with an iodine-containing drug, amiodarone. Clin Endocrinol (Oxf)
1978; 9(1):27–35.
130. Singh BN, Nademanee K. Amiodarone and thyroid function: Clinical implications
during antiarrhythmic therapy. Am Heart J 1983; 106(4 Pt 2):857–869.
131. Bogazzi F, Bartalena L, Tomisti L, et al. Glucocorticoid response in amiodarone-
induced thyrotoxicosis resulting from destructive thyroiditis is predicted by thyroid
volume and serum free thyroid hormone concentrations. J Clin Endocrinol Metab
2007; 92(2):556–562.
132. Dickstein G, Shechner C, Adawi F, et al. Lithium treatment in amiodarone-induced
thyrotoxicosis. Am J Med 1997; 102(5):454–458.
133. Chopra IJ, Baber K. Use of oral cholecystographic agents in the treatment of
amiodarone-induced hyperthyroidism J Clin Endocrinol Metab 2001; 86(10):4707–
4710.
134. Brennan MD, van Heerden JA, Carney JA. Amiodarone-associated thyrotoxico-
sis (AAT): Experience with surgical management. Surgery 1987; 102(6):1062–
1067.
135. Farwell AP, Abend SL, Huang SK, et al. Thyroidectomy for amiodarone-induced
thyrotoxicosis. JAMA 1990; 263(11):1526–1528.
136. Houghton SG, Farley DR, Brennan MD, et al. Surgical management of amiodarone-
associated thyrotoxicosis: Mayo Clinic experience. World J Surg 2004; 28(11):1083–
1087.
137. Bogazzi F, Dell’Unto E, Tanda ML, et al. Long-term outcome of thyroid func-
tion after amiodarone-induced thyrotoxicosis, as compared to subacute thyroiditis.
J Endocrinol Invest 2006; 29(8):694–699.
138. Basaria S, Cooper DS. Amiodarone and the thyroid. Am J Med 2005; 118(7):706–
714.
139. Koh LK, Greenspan FS, Yeo PP. Interferon-alpha induced thyroid dysfunction:
Three clinical presentations and a review of the literature. Thyroid 1997; 7(6):891–
896.
140. Caraccio N, Dardano A, Manfredonia F, et al. Long-term follow-up of 106 mul-
tiple sclerosis patients undergoing interferon-beta 1 a or 1b therapy: Predictive
factors of thyroid disease development and duration. J Clin Endocrinol Metab 2005;
90(7):4133–4137.
141. Lisker-Melman M, Di Bisceglie AM, Usala SJ, et al. Development of thyroid disease
during therapy of chronic viral hepatitis with interferon alfa. Gastroenterology 1992;
102(6):2155–2160.
142. Preziati D, La Rosa L, Covini G, et al. Autoimmunity and thyroid function in patients
with chronic active hepatitis treated with recombinant interferon alpha-2 a. Eur J
Endocrinol 1995; 132(5):587–593.
143. Hsieh MC, Yu ML, Chuang WL, et al. Virologic factors related to interferon-alpha-
induced thyroid dysfunction in patients with chronic hepatitis C. Eur J Endocrinol
2000; 142(5):431–437.
144. Wong V, Fu AX, George J, et al. Thyrotoxicosis induced by alpha-interferon therapy
in chronic viral hepatitis. Clin Endocrinol (Oxf) 2002; 56(6):793–798.
145. Prummel MF, Laurberg P. Interferon-alpha and autoimmune thyroid disease. Thyroid
2003; 13(6):547–551.
Thyroiditis and Other Forms of Hyperthyroidism 139

146. Kakizaki S, Takagi H, Murakami M, et al. HLA antigens in patients with interferon-
alpha-induced autoimmune thyroid disorders in chronic hepatitis C. J Hepatol 1999;
30(5):794–800.
147. Vialettes B, Guillerand MA, Viens P, et al. Incidence rate and risk factors for thyroid
dysfunction during recombinant interleukin-2 therapy in advanced malignancies.
Acta Endocrinol (Copenh) 1993; 129(1):31–38.
148. Mandac JC, Chaudhry S, Sherman KE, et al. The clinical and physiological spectrum
of interferon-alpha induced thyroiditis: Toward a new classification. Hepatology
2006; 43(4):661–672.
149. Tran A, Quaranta JF, Benzaken S, et al. High prevalence of thyroid autoantibodies
in a prospective series of patients with chronic hepatitis C before interferon therapy.
Hepatology 1993; 18(2):253–257.
150. Carella C, Mazziotti G, Amato G, et al. Clinical review 169: Interferon-alpha-related
thyroid disease: Pathophysiological, epidemiological, and clinical aspects. J Clin
Endocrinol Metab 2004; 89(8):3656–3661.
151. Minelli R, Valli MA, Di Secli C, et al. Is steroid therapy needed in the treatment of
destructive thyrotoxicosis induced by alpha-interferon in chronic hepatitis C? Horm
Res 2005; 63(4):194–199.
152. Baudin E, Marcellin P, Pouteau M, et al. Reversibility of thyroid dysfunction induced
by recombinant alpha interferon in chronic hepatitis C. Clin Endocrinol (Oxf) 1993;
39(6):657–661.
153. Barclay ML, Brownlie BE, Turner JG, et al. Lithium associated thyrotoxicosis: A
report of 14 cases, with statistical analysis of incidence. Clin Endocrinol (Oxf) 1994;
40(6):759–764.
154. McDermott MT, Burman KD, Hofeldt FD, et al. Lithium-associated thyrotoxicosis.
Am J Med 1986; 80(6):1245–1248.
155. Miller KK, Daniels GH. Association between lithium use and thyrotoxicosis caused
by silent thyroiditis. Clin Endocrinol (Oxf) 2001; 55(4):501–508.
156. Bogazzi F, Bartalena L, Scarcello G, et al. The age of patients with thyrotoxi-
cosis factitia in Italy from 1973 to 1996. J Endocrinol Invest 1999; 22(2):128–
133.
157. Cohen JD, Utiger RD. Metastatic choriocarcinoma associated with hyperthyroidism.
J Clin Endocrinol Metab 1970; 30(4):423–429.
158. Ohye H, Fukata S, Kanoh M, et al. Thyrotoxicosis caused by weight-reducing herbal
medicines. Arch Intern Med 2005; 165(8):831–834.
159. Kinney JS, Hurwitz ES, Fishbein DB, et al. Community outbreak of thyrotoxicosis:
Epidemiology, immunogenetic characteristics, and long-term outcome. Am J Med
1988; 84(1):10–18.
160. Hedberg CW, Fishbein DB, Janssen RS, et al. An outbreak of thyrotoxicosis caused
by the consumption of bovine thyroid gland in ground beef. N Engl J Med 1987;
316(16):993–998.
161. Mariotti S, Martino E, Cupini C, et al. Low serum thyroglobulin as a clue to the
diagnosis of thyrotoxicosis factitia. N Engl J Med 1982; 307(7):410–412.
162. Bouillon R, Verresen L, Staels F, et al. The measurement of fecal thyroxine in the
diagnosis of thyrotoxicosis factitia. Thyroid 1993; 3(2):101–103.
163. Beck-Peccoz P, Brucker-Davis F, Persani L, et al. Thyrotropin-secreting pituitary
tumors. Endocr Rev 1996; 17(6):610–638.
140 Meek and Smallridge

164. Losa M, Giovanelli M, Persani L, et al. Criteria of cure and follow-up of central
hyperthyroidism due to thyrotropin-secreting pituitary adenomas. J Clin Endocrinol
Metab 1996; 81(8):3084–3090.
165. Wilson CB. A decade of pituitary microsurgery. The Herbert Olivecrona lecture. J
Neurosurg 1984; 61(5):814–833.
166. Mindermann T, Wilson CB. Thyrotropin-producing pituitary adenomas. J Neurosurg
1993; 79(4):521–527.
167. Smallridge RC. Thyrotropin-secreting pituitary tumors. Endocrinol Metab Clin
North Am 1987; 16(3):765–792.
168. Socin HV, Chanson P, Delemer B, et al. The changing spectrum of TSH-secreting
pituitary adenomas: Diagnosis and management in 43 patients. Eur J Endocrinol
2003; 148(4):433–442.
169. Cooper DS, Wenig BM. Hyperthyroidism caused by an ectopic TSH-secreting pitu-
itary tumor. Thyroid 1996; 6(4):337–343.
170. Pasquini E, Faustini-Fustini M, Sciarretta V, et al. Ectopic TSH-secreting pituitary
adenoma of the vomerosphenoidal junction. Eur J Endocrinol 2003; 148(2):253–
257.
171. Gesundheit N, Petrick PA, Nissim M, et al. Thyrotropin-secreting pituitary adeno-
mas: Clinical and biochemical heterogeneity. Case reports and follow-up of nine
patients. Ann Intern Med 1989; 111(10):827–835.
172. Beck-Peccoz P, Persani L. Thyrotropinomas. Endocrinol Metab Clin North Am
2008; 37(1):123–134, viii-ix.
173. Weintraub B, Gershengorn M, Kourides E, et al. Inappropriate secretion of thyroid-
stimulating hormone. Ann Intern Med 1981; 95:339–351.
174. Kourides IA, Ridgway EC, Weintraub BD, et al. Thyrotropin-induced hyperthy-
roidism: Use of alpha and beta subunit levels to identify patients with pituitary
tumors. J Clin Endocrinol Metab 1977; 45(3):534–543.
175. Smallridge RC, Smith CE. Hyperthyroidism due to thyrotropin-secreting pitu-
itary tumors. Diagnostic and therapeutic considerations. Arch Intern Med 1983;
143(3):503–507.
176. Beck-Peccoz P, Roncoroni R, Mariotti S, et al. Sex hormone–binding globulin mea-
surement in patients with inappropriate secretion of thyrotropin (IST): Evidence
against selective pituitary thyroid hormone resistance in nonneoplastic IST. J Clin
Endocrinol Metab 1990; 71(1):19–25.
177. Mannavola D, Persani L, Vannucchi G, et al. Different responses to chronic somato-
statin analogues in patients with central hyperthyroidism. Clin Endocrinol (Oxf)
2005; 62(2):176–181.
178. Frank SJ, Gesundheit N, Doppman JL, et al. Preoperative lateralization of pituitary
microadenomas by petrosal sinus sampling: Utility in two patients with non-ACTH-
secreting tumors. Am J Med 1989; 87(6):679–682.
179. Smallridge RC, Czervionke LF, Fellows DW, et al. Corticotropin- and thyrotropin-
secreting pituitary microadenomas: Detection by dynamic magnetic resonance imag-
ing. Mayo Clin Proc 2000; 75(5):521–528.
180. Brucker-Davis F, Oldfield EH, Skarulis MC, et al. Thyrotropin-secreting pituitary
tumors: Diagnostic criteria, thyroid hormone sensitivity, and treatment outcome in
25 patients followed at the National Institutes of Health. J Clin Endocrinol Metab
1999; 84(2):476–486.
Thyroiditis and Other Forms of Hyperthyroidism 141

181. Chanson P, Weintraub BD, Harris AG. Octreotide therapy for thyroid-stimulating
hormone-secreting pituitary adenomas. A follow-up of 52 patients. Ann Intern Med
1993; 119(3):236–240.
182. Comi RJ, Gesundheit N, Murray L, et al. Response of thyrotropin-secreting pituitary
adenomas to a long-acting somatostatin analogue. N Engl J Med 1987; 317(1):12–17.
183. Gancel A, Vuillermet P, Legrand A, et al. Effects of a slow-release formulation of
the new somatostatin analogue lanreotide in TSH-secreting pituitary adenomas. Clin
Endocrinol (Oxf) 1994; 40(3):421–428.
184. Francis TB, Smallridge RC, Kane J, et al. Octreotide changes serum thyrotropin
(TSH) glycoisomer distribution as assessed by lectin chromatography in a TSH
macroadenoma patient. J Clin Endocrinol Metab 1993; 77(1):183–187.
185. Caron P, Gerbeau C, Pradayrol L, et al. Successful pregnancy in an infertile
woman with a thyrotropin-secreting macroadenoma treated with somatostatin analog
(octreotide). J Clin Endocrinol Metab 1996; 81(3):1164–1168.
186. Kienitz T, Quinkler M, Strasburger CJ, et al. Long-term management in five cases of
TSH-secreting pituitary adenomas: A single center study and review of the literature.
Eur J Endocrinol 2007; 157(1):39–46.
187. Dhillon KS, Cohan P, Kelly DF, et al. Treatment of hyperthyroidism associated
with thyrotropin-secreting pituitary adenomas with iopanoic acid. J Clin Endocrinol
Metab 2004; 89(2):2708–2711.
188. Daousi C, Foy PM, MacFarlane IA. Ablative thyroid treatment for thyrotoxicosis
due to thyrotropin-producing pituitary tumours. J Neurol Neurosurg Psychiatry 2007;
78(1):93–95.
189. Refetoff S, DeWind LT, DeGroot LJ. Familial syndrome combining deaf-mutism,
stippled epiphyses, goiter and abnormally high PBI: Possible target organ refractori-
ness to thyroid hormone. J Clin Endocrinol Metab 1967; 27(2):279–294.
190. Gershengorn MC, Weintraub BD. Thyrotropin-induced hyperthyroidism caused by
selective pituitary resistance to thyroid hormone. A new syndrome of “inappropriate
secretion of TSH”. J Clin Invest 1975; 56(3):633–642.
191. Lafranchi SH, Snyder DB, Sesser DE, et al. Follow-up of newborns with elevated
screening T4 concentrations. J Pediatr 2003; 143(3):296–301.
192. Wu SY, Cohen RN, Simsek E, et al. A novel thyroid hormone receptor-beta mutation
that fails to bind nuclear receptor corepressor in a patient as an apparent cause of
severe, predominantly pituitary resistance to thyroid hormone. J Clin Endocrinol
Metab 2006; 91(5):1887–1895.
193. Beck-Peccoz P, Persani L, Calebiro D, et al. Syndromes of hormone resistance in the
hypothalamic-pituitary-thyroid axis. Best Pract Res Clin Endocrinol Metab 2006;
20(4):529–546.
194. Tsunekawa K, Onigata K, Morimura T, et al. Identification and functional analysis
of novel inactivating thyrotropin receptor mutations in patients with thyrotropin
resistance. Thyroid 2006; 16(5):471–479.
195. Refetoff S, Dumitrescu AM. Syndromes of reduced sensitivity to thyroid hormone:
Genetic defects in hormone receptors, cell transporters and deiodination. Best Pract
Res Clin Endocrinol Metab 2007; 21(2):277–305.
196. Mamanasiri S, Yesil S, Dumitrescu AM, et al. Mosaicism of a thyroid hormone
receptor-beta gene mutation in resistance to thyroid hormone. J Clin Endocrinol
Metab 2006; 91(9):3471–3477.
142 Meek and Smallridge

197. Refetoff S, Weiss R, Usala S. The syndromes of resistance to thyroid hormone.


Endocr Rev 1993; 14(3):348–399.
198. Adams M, Matthews C, Collingwood TN, et al. Genetic analysis of 29 kindreds with
generalized and pituitary resistance to thyroid hormone. Identification of thirteen
novel mutations in the thyroid hormone receptor beta gene. J Clin Invest 1994;
94(2):506–515.
199. Mixson AJ, Renault JC, Ransom S, et al. Identification of a novel mutation in the
gene encoding the beta-triiodothyronine receptor in a patient with apparent selective
pituitary resistance to thyroid hormone. Clin Endocrinol (Oxf) 1993; 38(3):227–234.
200. Brucker-Davis F, Skarulis MC, Grace MB, et al. Genetic and clinical features of
42 kindreds with resistance to thyroid hormone. The National Institutes of Health
Prospective Study. Ann Intern Med 1995; 123(8):572–583.
201. Brucker-Davis F, Skarulis MC, Pikus A, et al. Prevalence and mechanisms of hearing
loss in patients with resistance to thyroid hormone. J Clin Endocrinol Metab 1996;
81(8):2768–2772.
202. Beck-Peccoz P, Persani L, Faglia G. Glycoprotein hormone alpha subunit in pituitary
adenomas. Trends Endocrinol Metab 1992; 3:41–45.
203. Gurnell M, Rajanayagam O, Barbar I, et al. Reversible pituitary enlargement in the
syndrome of resistance to thyroid hormone. Thyroid 1998; 8(8):679–682.
204. Rosler A, Litvin Y, Hage C, et al. Familial hyperthyroidism due to inappropriate
thyrotropin secretion successfully treated with triiodothyronine. J Clin Endocrinol
Metab 1982; 54(1):76–82.
205. Hamon P, Bovier-Lapierre M, Robert M, et al. Hyperthyroidism due to selec-
tive pituitary resistance to thyroid hormones in a 15-month-old boy: Efficacy of
D-thyroxine therapy. J Clin Endocrinol Metab 1988; 67(5):1089–1093.
206. Dorey F, Strauch G, Gayno JP. Thyrotoxicosis due to pituitary resistance to thyroid
hormones. Successful control with D thyroxine: A study in three patients. Clin
Endocrinol (Oxf) 1990; 32(2):221–228.
207. Beck-Peccoz P, Piscitelli G, Cattaneo MG, et al. Successful treatment of
hyperthyroidism due to nonneoplastic pituitary TSH hypersecretion with 3,5,3’-
triiodothyroacetic acid (TRIAC). J Endocrinol Invest 1983; 6(3):217–223.
208. Salmela PI, Wide L, Juustila H, et al. Effects of thyroid hormones (T4,T3),
bromocriptine and Triac on inappropriate TSH hypersecretion. Clin Endocrinol (Oxf)
1988; 28(5):497–507.
209. Darendeliler F, Bas F. Successful therapy with 3,5,3’-triiodothyroacetic acid
(TRIAC) in pituitary resistance to thyroid hormone. J Pediatr Endocrinol Metab
1997; 10(5):535–538.
210. Beck-Peccoz P, Mariotti S, Guillausseau PJ, et al. Treatment of hyperthyroidism due
to inappropriate secretion of thyrotropin with the somatostatin analog SMS 201–995.
J Clin Endocrinol Metab 1989; 68(1):208–214.
211. Connell JM, McCruden DC, Davies DL, et al. Bromocriptine for inappropriate
thyrotropin secretion [letter]. Ann Intern Med 1982; 96(2):251–252.
212. Devaney K, Snyder R, Norris HJ, et al. Proliferative and histologically malignant
struma ovarii: A clinicopathologic study of 54 cases. Int J Gynecol Pathol 1993;
12(4):333–343.
213. Dunzendorfer T, deLas Morenas A, Kalir T, et al. Struma ovarii and hyperthyroidism.
Thyroid 1999; 9(5):499–502.
Thyroiditis and Other Forms of Hyperthyroidism 143

214. Kempers RD, Dockerty MB, Hoffman DL, et al. Struma ovarii–ascitic, hyperthyroid,
and asymptomatic syndromes. Ann Intern Med 1970; 72(6):883–893.
215. Gould SF, Lopez RL, Speers WC. Malignant struma ovarii. A case report and
literature review. J Reprod Med 1983; 28(6):415–419.
216. Scully RE. Recent progress in ovarian cancer. Hum Pathol 1970; 1(1):73–98.
217. Roth LM, Talerman A. The enigma of struma ovarii. Pathology 2007; 39(1):139–146.
218. Pardo-Mindan FJ, Vazquez JJ. Malignant struma ovarii. Light and electron micro-
scopic study. Cancer 1983; 51(2):337–343.
219. Tamsen A, Mazur MT. Ovarian strumal carcinoid in association with multiple
endocrine neoplasia, type IIA. Arch Pathol Lab Med 1992; 116(2):200–203.
220. Ciccarelli A, Valdes-Socin H, Parma J, et al. Thyrotoxic adenoma followed by
atypical hyperthyroidism due to struma ovarii: Clinical and genetic studies. Eur J
Endocrinol 2004;150(4):431–437.
221. Brennan MD, Erickson DZ, Carney JA, et al. Nongoitrous (type I) amiodarone-
associated thyrotoxicosis: Evidence of follicular disruption in vitro and in vivo.
Thyroid 1995; 5(3):177–183.
222. Smith FG. Pathology and physiology of struma ovarii. Arch Surg 1946; 53:603–626.
223. Brown WW, Shetty KR, Rosenfeld PS. Hyperthyroidism due to struma ovarii:
Demonstration by radioiodine scan. Acta Endocrinol (Copenh) 1973; 73(2):266–
272.
224. Nodine J, Maldia G. Pseudostruma ovarii. Obstet Gynecol 1961; 17:460–463.
225. Rosenblum NG, LiVolsi VA, Edmonds PR, et al. Malignant struma ovarii. Gynecol
Oncol 1989; 32(2):224–227.
226. Hershman JM. Human chorionic gonadotropin and the thyroid: Hyperemesis gravi-
darum and trophoblastic tumors. Thyroid 1999; 9(7):653–657.
227. Rajatanavin R, Chailurkit LO, Srisupandit S, et al. Trophoblastic hyperthyroidism:
Clinical and biochemical features of five cases. Am J Med 1988; 85(2):237–241.
228. Desai RK, Norman RJ, Jialal I, et al. Spectrum of thyroid function abnormalities in
gestational trophoblastic neoplasia. Clin Endocrinol (Oxf) 1988; 29(6):583–592.
229. Goodwin TM, Hershman JM. Hyperthyroidism due to inappropriate production of
human chorionic gonadotropin. Clin Obstet Gynecol 1997; 40(1):32–44.
230. Fisher PM, Hancock BW. Gestational trophoblastic diseases and their treatment.
Cancer Treat Rev 1997; 23(1):1–16.
231. Karp PJ, Hershman JM, Richmond S, et al. Thyrotoxicosis from molar thyrotropin.
Arch Intern Med 1973; 132(3):432–436.
232. Giralt SA, Dexeus F, Amato R, et al. Hyperthyroidism in men with germ cell tumors
and high levels of beta-human chorionic gonadotropin. Cancer 1992; 69(5):1286–
1290.
233. Schimke RN, Madigan CM, Silver BJ, et al. Choriocarcinoma, thyrotoxicosis, and
the Klinefelter syndrome. Cancer Genet Cytogenet 1983; 9(1):1–7.
234. Kato K, Mostafa MH, Mann K, et al. The human chorionic gonadotropin molecule
from patients with trophoblastic diseases has a high thyrotropic activity but is less
active in the ovary. Gynecol Endocrinol 2004; 18(5):269–277.
235. Higgins HP, Hershman JM, Kenimer JG, et al. The thyrotoxicosis of hydatidiform
mole. Ann Intern Med 1975; 83(3):307–311.
236. Morley JE, Jacobson RJ, Melamed J, et al. Choriocarcinoma as a cause of thyrotox-
icosis. Am J Med 1976; 60(7):1036–1040.
144 Meek and Smallridge

237. Nisula BC, Taliadouros GS. Thyroid function in gestational trophoblastic neopla-
sia: Evidence that the thyrotropic activity of chorionic gonadotropin mediates the
thyrotoxicosis of choriocarcinoma. Am J Obstet Gynecol 1980; 138(1):77–85.
238. Norman RJ, Green-Thompson RW, Jialal I, et al. Hyperthyroidism in gestational
trophoblastic neoplasia. Clin Endocrinol (Oxf) 1981; 15(4):395–401.
239. Galton VA, Inggar SH, Jimenez-Fonseca J, et al. Alterations in thyroid hormone
economy in patients with hydatidiform mole. J Clin Invest 1971; 50(6):1345–
1354.
240. Hershman JM. Physiological and pathological aspects of the effect of human chori-
onic gonadotropin on the thyroid. Best Pract Res Clin Endocrinol Metab 2004;
18(2):249–265.
241. Gross JL, Vasques Moraes I. Thyroid hormone-producing metastases in differenti-
ated thyroid cancer. J Endocrinol Invest 1996; 19(1):21–24.
242. Paul SJ, Sisson JC. Thyrotoxicosis caused by thyroid cancer. Endocrinol Metab Clin
North Am 1990; 19(3):593–612.
243. Ober KP, Cowan RJ, Sevier RE, et al. Thyrotoxicosis caused by functioning
metastatic thyroid carcinoma. A rare and elusive cause of hyperthyroidism with
low radioactive iodine uptake. Clin Nucl Med 1987; 12(5):345–348.
244. Hunt WB, Crispell KR, McKee J. Functioning metastatic carcinoma of the thyroid
producing clinical hyperthyroidism. Am J Med 1960; 28:995–1001.
245. Studer H, Veraguth P, Wyss F. Thyrotoxicosis due to a solitary hepatic metastasis of
thyroid carcinoma. J Clin Endocrinol Metab 1961; 21:1334–1338.
246. Verschueren CP, Rutteman GR, Vos JH, et al. Thyrotrophin receptors in normal
and neoplastic (primary and metastatic) canine thyroid tissue. J Endocrinol 1992;
132(3):461–468.
247. Bowden WD, Jones RE. Thyrotoxicosis associated with distant metastatic follicular
carcinoma of the thyroid. South Med J 1986; 79(4):483–486.
248. Shimaoka K, VanHerle AJ, Dindogru A. Thyrotoxicosis secondary to involvement of
the thyroid with malignant lymphoma. J Clin Endocrinol Metab 1976; 43(1):64–68.
249. Samuels MH, Launder T. Hyperthyroidism due to lymphoma involving the thyroid
gland in a patient with acquired immunodeficiency syndrome: Case report and review
of the literature. Thyroid 1998; 8(8):673–677.
250. Nakashima T, Inoue K, Shiro-ozu A, et al. Predominant T3 synthesis in the metastatic
thyroid carcinoma in a patient with T3-toxicosis. Metabolism 1981; 30(4):327–330.
251. Sung LC, Cavalieri RR. T3 thyrotoxicosis due to metastatic thyroid carcinoma. J
Clin Endocrinol Metab 1973; 36(2):215–217.
252. Cerletty JM, Listwan WJ. Hyperthyroidism due to functioning metastatic thyroid
carcinoma. Precipitation of thyroid storm with therapeutic radioactive iodine. JAMA
1979; 242(3):269–270.
4
Hypothyroidism

Michael T. McDermott and E. Chester Ridgway


University of Colorado Denver, School of Medicine, Aurora, Colorado, U.S.A.

INTRODUCTION
Hypothyroidism is a condition in which the thyroid gland produces amounts
of thyroid hormones that are insufficient to satisfy the requirements of peripheral
tissues. The normal thyroid gland secretes both thyroxine (T4) and triiodothyronine
(T3). T4 is converted to T3 in peripheral tissues by the enzyme 5’ deiodinase (1);
85% of circulating T3 is produced by peripheral conversion from T4, whereas 15%
is directly secreted from the thyroid gland (2,3). Both T4 and T3 are transported
across target cell membranes to enter the cytoplasm and later the nucleus where
they bind to specific thyroid hormone receptors. T3 binds to these receptors with
about 10-fold higher affinity than does T4. The thyroid hormone-receptor complex
then binds with other cofactors to the regulatory regions of thyroid hormone
responsive genes, where it governs the production of various proteins that mediate
thyroid hormone effects (4–7). Receptors for thyroid hormones are located in
multiple tissues throughout the body, being most abundant in pituitary, brain, liver,
kidney, heart, muscle, bone, and skin cells (6,7). The multitude of thyroid hormone
responsive tissues throughout the body underlies the diverse array of clinical
features that may be seen in patients experiencing thyroid hormone deficiency.
The clinical evaluation and management of patients with hypothyroidism is
relatively straightforward in many cases. However, the common symptoms and
signs of hypothyroidism are neither sensitive nor specific and may therefore pose
numerous potential diagnostic pitfalls. Even the most experienced clinicians may
be challenged by the protean and variable clinical features of hypothyroidism.
This chapter will discuss the epidemiology, etiology, clinical manifestations, and

145
146 McDermott and Ridgway

diagnosis of hypothyroidism but will focus primarily on practical issues in the


management of patients with thyroid hormone deficiency.

EPIDEMIOLOGY
Primary hypothyroidism is the most common functional disorder of the thyroid
gland. Overt hypothyroidism, or thyroid failure, is defined biochemically as an
elevated serum thyroid-stimulating hormone (TSH) and a serum free T4 concen-
tration that is below the population reference range. Subclinical hypothyroidism,
or mild thyroid failure, is defined as an elevated serum TSH with a serum free T4
concentration that is still within the population reference range. Hypothyroidism,
both overt and subclinical hypothyroidism, is distinctly more common in women
than in men and increases in frequency with age. Published statistics regarding
prevalence and incidence are variable because existing studies have differed sig-
nificantly in regard to population age range, geographic location, and criteria used
to define the presence and degree of thyroid failure. The annual incidence has
been estimated to be 4 to 5 per 1000 population per year in women and 0.6 to
0.9 per 1000 per year in men (8,9). The prevalence of overt hypothyroidism has
been reported to be approximately 1% to 3% in large population-based screening
studies (9–11).
Subclinical hypothyroidism is much more prevalent having been found in 4%
to 10% of multiple populations (9–19). The Colorado Thyroid Disease Prevalence
Study of 25,862 state residents reported an elevated serum TSH concentration in
9.5% of all subjects and in 8.9% of those who were not already taking thyroid
hormone; nearly 75% of these individuals had serum TSH values between 5 and
10 mU/L and ⬎95% had normal serum total T4 levels (18). The National Health
and Nutrition Examination Survey III (NHANES III) screened 17,353 adults
and children throughout the United States and found elevated serum TSH levels
(⬎4.5 mU/L) in 1.4% to 8.1% of subjects in all age brackets younger than 60 years
of age (19). This study and others have reported significantly higher prevalence
rates in the elderly population, varying from 7% to ⬎17% in subjects older than
60 years of age (10,13–19). A recent reanalysis of the NHANES data suggested,
however, that the 97th percentile for TSH in persons older than 80 years without
antithyroid antibodies extends to 7.5 mU/L, suggesting that the true prevalence of
SCH in the elderly may be considerably lower (20).
Progression from subclinical to overt hypothyroidism has been reported to
occur in 5% to 18% of patients per year (8,12,16,17,21,22). Individuals most
likely to undergo progression are those with higher initial serum TSH levels,
positive antithyroid antibodies, and a prior history of radioiodine or external beam
radiation therapy (13,23,24). On the other hand, individuals with minimal TSH
elevations may remain stable for years without developing overt hypothyroidism
(23–26), and up to 37% may have their TSH values return to within the population
reference range (24,25). Patients with milder TSH elevations are more likely to
Hypothyroidism 147

achieve normal serum TSH levels over a three- to five-year follow-up period
(24–26).

ETIOLOGY
Thyroid hormone deficiency may result from disease of the thyroid gland itself
(primary hypothyroidism), disorders of TSH secretion due to pituitary disease
(secondary hypothyroidism), or abnormalities of thyrotropin-releasing hormone
(TRH) production from the hypothalamus (tertiary hypothyroidism). The latter two
situations are more generally referred to as “central hypothyroidism” (27,28). Pri-
mary hypothyroidism accounts for over 99.5% of all diagnosed cases of hypothy-
roidism whereas central hypothyroidism is responsible for less than 0.5% (28),
although some sources suggest this may be as high as 5% (29). Conditions that
cause primary and central hypothyroidism are listed in Table 1 (30–39). Medica-
tions that have been reported to cause primary hypothyroidism (40) include amio-
darone (41,42), lithium (43–46), interferon-alpha (47–49), interleukin-2 (50), and
the tyrosine kinase inhibitors, sunitinib (51–54) and sorafenib (55). Bexarotene
causes central hypothyroidism by interfering with TSH secretion (56,57).

CLINICAL MANIFESTATIONS
Overt Hypothyroidism
General Features
Thyroid hormone deficiency results in a broad spectrum of symptoms, signs,
and laboratory abnormalities (58–62) that progress gradually over time. Affected
patients often experience fatigue, weakness, weight gain (usually ⬍10 lb), cold
intolerance, dry skin, facial puffiness, hair loss, constipation, arthralgias, myalgias,
decreased libido, menstrual irregularity and menorrhagia, difficulty concentrating,
and depression. Common physical findings include bradycardia, hypertension,
cool skin with a yellowish discoloration, periorbital edema, coarse hair, thinning
of the lateral eyebrow regions, muscle weakness, carpal tunnel syndrome, and
delayed relaxation phase of the deep tendon reflexes. The most frequently reported
features from studies in the 1930s (60) and in the 1990s (61) are shown in Table 2.
Depending upon the underlying etiology, the thyroid gland may be visibly or pal-
pably enlarged, normal, or nonpalpable. Patients whose hypothyroidism occurs
acutely (withdrawal of thyroid hormone therapy prior to a radioiodine scan or
treatment for thyroid cancer) tend to have more severe symptoms but less promi-
nent physical findings than do patients whose hypothyroidism developed over a
more prolonged period of time. Also, elderly patients tend to have fewer symptoms
of hypothyroidism than younger patients (63).
Hypothyroid patients may also exhibit a number of general laboratory abnor-
malities such as anemia (macrocytic, normocytic, or microcytic), hyponatremia,
hypercholesterolemia, elevated liver-associated enzymes, and increased creatine
148 McDermott and Ridgway

Table 1 Etiology of Hypothyroidism


A. Primary hypothyroidism
1. Chronic lymphocytic thyroiditis (Hashimoto’s disease)
2. Thyroidectomy
3. Radioiodine therapy
4. External radiation therapy (⬎2000 R)
5. Infiltrative/infectious diseases
6. Iodine deficiency
7. Genetic disorders
a. TSH receptor gene mutations
b. Iodide symporter gene mutations
c. Defects in thyroid hormone synthesis
8. Disruptive thyroiditis (usually transient)
a. Postpartum thyroiditis
b. Silent (painless) thyroiditis
c. Subacute (granulomatous) thyroiditis
9. Drug-induced hypothyroidism
a. Thionamides
b. Iodine excess
c. Amiodarone
d. Lithium
e. Interferon alpha
f. Tyrosine kinase inhibitors
g. Retinoid X receptor ligands
B. Central hypothyroidism
1. Mass (tumor, aneurysm)
2. Infiltrative/infectious diseases
3. Pituitary/hypothalamic surgery
4. Pituitary/hypothalamic radiation therapy
5. Genetic disorders
a. Pit-1/Prop-1 gene mutations
b. TSH␤ gene mutations
c. TRH receptor gene mutations

kinase (CK). Prolongation of the bleeding time, suggesting platelet dysfunction,


is another frequent finding. Elevation of the serum prolactin level is commonly
present and may occasionally present a diagnostic dilemma (discussed further
later). Pleural and pericardial effusions may be seen on chest X-rays while elec-
trocardiograms are frequently characterized by bradycardia, diffuse low voltage,
and nonspecific ST segment- and T-wave abnormalities.
Symptoms of central hypothyroidism are similar to, but tend to be milder
than, those seen in patients with primary hypothyroidism. Since this condition most
commonly results from neoplastic or inflammatory processes involving the pitu-
itary gland and/or the hypothalamus, central hypothyroidism is often accompanied
by pituitary mass effects such as headaches, visual field defects, ophthalmoplegia,
Hypothyroidism 149

Table 2 Clinical Symptoms in Patients with Overt Hypothyroidism


Symptom % of Cases Symptom % of Cases

A. Early Studies—1930s (57)


Weakness 99 Constipation 61
Dry skin 97 Weight gain 59
Coarse skin 97 Hair loss 57
Lethargy 91 Lip pallor 57
Slow speech 91 Dyspnea 55
Eyelid puffiness 90 Peripheral edema 55
Cold intolerance 89 Hoarseness 52
Decreased sweating 89 Anorexia 45
Cold skin 83 Nervousness 35
Thick tongue 82 Menorrhagia 32
Facial edema 79 Palpitations 31
Coarse hair 76 Deafness 30
Pale skin 67 Precordial pain 25
Memory impairment 66
B. Recent Study—1990s (58)
Dry skin 76 Paresthesias 52
Cold intolerance 64 Cold skin 50
Coarse skin 60 Constipation 48
Periorbital puffiness 60 Slow movements 36
Diminished sweating 54 Hoarseness 34
Weight gain 54 Impaired learning 22

Source: Adapted from Refs. 57, 58.

deficiencies of other anterior pituitary hormones and, less commonly, diabetes


insipidus. These features often overshadow the manifestations of hypothyroidism
(27,28).

Pulmonary Abnormalities
The fatigue and decreased exercise tolerance commonly seen in hypothyroid
patients may be partly related to the associated disorders of pulmonary function.
Prominent among the pathophysiological features of thyroid failure are CO2 reten-
tion, hypoxemia, decreased diffusing capacity of carbon monoxide (DLCO), and
increased alveolar-arterial (A-a) oxygen gradients (64–68). Mechanisms for these
effects include upper airway obstruction from goiter and soft tissue enlargement,
decreased compliance of the chest wall, respiratory muscle weakness (69–71),
increased capillary permeability (72,73), pleural effusions (74,75) and impairment
of both hypoxic and hypercapneic ventilatory drives (76–78). This combination
of aberrations may also predispose to peripheral or central sleep apnea (79–83).
Respiratory muscle weakness and impaired ventilatory drives may also profoundly
150 McDermott and Ridgway

impair the ability of acutely ill hypothyroid patients to be weaned from assisted
ventilation devices (84).
Cardiovascular Abnormalities
Pericardial effusions may occur in up to 50% of patients with overt thyroid failure
(74,85,86); the effusions are usually small and have little clinical significance (87),
although pericardial tamponade can rarely occur (88). Myocardial dysfunction,
both systolic and diastolic, is another well-recognized feature of overt hypothy-
roidism (89–97); reversible asymmetric septal hypertrophy has been reported as
well (98). Thyroid hormone deficiency is known to increase systemic vascu-
lar resistance (90,99), which may further impair left ventricular performance.
Although hypothyroidism alone rarely causes congestive heart failure (29,100),
a severe dilated cardiomyopathy has been reported in a profoundly hypothyroid
young man with complete reversal upon institution of thyroid hormone replace-
ment therapy (101).
Hypothyroidism also predisposes to the development of coronary artery dis-
ease; the most commonly implicated mechanisms are the associated lipid disorders
and hypertension (18,102,103). Transient myocardial ischemia, due to regional
myocardial perfusion abnormalities, has been demonstrated in some patients with
severe thyroid failure (104). Hypothyroidism has also been shown to be associ-
ated with increased central arterial stiffness (105) and increased carotid intima-
medial thickness (106), both of which improve significantly after the institution of
levothyroxine replacement therapy (105,106). Serum CK levels are often increased
in hypothyroidism (107–110); although the CK is generally of the MM (skeletal
muscle) fraction, there may also be a component of the MB (myocardial) fraction
in some patients (111). Elevated CK levels may occasionally lead to a mistaken
diagnosis of acute myocardial infarction in hypothyroid patients experiencing
atypical chest pain (112).
Endocrine Abnormalities
Hypothyroidism can cause hyperprolactinemia (113,114). Patients with long-
standing hypothyroidism may sometimes develop marked prolactin elevations
associated with pituitary enlargement (pseudotumors) due to thyrotrope hyperpla-
sia (115–117). Reported patients have presented with high serum prolactin levels
and significant pituitary enlargement on imaging studies, suggesting the presence
of pituitary prolactinomas. Upon discovery of elevated serum TSH levels and
institution of thyroid hormone replacement therapy, the elevated prolactin levels
and pituitary enlargement have completely resolved. Hypothyroidism may, on
occasion, be associated with adrenal insufficiency. This most commonly occurs
in patients with type 2 autoimmune polyendocrine syndrome (APS 2; Schmidt’s
syndrome), humen leucocyte antigen (HLA)-DR3/DR4 related condition in
which circulating organ-specific autoantibodies cause thyroid failure; adrenal
failure; and, less often, type 1 diabetes mellitus (118). The two disorders may
also coexist in patients with tumors or infiltrative disorders of the pituitary gland
Hypothyroidism 151

or hypothalamus, resulting in central hypothyroidism and central adrenal insuf-


ficiency (119); such patients often have evidence of pituitary mass effects, other
anterior pituitary hormone deficiencies and, less frequently, diabetes insipidus
(27,28). Since thyroid hormone replacement may acutely lower serum cortisol
levels by increasing the cortisol metabolic clearance rate (120–122), initiating
thyroid hormone replacement without recognizing and treating coexisting adrenal
disease may precipitate an acute adrenal crisis (119). Adrenal function should
be assessed in any hypothyroid patient with a positive family history of adrenal
insufficiency or Schmidt’s syndrome, hyperkalemia, visual field defects, a known
pituitary mass, other anterior pituitary hormone disorders, or diabetes insipidus.
Children with hypothyroidism often have short stature, exhibiting an abrupt
departure from the normal growth curve corresponding to the onset of thyroid
failure; deficiencies of growth hormone secretion and action have been described
in these patients (123–126). Both delayed puberty (127,128) and precocious pseu-
dopuberty (129,130) have been described in hypothyroid children; the mechanism
for the latter is uncertain but may result from a type of hormonal cross-talk involv-
ing the effects of increased TRH and/or TSH on the gonadal axis (130). In adult
women, hypothyroidism may cause infertility, anovulation, irregular or heavy
menses, amenorrhea, and galactorrhea (128), while in men it may cause infertility,
defective spermatogenesis, and erectile dysfunction (131–133).

Psychiatric Disorders
Depression and other psychiatric disturbances in patients with hypothyroidism
were initially reported in the late 19th century (134–137). The term “myxedema
madness,” first used in 1949, referred to the relatively frequent discovery of
hypothyroidism in patients who were residents of mental hospitals (138). Neu-
ropsychiatric alterations that have since been described in association with
hypothyroidism span a wide spectrum of conditions, including irritability, poor
concentration, impaired memory, cognitive dysfunction, depression, paranoia, hal-
lucinations, and schizophrenia (139–141).

Neurological Disorders
Cretinism with mental retardation is a well-known consequence of endemic iodine
deficiency and of untreated congenital hypothyroidism, emphasizing the critical
role thyroid hormone plays in brain development (142). It has been well demon-
strated that euthyroid children whose mothers had untreated or inadequately treated
hypothyroidism during pregnancy have an increased likelihood of neuropsycho-
logical or cognitive impairment (143–149). Hypothyroid adults may, as discussed
earlier, exhibit a variety of psychiatric disorders (139–141). Peripheral metabolic
polyneuropathies and multiple entrapment neuropathies, including but not limited
to carpal tunnel syndrome, are also well described (150–153). Rarely, hypothy-
roidism has been reported to cause a central focal neurological disorder such as
cerebellar ataxia (154).
152 McDermott and Ridgway

Musculoskeletal Disorders
Arthralgias, myalgias and variable degrees of proximal myopathy are well-
recognized features of thyroid hormone deficiency (107,109,155–158). Mod-
erate to marked elevations of serum muscle enzymes are also well described
(107–112) and acute exertional rhabdomyolysis has even been reported (159).
Less commonly, a peculiar myopathy characterized by muscle hypertrophy, stiff-
ness, weakness, and slowness of movement have been described in association
with hypothyroidism; this has been referred to as Hoffman’s syndrome in adults
(160,161) and as the Kocher-Debre-Semelaigne syndrome in children (162).
Infections
Hypothyroid patients have an increased propensity to develop infections, par-
ticularly those of the upper and lower respiratory tract, urinary tract, and skin.
Symptoms and signs related to these infections may occasionally be the initial
manifestations of hypothyroidism. This increased susceptibility to infections is
not well understood but likely involves hypothyroid-related alterations of respira-
tory and bladder function, reduced cutaneous blood flow, and decreased activity
of circulating and tissue phagocytic cells (163–165).

Subclinical Hypothyroidism
Subclinical hypothyroidism (SCH) (Table 3) is often asymptomatic, but up to
30% of affected patients may experience nonspecific physical and psychiatric
symptoms (18,166–177). In the Colorado Thyroid Disease Prevalence Study (18),
patients with only mild TSH elevations had slightly but statistically significantly
more symptoms on a validated thyroid health questionnaire than did euthyroid
controls (Table 4). Symptomatic improvement following institution of thyroid
hormone replacement has been reported in most published studies (166–170,175),
depending largely on the baseline TSH of the affected subjects, with more symp-
tomatic benefit being demonstrated in subjects with TSH levels ⬎ 10 to 12 mU/L
(175–179).
Subtle disorders of myocardial function have been well described in patients
with SCH (166,167,169,177,180–192), but their clinical significance is uncer-
tain (177). Reported abnormalities include impaired myocardial contractility
(166,167,169,180–185,187,191) and diastolic dysfunction (185–187,190), at rest
(167,180,181,184–187) or with exercise (169,182–187). Abnormal myocardial

Table 3 Subclinical Hypothyroidism/


Mild Thyroid Failure—Definition

Few or no clinical signs


Normal free T4 and free T3
Elevated basal TSH
Hypothyroidism 153

Table 4 Clinical Symptoms in Patients with Subclinical Hypothyroidism


Percent having symptoms

Subclinical hypothyroidism Euthyroid controls


Symptom (n = 1799) (n = 22,842) p

Dry skin 28 25 ⬍.001


Poor memory 24 20 ⬍.001
Slow thinking 22 18 ⬍.001
Muscle weakness 22 18 ⬍.001
Fatigue 18 16 ⬍.01
Muscle cramps 17 15 ⬍.001
Cold intolerance 15 12 ⬍.001
Puffy eyes 12 10 ⬍.05
Constipation 8 7 ⬍.05
Hoarseness 7 5 ⬍.05

Source: Adapted from Ref. 17.

texture has been demonstrated in SCH subjects by videodensitometric analysis


(187). SCH has also been reported to increase the risk of developing congestive
heart failure (193) (Fig. 1). Right ventricular systolic and diastolic performance
has also been reported to be impaired (192). The reported effects of thyroid
hormone replacement on cardiac function in SCH patients include enhanced car-
diac contractility (166,169,181–185,187,191), improvement of diastolic function
(186,187,190), normalization of videodensitometric myocardial texture (187), and
improved right ventricular performance (192). Increases in pulmonary vital capac-
ity, the anaerobic threshold and oxygen uptake at the anaerobic threshold have also
been demonstrated (169).
Considerable evidence implicates SCH as a risk factor for atheroscle-
rotic cardiovascular disease. Patients with SCH have well documented mild
elevations of serum total cholesterol, low density lipoprotein (LDL) cholesterol
(18,170,175,194–204) and non-high density lipoprotein (HDL) cholesterol
concentrations (203), and abnormal lipoprotein remnant metabolism (204).
However, one study of patients with short-term overt hypothyroidism reported
that while total cholesterol and LDL cholesterol were increased, the profile was
not atherogenic since the increases were predominantly in large LDL particles
with small, dense LDL particle concentrations remaining unchanged (205);
whether or not this is also true in chronic subclinical hypothyroidism remains to
be determined at this time. Some reports suggest that even high normal serum
TSH values may adversely affect serum lipid and lipoprotein levels (206–208). It
has been estimated that an increase in the serum TSH level of 1 mU/L is associated
with a rise in the serum total cholesterol concentration of 3.5 mg/dL in women
and 6.2 mg/dL in men (207). The condition is also reported to be associated with
insulin resistance and features of the metabolic syndrome (abdominal obesity,
154 McDermott and Ridgway

Figure 1 Congestive heart failure (CHF) cumulative events in older subjects in relation
to serum TSH levels. Higher TSH levels were associated with a higher rate of CHF events
(p = 0.03 for trend). CHF events were higher in subjects with TSH levels ≥7.0 mU/L
compared with euthyroid subjects (p = 0.006); CHF events were not higher in subjects
with TSH levels between 4.5 and 6.9 mU/L. Source: From Ref. 193. Copyright  c 2005,
American Medical Association. All Rights reserved.

hypertension, elevated triglycerides, and low HDL cholesterol) (209–215). One


large cross-sectional study in Australia, however, did not find an association
between SCH and hypertension (216). A positive association between serum TSH
levels and body mass index and obesity, even within the population reference
range for TSH, has also been reported (217,218). Levothyroxine replacement in
SCH patients with TSH levels ⬎10 to 12 mU/L results in statistically significant
reductions in LDL cholesterol (170,175,197–200,202) and non-HDL cholesterol
(203); several studies in patients with milder TSH elevations (5–12 mU/L) have
not been conclusive on this issue (177–179,219), but two carefully designed,
randomized, controlled studies have reported a similar beneficial reduction
in serum total cholesterol and LDL cholesterol (175,200). Two quantitative
literature reviews (197,198) of the prospective studies examining this issue have
concluded that levothyroxine treatment of patients with SCH lowers both serum
total cholesterol and LDL cholesterol by approximately 10 mg/dL.
Surrogate markers of cardiovascular disease have been reported to be
abnormal in patients with SCH. These include increased pulse wave velocity
(220), impaired endothelial function (175,221,222), increased levels of C-reactive
Hypothyroidism 155

Figure 2 Carotid intimal medial thickness (IMT) values in individual patients with sub-
clinical hypothyroidism at baseline and on treatment with LT4. (A) Patients whose mean-
IMT decreased (n = 19). (B) Patients whose mean-IMT did not change (n = 3) or increased
(n = 1). Source: From Ref. 200.

protein (210), increased carotid artery intima-media thickness (CIMT) (200)


(Fig. 2), and increased arterial stiffness (223). These abnormalities have also been
shown to improve with levothyroxine replacement therapy (175,200,220–223).
While some studies have demonstrated an association of SCH with the
development of atherosclerosis (169,224–229), not all studies are in agreement on
this issue (193,230,231,232) (Fig. 3). The Rotterdam Study reported that patients
with SCH have a significantly increased prevalence of both aortic atherosclerosis
156 McDermott and Ridgway

Figure 3 Meta-analysis of coronary heart disease (CHD) risk in subclinical hypothy-


roidism (SH). Odds ratio (OR; diamonds) and 95% confidence intervals (CI; horizontal
lines) for CHD in subjects with SH are shown. Abbreviations: CC, case-control study; CS,
cross-sectional study; PC, prospective cohort study. Source: From Ref. 228.

and myocardial infarctions and that, after controlling for other known cardiovas-
cular risk factors, SCH was found to be an independent and equally important
risk factor for myocardial infarction (224); however, no increase in myocardial
infarctions was observed in SCH over a 4.6-year follow-up period. A Japanese
study similarly reported that SCH was associated with ischemic heart disease and
with all-cause mortality in men only (226) (Fig. 4). The Busselton Health Study
also concluded that SCH was a significant and independent risk factor for coronary
heart disease (227). In contrast, other large population-based studies, including
the Health, Aging, and Body Composition Study (193), the New Mexico Elder
Health Study (230), and the Cardiovascular Health Study (231), did not show an
association between SCH and cardiovascular disorders or mortality (231). Another
study of subjects over age 85 actually reported enhanced four-year survival in a
cohort of very elderly subjects with untreated SCH (232). At this time, there is
little data concerning the effects of treating SCH on cardiovascular outcomes. One
small uncontrolled, retrospective analysis (233) showed progression of coronary
atherosclerosis in subjects with elevated serum TSH levels on levothyroxine ther-
apy compared to those with normal TSH levels (p ⬍ 0.02). An observational study
Hypothyroidism 157

Figure 4 Survival rates (Kaplan-Meier survival curves) were lower in men (A) with
subclinical hypothyroidism than in the controls, but this difference was not seen in women
(B). Source: From Ref. 226.

suggested that patients treated for hypothyroidism might have an increased risk of
cardiovascular events, although the authors attributed this finding to underlying
atherosclerosis from preexisting hypothyroidism, inappropriate thyroid hormone
dosing, or both (234). Randomized controlled trials to determine if treatment of
SCH improves cardiovascular outcomes are clearly needed (177,178).
158 McDermott and Ridgway

Cross-sectional studies have demonstrated evidence of specific neurobehav-


ioral and neuromuscular symptoms in mild thyroid failure patients (18,174,235–
244). Depression (235–239), memory loss (18,174,235,240,241), cognitive
impairment (174,241,242), and a variety of neuromuscular complaints (243,244)
have all been reported to occur more frequently in patients with this condition.
Other studies, however, have found no evidence of depression (245), cognitive
loss (245), or neuropsychological dysfunction (246) in SCH patients. Objective
evidence of neurological dysfunction, including decreased peripheral nerve con-
duction amplitude (247), an abnormal stapedial reflex (248), abnormal cerebral
nerve latency (249), and alterations of cerebral blood flow (250) have been demon-
strated in these patients. Skeletal muscle abnormalities, including elevated serum
creatine phosphokinase levels (110), increased circulating lactate levels during
exercise (243) and repetitive discharges on surface electromyography (244) have
also been reported. Memory has been shown to improve significantly in one ran-
domized controlled trial (RCT) (240) and in two uncontrolled studies in which
mild thyroid failure patients were given levothyroxine therapy (235,241). Other
reported benefits from uncontrolled interventional studies include reduction in
neuromuscular complaints (235,244) and normalization of abnormal electromyo-
grams (244).

Myxedema Coma
Myxedema coma is a life-threatening condition that represents the extreme end of
the spectrum of thyroid hormone deficiency (251–254). It usually occurs in elderly
patients, with inadequately treated or untreated hypothyroidism, who then have a
superimposed precipitating event. Conditions reported to precipitate myxedema
coma include prolonged cold exposure, infection, trauma, surgery, myocardial
infarction, congestive heart failure, pulmonary embolism, stroke, respiratory fail-
ure, gastrointestinal bleeding, and the use of medications, particularly central
nervous system depressants. Affected patients generally present with hypother-
mia, bradycardia, hypotension, and hypoventilation along with central nervous
system manifestations such as seizures, stupor, and coma. Deep tendon reflexes
are absent or exhibit a severely delayed relaxation phase. Typical myxedematous
changes of the skin are usually apparent. Pleural, pericardial, and peritoneal effu-
sions are frequent findings. An ileus is common and acute urinary retention may
also occur. Treatment, discussed later, must be instituted promptly because of the
very high mortality rate in this condition when appropriate therapy is delayed or
neglected (255,256).

DIAGNOSIS
Hormone Assays
The key to the accurate diagnosis of hypothyroidism is measurement and appropri-
ate interpretation of serum thyrotropin (TSH) and thyroid hormone levels. Thyroid
Hypothyroidism 159

Figure 5 Primary hypothyroidism development. In the earliest stage, mild hypothy-


roidism, the only detectable abnormality is an elevated serum TSH level. Progression
to moderate thyroid failure is characterized by a higher serum TSH concentration and a low
serum T4 value; the serum T3 level is relatively preserved by enhanced T4 to T3 conversion.
In severe hypothyroidism, the serum TSH is higher still, the T4 is further reduced and the
serum T3 declines below the normal range.

gland failure is a gradual but generally progressive process. Although it exists on


a continuum, it is instructive to characterize the disorder according to grades of
severity as being mild, moderate, or severe (Fig. 5, Table 5). When the thyroid
gland first begins to fail, its production and secretion of T4 and T3 diminish. The
slight drop in serum thyroid hormone levels is detected by the hypothalamus and
pituitary gland, which respond with an increase in pituitary TSH secretion. TSH
then stimulates the secretory activity of the damaged thyroid gland, achieving
serum T4 and T3 levels within the population reference range but still low for that
individual patient. The only detectable abnormality at this early stage, therefore, is
a mildly or moderately elevated serum TSH concentration, usually ⬍10 mU/L. As
thyroid hormone synthetic capacity worsens, T4 levels begin to decline but serum
T3 is preferentially maintained because of increased T4 to T3 conversion due to
enhanced 5’ deiodinase activity in both the thyroid gland and peripheral tissues
(257–259). Moderate hypothyroidism is thus characterized by a high serum TSH
level and low T4 but a normal T3 concentration. Once thyroid function becomes
severely impaired, insufficient T4 is generated to sustain adequate T3 production
and the T3 level also declines. Accordingly, severe hypothyroidism is character-
ized by a very high serum TSH level, very low serum T4 and free T4 and a low
serum T3 concentration.
160 McDermott and Ridgway

Table 5 TSH and Thyroid Hormone Levels in Primary Hypothyroidism, Central


Hypothyroidism, Euthyroid Sick Syndrome and Thyroid Hormone Resistance Syndromes

TSH T4 Free T4 T3 Free T3 T3 RU

Primary hypothyroidism
Mild ↑ N N N N N
Moderate ↑↑ ↓ ↓ N N N
Severe ↑↑↑ ↓↓ ↓↓ ↓ ↓ ↓
Central hypothyroidism
Mild N N N N N N
Moderate ↓, N ↓ ↓ N N N
Severe ↓, N ↓↓ ↓↓ ↓ ↓ ↓
Euthyroid sick syndrome
Mild–Moderate N N N ↓ ↓ N
Moderate–Severe N, ↓ ↓ N ↓↓ ↓↓ ↑
Recovery N, ↑ N N, ↓ N N N
Thyroid hormone resistance
N, ↑ ↑ ↑ ↑ ↑ N, ↑

When thyroid gland failure is secondary to inadequate pituitary TSH secre-


tion (central hypothyroidism), the serum TSH level fails to rise, making the diag-
nosis of mild central hypothyroidism very difficult to make. Progressively more
severe degrees of central hypothyroidism are characterized by declining serum
T4 and T3 profiles similar to those seen with primary hypothyroidism (Fig. 6,
Table 5) with the exception that the serum TSH level remains normal or low
(27,28,260). In some cases, however, mildly elevated serum TSH concentrations
have been observed; such patients apparently produce a TSH molecule with nor-
mal immunological but reduced biological activity due to altered glycosylation
(261,262).
Hypothyroidism that occurs after antithyroid drug therapy or radioiodine
treatment of thyrotoxicosis is often characterized by a mixed hormone profile.
Serum T4 and T3 levels commonly decrease rapidly after therapy and become
frankly low by six to eight weeks. In approximately 90% of patients, the TSH
levels remain low or undetectable for several weeks, giving a picture similar to
that seen in central hypothyroidism (263). This transient phenomenon is likely
due to delayed recovery of chronically suppressed pituitary thyrotropes caused
by the antecedent thyrotoxicosis. Eventually, the thyrotrope cells do recover and
serum TSH levels increase as the majority of patients progress to develop typical
primary hypothyroidism (263). Approximately 15% of patients, however, may
later recover sufficient thyroid function to return to a euthyroid state (264).
TSH elevations are almost always indicative of some degree of primary
thyroid failure. However, variable elevations of the serum TSH level can sometimes
be seen in other conditions such as the recovery phase of the nonthyroidal illness
Hypothyroidism 161

Figure 6 Central hypothyroidism development. This disorder results from impaired pitu-
itary TSH secretion and thus there is no increase in serum TSH to signal early or mild
dysfunction. Moderate central hypothyroidism is characterized by a low serum T4 concen-
tration, a low or low-normal TSH level and a normal T3 value. In severe central hypothy-
roidism, the serum T4 is lower still, the T3 becomes depressed and the TSH remains in the
low or low-normal range.

syndrome (euthyroid sick syndrome) (265,266), untreated or inadequately treated


adrenal insufficiency (267), and rare genetic mutations in the TSH receptor (268).

Antithyroid Antibodies
Measurement of antithyroid antibodies is useful for determining the etiology of
primary hypothyroidism. The major thyroid antigens known to elicit autoanti-
body formation are thyroglobulin (TG), thyroid peroxidase (TPO), and the TSH
receptor. Lymphocytic thyroiditis (Hashimoto’s disease) is characterized by the
presence of high titers of anti-TG and anti-TPO antibodies. Anti-TG antibodies
are present in approximately 60% of patients with lymphocytic thyroiditis whereas
anti-TPO antibodies are present in about 95%; thus, anti-TPO antibodies are the
best autoantibody marker for this disorder (30). Knowledge of antithyroid antibody
titers may be particularly useful when one is attempting to predict whether patients
with subclinical hypothyroidism will progress to overt hypothyroidism since this
progression is much more likely in patients who have circulating anti-TPO anti-
bodies (13,23,24). Hypothyroidism may occasionally result from the production
of TSH receptor blocking antibodies (269–271) without evidence of anti-TG and
anti-TPO antibodies.
162 McDermott and Ridgway

Screening for Hypothyroidism


Screening for hypothyroidism among asymptomatic persons is an issue that has
sparked significant controversy (11,178,179,272–279). A comprehensive 1996
cost-utility analysis (273) estimated that screening for hypothyroidism at the
routine health examination starting at age 35 years in both sexes was as cost
effective as several other generally accepted medical practices. American College
of Physicians (ACP) publications in 1998 and 2004 (11,274–276), considering
both prevalence and cost-effectiveness, recommended screening women over age
50 years with a serum TSH determination; a free T4 measurement was recom-
mended only if the TSH value is undetectable or ⬎10 mU/L. The Institute of
Medicine recommended against routine screening of the Medicare population,
however, citing insufficient data showing benefit (277). The American Thyroid
Association, in contrast, recommended screening women at age 35 and men at
age 45, and repeated screening in both genders every five years thereafter (278).
A 2004 Consensus Conference, consisting of a multidisciplinary panel, conducted
an evidence-based review based on prevalence, outcomes, and cost-effectiveness
data available up to 2001, and concluded that there was insufficient evidence
to recommend for or against routine population screening for thyroid disease in
adults (178,179); however, they did recommend aggressive case finding in select
high-risk patients. Subsequently, a panel of thyroidologists reviewed the same and
subsequent evidence in 2004 and concluded instead that population screening for
thyroid disease was generally warranted (279).
Many experts also recommend screening in all women who are planning
pregnancy or early in the first trimester of pregnancy, but this is not universally
accepted practice (280). Data favoring screening include studies showing that
untreated or inadequately treated maternal hypothyroidism may have adverse con-
sequences on fetal neuropsychological and cognitive development (143–149).
It has been demonstrated that undertaking a high-risk case finding strategy,
as opposed to general screening, will fail to identify a substantial number of
hypothyroid mothers (281). TSH testing is also indicated for all patients who have
symptoms compatible with thyroid hormone deficiency, palpably enlarged thy-
roid glands, and conditions associated with an increased prevalence of associated
hypothyroidism (Table 6).

TREATMENT
Thyroid Hormone Preparations
Thyroid hormone replacement therapy was introduced in 1891 when George Mur-
ray reported on studies treating myxedema with injections of an extract of sheep
thyroid glands (282). Subsequently, Hector MacKenzie demonstrated similar ben-
eficial effects from an oral preparation of whole sheep thyroid and later from a des-
iccated extract of ovine thyroid (283). Following these reports, desiccated extracts
Hypothyroidism 163

Table 6 Conditions Associated with an Increased Risk of


Hypothyroidism

A. High-risk patients and conditions (prevalence ⬎10%)


1. Chronic lymphocytic thyroiditis
2. Previous treatment for thyrotoxicosis
3. Previous high dose neck radiation therapy
4. Suspected hypopituitarism
5. Amiodarone therapy
B. Moderate risk patients and conditions (prevalence 3–10%)
1. Goiter or thyroid nodular disease
2. Hypercholesterolemia
3. Graves’ ophthalmopathy
4. Postpartum women
5. Lithium carbonate therapy
6. Interferon-alpha therapy
7. Tyrosine kinase inhibitor therapy
8. Associated autoimmune disease
C. Low-risk patients and conditions (prevalence ⬍2%)
1. Adults and children at routine visits
2. Dementia
3. Psychiatric patients
4. Elderly patients
5. Sleep apnea

from sheep, cow, and pig thyroids became the standard treatment for hypothy-
roidism. Synthetic liothyronine (LT3) was introduced in 1956 and levothyroxine
(LT4) became available in 1958; however, thyroid extracts remained the treatment
of choice because of the assumption that the combination of T4 and T3 in these
extracts was more physiological than either component given alone. The discovery
in the early 1970s that T4 is converted to T3 in peripheral tissues and that only
15% to 20% of circulating T3 comes from direct thyroid secretion (2,3) led to a
reconsideration of this premise. Clinical studies subsequently demonstrated the
advantages of using LT4 alone as replacement therapy (284,285) and that serum
TSH measurements were a superior tool for monitoring LT4 dosage requirements
(286,287).
Multiple thyroid hormone preparations are currently marketed for use
as thyroid hormone replacement therapy. Pure LT4 is available in various oral
brand-name [Synthroid R
(Abbott), LevoxylR
(Jones Pharma/King), LevothroidR

(Lloyd, distributed by Forest), Unithroid (Jerome Stevens)] and generic


R

[Levo-T R
(Alara), Levothyroxine R
(Mylan), Levothyroxine R
(Genpharm),
R 
R
Tirosint (Institute Biochimique), and Levolet (Vintage)] products: there is
also a parenteral formulation (Synthroid R
). Pure LT3 is available as an oral
preparation (Cytomel ) and a parenteral product (Triostat
R R
). Mixtures of LT4

R R
and LT3 (Liotrix , Thyrolar ) and several brands of desiccated thyroid are also
164 McDermott and Ridgway

manufactured; their use is generally not recommended because these products


contain a higher T3/T4 ratio (1:4) (288,289) than is present in human thyroid
secretions (approximately 1:10–20). Due to the high T3 content and the rapid,
essentially complete absorption of T3 from the intestinal tract, serum T3 levels
rise significantly, often into the supraphysiological range, two to six hours after
ingestion of these medications (290).
Most thyroid specialists today consider LT4 to be the treatment of choice for
hypothyroidism (285,291). This is based on the principal that T4 is converted to T3
in peripheral tissues at a regulated rate that is appropriate for the overall metabolic
needs of the body. Physiological replacement of LT4 relieves symptoms and
normalizes serum TSH in the vast majority of hypothyroid patients. Furthermore,
LT4 therapy given to thyroidectomized individuals in doses sufficient to normalize
serum TSH levels has been demonstrated to produce serum T3 levels that are
similar to those in the euthyroid, prethyroidectomy state (292). LT4 is available
in multiple dose sizes, allowing precise titration of the dosage in each individual
patient according to their symptoms and serum TSH levels. Furthermore, its
relatively slow intestinal absorption and long serum half-life provide stable serum
concentrations with minimal diurnal variation. Side effects and toxicity occur
only with overtreatment, a condition that can be avoided by choosing a treatment
schedule that is appropriate for each patient and by regular monitoring of the
serum TSH.
Brand-name LT4 preparations are recommended by many thyroidologists.
Several early studies suggested that different LT4 brands were not equal in potency
(293–296), while a subsequent study called this into question, reporting that four
generic and brand-name preparations were all bioequivalent (297); this sparked a
burst of controversy over study design, interpretation of results and conflicts of
interest (298–300). In 2005, the FDA determined by its own methodology that
many of the brands and generic preparations of levothyroxine were bioequiva-
lent. However, subsequent investigations suggested that the FDA bioequivalence
methodology was imperfect, indicating that patients who were switched between
levothyroxine preparations may need to be retested with serum TSH measure-
ments five to six weeks later to assure adequate control. This would be especially
important for patients whose thyroid hormone levels require precise control, most
notably thyroid cancer patients and hypothyroid women during pregnancy. In
2009, the FDA will take a step toward enhancing the quality of all levothyroxine
products by requiring that they retain from 95% to 105% of their stated potency
throughout their entire shelf life; the previous standard was 90% to 110%. At
present, considering their overall low cost, and our desire to provide our hypothy-
roid patients with the most reliable supply of thyroid hormone, we continue to
recommend the use of brand-name LT4 preparations for thyroid hormone replace-
ment in order to assure that there is consistency every time patients refill their
prescription. If a patient does switch brands or is placed on a generic preparation
of thyroxine, reassessment of the serum TSH is recommended six weeks later,
since the preparations may not yield the same target serum TSH.
Hypothyroidism 165

The absorption of oral LT4 is approximately 80% whereas that of LT3 is


85% to 100% (301–303). Absorption of LT4 occurs at multiple sites throughout
the length of the small intestine but nearly two-thirds occurs in the proximal
small bowel (301,302). It is recommended that LT4 be taken fasting, on an empty
stomach, in the morning since the presence of food decreases LT4 absorption
by about 10% (304). Alternatively, taking LT4 in the evening or at bedtime may
provide better medication absorption than morning dosing (305), although this
study did not control for the fact that study subjects ate after the morning dose
but not after the bedtime dose. Administered LT4 accumulates slowly and has a
serum half-life of about seven days. It requires approximately five to six weeks for
serum levels of T4 to reach a new steady state on a given dose of LT4; therefore,
sampling for serum TSH and/or T4 levels should be performed no sooner than
five to six weeks after therapy is initiated or a dosage change has been made.
Both euthyroid and hypothyroid individuals exhibit a small but significant diurnal
variation in TSH secretion, having slightly higher serum TSH levels between the
hours of 11:00 p.m. and 4:00 a.m. (306). In addition, hypothyroid patients have
small decreases in serum TSH levels (307,308) and increases in serum total and
free T4 concentrations (307) several hours after exogenous LT4 administration.
When possible, TSH measurements should be obtained prior to LT4 ingestion.

Patient-Oriented Approach to Treatment


Thyroid hormone replacement therapy must be individualized for each patient.
The average full replacement dose of LT4 is 1.6 to 1.7 ␮g/kg/day (303,309,310).
Since dose requirements relate more strongly to lean body mass than to total body
mass (311), it may be preferable to estimate a starting dose based on ideal body
weight rather than actual weight in obese individuals. Alternatively, the initial
LT4 dose can be estimated based on the magnitude of the TSH elevation (312).
However, the starting dose, titration schedule, final dose, and ultimate goals of
therapy depend upon multiple factors such as the patient’s age, symptoms and
general state of health as well as the severity, duration, and underlying cause of
hypothyroidism.

Overt Primary Hypothyroidism


The young, otherwise healthy, patient with overt primary hypothyroidism can usu-
ally be started directly on a full LT4 replacement dose of 1.6 to 1.7 ␮g/kg/day
(310,313). After six weeks, the serum TSH level should be measured and the
results used to guide dosage titration until the serum TSH is normal (314)
(Table 7). Since TSH levels in the general population are not normally distributed
over the reported normal range, but are skewed with the majority of individuals
having values at the low end of the range, most practitioners attempt to maintain
serum TSH levels between 0.5 and 2.0 mU/L. Once the TSH concentration is in the
desired range, it should be rechecked in three months and then on an annual basis.
166 McDermott and Ridgway

Table 7 Fine Tuning LT4 Dosages

If serum TSH is Then

⬎ 5.0 mU/L Increase daily LT4 dose by 12.5–25 ␮g/day


0.5–5.0 mU/L Continue dose; recheck annually
⬍0.5 mU/L Decrease daily LT4 dose by 12.5–25 ␮g/day

When a hypothyroid patient is older than 60 years, it is prudent to adopt a


more cautious approach to LT4 replacement therapy. This is because the likelihood
is greater that such patients may harbor subclinical coronary artery disease that
could become symptomatic if they are initially given a full replacement dose.
Therefore, a starting LT4 dose of 50 ␮g/day and a 25 ␮g/day increment six
weeks later is recommended. Once a dose of 75 ␮g/day has been reached, further
adjustments should be guided by the results of TSH testing. As in younger patients,
the goal of treatment is to relieve symptoms and ideally to maintain the serum
TSH between 0.5 and 2.0 mU/L.
Hypothyroid patients with known or suspected coronary artery disease
should be treated even more carefully. In such patients, many practitioners begin
with a low LT4 dose of 12.5 to 25 ␮g/day and increase the dose in 12.5 to 25 ␮g
increments every six to eight weeks. A goal TSH level of 0.5 to 2.0 mU/L is appro-
priate if it can be achieved without precipitation or exacerbation of cardiac symp-
toms. However, for patients who develop chest pain or palpitations at these doses,
a more conservative TSH goal of 2.0 to 5.0 mU/L or slightly higher is preferable.
Subclinical Hypothyroidism
Once a diagnosis of subclinical hypothyroidism is confirmed by retesting, the
decision to be made is whether or not thyroid hormone replacement therapy should
be initiated. This question has inspired much study and has engendered significant
controversy (176–179,279,315,316). The 2004 Consensus Conference evidence-
based review concluded that there was insufficient evidence to recommend for
or against routine treatment of patients with mildly elevated serum TSH levels
of 5 to 10 mU/L but considered the evidence to be fair in support of treating
patients whose serum TSH values are ⬎10 mU/L (178,179). A similar view was
expressed in a meta-analysis of thyroid hormone replacement therapy in SCH,
which noted that such treatment did not result in improved survival, prevent
cardiovascular morbidity, or improve quality of life or symptoms (317). The
2005 thyroid specialist panel report recommended instead that all patients with
elevated serum TSH values be considered for levothyroxine therapy and that
the treating provider should make this determination based on clinical judgment
(279). Because patients with SCH may have a variety of nonspecific symptoms
(18,166–177), such as subtle cardiac dysfunction (166,167,169,177,180–192),
lipid abnormalities (18,170,175,194–204), increases in other cardiovascular risk
Hypothyroidism 167

Table 8 Mild Thyroid Failure—Benefits and Costs of Early


Detection and Therapy

Benefits
1. Treat symptoms related to mild thyroid hormone deficiency
2. Control associated hypercholesterolemia
3. Prevent progression to overt hypothyroidism
Costs
1. Serum TSH assays
2. Levothyroxine therapy
3. Follow-up visits
4. Risk of iatrogenic hyperthyroidism

factors (209–215), and an increased risk of cardiovascular disease (169,224–229),


and because of the tendency for the condition to progress to overt hypothyroidism
(8,12,16,17,21,22), many experts recommend that most, if not all, patients with
mild thyroid failure should be considered for treatment with LT4 in doses sufficient
to reduce the serum TSH levels to 0.5 to 2.0 mU/L (Table 8).
The management of subclinical hypothyroidism in young patients may vary
according to the preference of the provider. One approach is to start LT4 in a dose
of 25 to 50 ␮g/day, recheck the serum TSH in six weeks, and adjust the dose in 12.5
to 25 ␮g increments, if needed, until the first dose is reached that brings the TSH
into the desired range. That dose is then maintained and the patient is monitored
annually. When, and if, the TSH again rises above normal, the LT4 dose is increased
in 12.5 to 25 ␮g increments to the next level that normalizes the serum TSH.
This method is intended to avoid initial overtreatment in patients who may have
significant residual thyroid function and possibly autonomous activity (313,318).
An alternative approach is to start patients directly on an estimated full
LT4 replacement dose of 1.6 to 1.7 ␮g/kg/day. The serum TSH is then checked
six weeks later and the LT4 dose adjusted to bring the TSH level into the 0.5
to 2.0 mU/L range. Once there, the patient should be monitored on an annual
basis. The rationale for this method is that these patients will likely develop
overt hypothyroidism eventually and will require full replacement doses at that
time. By starting them on full replacement at the outset, the time and expense
required for multiple stepwise adjustments in the future are avoided and patients
do not have recurrent periods of subclinical hypothyroidism as their thyroid glands
progressively fail. The risk of this option is that the combination of exogenous
and endogenous thyroid hormone may prove to be excessive, especially if there is
residual autonomous thyroid activity, producing high-normal serum free T4 levels
and suppressed serum TSH concentrations (subclinical thyrotoxicosis).
Subclinical hypothyroidism in patients over age 60 should be treated
cautiously, again because overtreatment can be hazardous at this age. The
recommended approach is to start them on 25 to 50 ␮g/day of LT4 and to increase
the dose, as described earlier for young patients, until the first dose that lowers the
168 McDermott and Ridgway

serum TSH into the desired range of 0.5 to 2.0 mU/L is reached (300,303). In the
oldest old (patients over age 80 years), the data suggest that mild thyroid failure
is associated with decreased mortality (232). Therefore, it has been recommended
that treatment be initiated when the serum TSH is higher (e.g., ⬎10 mU/L), and
that the target serum TSH be higher than it would be for younger patients, for
example, 4 to 6 mU/L. (319).
Mild thyroid failure in patients with known or suspected coronary artery
disease should be treated carefully using the more gradual approach described
earlier for heart patients with overt hypothyroidism. The goal TSH range of 0.5 to
2.0 mU/L is reasonable if the required LT4 dose does not worsen cardiac symptoms.
Otherwise, a more conservative TSH goal of 2.0 to 5.0 mU/L or slightly higher
would be recommended.

Postradioiodine Hypothyroidism
Since an elevated serum TSH level often occurs relatively late in the development
of postradioiodine hypothyroidism (263), LT4 replacement should be initiated
when the patient becomes symptomatic or the serum T4 or free T4 level first
becomes subnormal, even if the TSH is still suppressed. Such early replacement
may reduce the risk of developing subsequent Graves’ orbitopathy (320). Since
most of these patients are young and their hypothyroidism is of short duration,
initiating low dose LT4 with gradual titration upward is rarely necessary. Nonethe-
less, some individuals may have residual autonomous thyroid function and become
symptomatically thyrotoxic on full LT4 replacement doses. Therefore an initial
LT4 replacement dose of 1.2 to 1.3 ␮g/kg/day in these patients may be appropri-
ate, with careful monitoring of their clinical status, free T4 and TSH levels. About
three months after radioiodine treatment, TSH secretory capacity usually recovers,
after which serum TSH levels can be used to monitor therapy as in other patients
with primary hypothyroidism. The goal serum TSH range for these patients should
be 0.5 to 2.0 mU/L.

Central Hypothyroidism
This disorder is significantly more difficult to manage because of the inability to
use serum TSH measurements as an indicator of when to treat and as a guide to the
appropriate LT4 replacement dose (27,28,260). These decisions therefore require
a greater degree of clinical judgment. Treatment should generally be started when
patients develop symptoms consistent with thyroid hormone deficiency and the
serum free T4 level is in the low normal or frankly low range. As with primary
hypothyroidism, LT4 dosing schedules need to be determined with consideration
given to the patient’s age, severity of hypothyroidism, and the presence of under-
lying illnesses. Since TSH levels are usually uninformative and often misleading
in this condition, serum free T4 measurements should be used to monitor the
adequacy of LT4 replacement (260). The goal of treatment should be amelioration
of pertinent symptoms and maintenance of serum free T4 levels in the mid-normal
Hypothyroidism 169

to high-normal range (260,321). Because of the possibility of coexisting central


adrenal insufficiency, adrenal function should be assessed prior to LT4 treatment.
If adrenal function is found to be inadequate, glucocorticoid replacement should
be initiated at the same time LT4 replacement is started.

Thyroid Cancer
Patients with hypothyroidism resulting from thyroidectomy for thyroid cancer
may require as much as 20% more LT4 for replacement and suppression than
do patients whose hypothyroidism has resulted from autoimmune thyroid disease
or radioiodine treatment of Graves’ disease (322). The serum TSH concentration
remains the most accurate guide to the proper LT4 dosage (286,287,323). When
TSH suppression is the goal, serum TSH levels should be maintained below
or at the lower limits of the normal range, whereas for simple replacement the
goal TSH range should be 0.5 to 2.0 mU/L. There is considerable controversy
over the appropriate degree of TSH suppression in patients with thyroid cancer.
Some believe that LT4 doses sufficient to reduce serum TSH levels to between
0.1and 0.5 mU/L are satisfactory for all patients with thyroid cancer (322). Others
believe that full suppression of TSH levels to below the detection limits of third
generation TSH assays (⬍0.01 mU/L) is ideal (323–325). Evidence-based practice
guidelines from the American Thyroid Association recommend that in patients
with low-risk disease who have been shown to be free of disease following surgery
and radioiodine ablation therapy, the serum TSH be kept between 0.3 and 2 mU/L.
For patients with high-risk invasive or metastatic disease, it is recommended that
the TSH level be maintained at ⬍0.1 mU/L indefinitely. For patients with high-risk
disease who have no evidence of active disease, the TSH target is 0.1 to 0.5 mU/L
(326).

Transient Hypothyroidism
Hypothyroidism may occasionally be transient, as in the early recovery phase of
postpartum thyroiditis, silent thyroiditis, and subacute thyroiditis. Thyroid hor-
mone deficiency is rarely severe during recovery from these conditions but it may
be symptomatic and may last as long as several months before the patient returns
to a euthyroid state. Permanent hypothyroidism is uncommon in subacute and
silent thyroiditis, but may occur in up to 25% of patients with postpartum thy-
roiditis (327–329). Thyroid hormone replacement is advisable for patients who
are symptomatic and for those whose serum TSH levels are ⬎20 mU/L. We rec-
ommend initiating treatment with an LT4 dose of 50 to 75 ␮g/day and a goal
TSH of 1.5 to 3.0 mU/L, thus permitting sufficient TSH secretion to stimulate
recovery of thyroid function as the inflammatory process subsides. After three to
six months of therapy, LT4 should be tapered in 25 ␮g decrements until eventual
discontinuation, provided serum TSH concentrations remain within the normal
range.
170 McDermott and Ridgway

Drug-Induced Hypothyroidism
Amiodarone is an anti-arrhythmic agent that is highly lipophilic, having a long
half-life of approximately 100 days and a propensity to accumulate in multiple
tissues, including the thyroid gland. Each molecule of amiodarone contains two
iodine atoms and the compound is 37% organic iodine by weight; a 200 mg
tablet delivers about 9 mg of iodine into the circulation. This medication inhibits
peripheral T4 to T3 conversion and thereby alters circulating thyroid hormone
concentrations (high T4, low T3, and transiently high TSH). It may also precipi-
tate overt symptomatic hypothyroidism or hyperthyroidism (40,41), particularly in
patients with underlying goiters or antithyroid antibodies living in iodine-replete
areas of the world, such as the United States. Hypothyroidism occurs in approx-
imately 20% of patients on chronic amiodarone therapy. The major mechanism
of this effect is probably high iodine delivery to the thyroid gland. High concen-
trations of intrathyroidal iodine acutely inhibit new thyroid hormone synthesis, a
phenomenon known as the Wolff-Chaikoff effect (330,331); while patients with
normal thyroid glands usually escape from this effect, patients with underlying
thyroid disease often do not and hypothyroidism ensues. Other potential mecha-
nisms for amiodarone-induced hypothyroidism include direct cytotoxic effects of
the drug on thyroid cells and possible initiation or augmentation of thyroid autoim-
munity. The condition resolves in approximately 50% of patients who discontinue
amiodarone administration but persists transiently or permanently in the remain-
der. We recommend treatment in most cases, particularly if they are symptomatic.
Because these patients have underlying heart disease, LT4 should be started in low
doses of 25 to 50 ␮g/day and increased carefully in 12.5 to 25 ␮g increments every
six to eight weeks in order to bring the TSH into the normal range. If arrhythmias
appear to be exacerbated with LT4 therapy, mildly elevated serum TSH levels may
be a more appropriate goal. Amiodarone treated patients have been reported to
require higher than expected LT4 replacement doses, possibly due to impaired T4
to T3 conversion within the pituitary gland (42).
Patients treated chronically with lithium also have a high likelihood of devel-
oping thyroid dysfunction. Approximately 50% of patients develop goiters, 20%
develop subclinical hypothyroidism, and up to 20% develop overt hypothyroidism
(40). Lithium administration is known to increase the intrathyroidal iodine con-
centration (43); lithium-induced hypothyroidism might therefore be due to the
Wolff-Chaikoff effect. Alternatively, since the majority of affected patients have
circulating antithyroid antibodies, lithium may also somehow enhance underlying
thyroid autoimmunity (44–46). If lithium withdrawal is not practical, we recom-
mend LT4 treatment for all patients who have goiters or elevated serum TSH
levels. The goal TSH should be 0.5 to 2.0 mU/L.
Interferon-alpha administration has been reported to cause thyroid dysfunc-
tion in up to 6% of patients; hypothyroidism and silent thyroiditis are the most
commonly found conditions and are believed to result from induction or aug-
mentation of preexisting thyroid autoimmunity (40,47–49). Interleukin-2 has also
Hypothyroidism 171

been implicated as a possible cause of silent (painless) thyroiditis (40,50). The


tyrosine kinase inhibitors, sunitinib and sorafenib, have been reported to cause
hypothyroidism, most likely by inducing thyroiditis or blocking iodine entry into
the thyroid (51–55). Discontinuation of these medications is often not desirable,
especially if they are showing efficacy against the serious disorders they are used
to treat. Accordingly, we recommend LT4 replacement therapy if these patients
are symptomatic, if their TSH levels are ⬎20 mU/L, or if TSH elevations of any
degree persist for more than three months. The goal TSH levels on treatment
should be 0.5 to 2.0 mU/L. Bexarotene, a retinoid x receptor ligand (rexinoid)
has been reported to cause central hypothyroidism through suppression of TSH
secretion (56,57). Administration of LT4 to normalize free T4 levels is the therapy
of choice.

Myxedema Coma
Patients suspected of having myxedema coma should be managed in an inten-
sive care unit. The cornerstone of treatment is rapid restoration of the thy-
roid hormone deficit (251–254). Investigators have recommended intravenous
LT4 (332,333) or LT3 (334,335) or a mixture of the two. Current com-
monly used regimens include the following. LT4 regimen: LT4 200 to 300 ␮g
intravenously over more than five minutes followed by oral or intravenous LT4
50 to 100 ␮g/day. LT3 followed by LT4 regimen: LT3 50 to 100 ␮g intravenously
over more than five minutes followed by oral or intravenous LT4 50 to 100 ␮g/day.
LT4 plus LT3 regimen: LT4 200 to 300 ␮g plus LT3 20 to 50 ␮g intravenously
over more than five minutes followed by oral or intravenous LT4 50 to 100 ␮g/day
plus LT3 20 to 25 ␮g/day in two divided doses. There are no randomized trials
available to settle the issue of which approach is best.
Myxedema coma is a syndrome in which severe thyroid hormone deficiency
is complicated by one or more precipitating events, as discussed earlier. It is there-
fore critical to identify and treat these precipitating events just as aggressively as
the thyroid hormone deficiency (251–254). Ventilatory and circulatory monitor-
ing and support are essential. Hypothermia may be corrected by slow rewarming
with blankets. Underlying infections should be actively sought and aggressively
treated when present. Hyponatremia and anemia are common features that should
be corrected with isotonic fluids, colloid and blood transfusions when indicated.
Hydrocortisone, 75 to 100 ␮g intravenously every six to eight hours, should also
be started without delay because thyroid hormone administration acutely increases
the metabolic clearance rate of cortisol in these patients (120–122) who may have
limited adrenal reserve or frank adrenal insufficiency. When initially described,
myxedema coma had a mortality rate of 100%, but with appropriate treatment, the
outlook for these patients is considerably improved. Even so, the mortality rates
in recent studies have been as high as 45% (255,256).
172 McDermott and Ridgway

Pitfalls in the Management of Hypothyroidism


Patients Whose Symptoms Do Not Resolve on LT4 Replacement
Presenting symptoms persist in some patients despite seemingly adequate LT4
replacement with normalization of serum TSH concentrations (336–339); this is
particularly true of patients with mild thyroid failure who have nonspecific symp-
toms such as fatigue, weight gain, and depression. The most likely explanation is
that these patients’ symptoms are unrelated to their thyroid hormone deficiency.
Coexisting disorders such as poor physical conditioning, ineffective sleep habits,
stress, endogenous depression, anemia, electrolyte and mineral abnormalities,
diabetes mellitus, and other systemic disorders should be evaluated and managed
appropriately.
Theoretically, using an LT4/LT3 combination that approximates the normal
thyroidal secretion ratio might be beneficial, particularly in patients whose initial
symptoms persist. In 1999, Bunevicius et al. reported that replacing 50 ␮g of
the total LT4 dose with 12.5 ␮g of LT3 for a five-week period resulted in sig-
nificantly improved neuropsychiatric function, memory, mood, and overall sense
of well-being compared to full replacement doses of LT4 alone (340,341). The
main criticism of this study was that supraphysiological amounts of T3 were used,
since 12.5 ␮g of T3 is more than twice what a normal thyroid gland secretes
daily (6 ␮g). Multiple subsequent studies of combination LT4/LT3 therapy fol-
lowed this initial report (342–350). Although a few reported a mild subjective
benefit (340,346,347,349), there has been no statistically significant evidence that
patients benefit from combined LT4/LT3 therapy compared to monotherapy with
levothyroxine alone. A review (351) and a meta-analysis (352) of all the trials
concluded there was no benefit from combination T4/T3 therapy. Moreover, it
has been clearly demonstrated that LT4 therapy alone given to thyroidectomized
individuals in doses that normalize the serum TSH levels produce serum T3 lev-
els that are indistinguishable from those in the euthyroid, prethyroidectomy state
(292). Nonetheless, the possibility remains that some patients with suboptimal
clinical responses to LT4 treatment might benefit from the addition of a small
dose of LT3 (5 ␮g/day) and a trial of combination therapy may be reasonable in
select patients. The future use of combination LT4/LT3 therapy will likely involve
the development of slow release T3 preparations and careful comparative studies,
using physiological amounts of LT4 and LT3, testing their efficacy, safety, and
optimal dosing regimens.

Changing Thyroid Hormone Requirements


LT4 dose requirements may change under various circumstances (Tables 9 and 10).
Pregnant women without endogenous thyroid function often require a 30% to 50%
LT4 dose increase during the first 20 weeks of pregnancy (280,353–356). This is
particularly important to recognize since inadequately treated maternal hypothy-
roidism may impair intellectual development of the fetus in utero (143–149).
Hypothyroidism 173

Table 9 Conditions Associated with Altered LT4


Dose Requirements

A. Reasons for increased LT4 dose requirements


1. Pregnancy
2. Use of drugs that decrease LT4 absorption
3. Use of drugs that increase LT4 metabolism
4. Use of drugs that decrease T4 to T3 conversion
5. Estrogen use
6. Malabsorption disorders
7. Nephrotic syndrome
8. Nonadherence
9. Progression of endogenous thyroid disease
B. Reasons for decreased LT4 dose requirements
1. Aging
2. Androgen use
3. Metformin use
4. Self-administration of excess LT4
5. Reactivation of Graves’ disease
6. Development of autonomous thyroid nodules

Table 10 Drugs That May Increase


Exogenous LT4 Dose Requirements

A. Drugs that decrease LT4 absorption


1. Ferrous sulfate
2. Calcium carbonate
3. Aluminum hydroxide
4. Sucralfate
5. Cholestyramine/colestipol
6. Fiber supplements
7. Soy supplements
8. Sevelamer
9. Raloxifene
10. Proton pump inhibitors
B. Drugs that increase LT4 metabolism
1. Phenytoin
2. Phenobarbital
3. Carbamazepine
4. Rifampin
C. Drugs that inhibit T4 to T3 conversion
1. Amiodarone
2. Glucocorticoids
3. Propranolol
4. Propylthiouracil
5. Ipodate
174 McDermott and Ridgway

Food and medications can interfere with the absorption of thyroid hormone in
the intestine; the most prominent effects have been reported with iron, calcium,
antacids, bile acid resins, fiber supplements, soy protein, raloxifene, and sevelamer
(357–365). Taking these products at a time of day that is separated by at least four
to eight hours from the LT4 dose will often bring about resolution of this problem.
Since gastric acid is important for LT4 absorption, patients with achlorhydria due
to chronic proton pump inhibitor use (366), Helicobacter pylori-related gastritis
(366), atrophic gastritis (366), or autoimmune gastritis (367) may have increased
LT4 requirements due to malabsorption of their medication (366,367). Intestinal
malabsorption due to Celiac disease (gluten sensitive enteropathy) (368) or to
lactose intolerance (369) should also be considered as a cause of escalating LT4
requirements and should be pursued when other causes are not evident. Moreover,
both autoimmune gastritis and Celiac disease occur more commonly in patients
with autoimmune thyroid disease (118,367,368). Nephrotic syndrome with mas-
sive urinary protein losses can also be a cause of increasing thyroid hormone
requirements (370). Other medications, particularly the anti-seizure and anti-
tuberculous agents, can significantly enhance hepatic T4 metabolism, resulting
in lower circulating serum T4 concentrations and the need for higher LT4 doses in
hypothyroid patients (371). Oral estrogen therapy has been shown to increase LT4
dose requirements by increasing circulating concentrations of thyroxine-binding
globulin and thereby reducing serum free thyroid hormone levels (372). Sertraline
may also increase thyroid hormone requirements but the mechanism of this effect
remains uncertain (373).
An apparent increase in LT4 requirements can also result from patient non-
adherence (374,375). One clue to this is the finding of an elevated serum TSH level
associated with a high-normal or elevated free T4. This profile suggests that the
patient missed multiple LT4 doses, resulting in elevation of the serum TSH, and
then took extra doses several days preceding blood sampling, acutely raising the
serum T4 concentration. The importance of adherence should be specifically dis-
cussed with all patients. An alternative solution to nonadherence is weekly, rather
than daily, ingestion of the calculated seven-day requirement (1.6–1.7 ␮g/kg ×
7) under the supervision of a health-care provider (376). Finally, patients who
have increasing LT4 needs may simply be having progression of their underlying
thyroid disease with worsening of endogenous thyroid function. When alteration
of the precipitating circumstances is not an option, the LT4 dose should be grad-
ually increased at six-week intervals as necessary to maintain the serum TSH
concentration in the desired range.
Conditions in which LT4 requirements are decreased are also listed in
Table 9. This is most commonly seen in elderly patients, in whom LT4 doses
are on average 25% lower than those in young patients (377,378). Prescribed or
surreptitious androgen use may also results in decreasing LT4 dose requirements
due to reductions in thyroxine-binding globulin with a consequent increase in free
thyroid hormone concentrations (379). Metformin has been reported to suppress
serum TSH levels by a postulated central mechanism (380); while this is not
Hypothyroidism 175

related to an increase in circulating thyroid hormone levels, dropping TSH levels


in response to metformin could lead one to erroneously believe that the LT4 dose
is too high. One must additionally consider the possibility that patients whose
LT4 dose needs are apparently dropping are intermittently or regularly taking
extra doses of their medications because they feel better on higher doses. While
difficult to confirm, this possibility can be investigated by reviewing pharmacy
refill records. Finally, increased endogenous thyroid hormone production due to
reactivation of Graves’ disease or to development of autonomously functioning
thyroid nodules must also be considered in patients whose LT4 dose requirements
decline. If the precipitating problem cannot be identified and corrected, the LT4
dose should be gradually decreased at six-week intervals in order to maintain
normal serum TSH levels.
Patients on stable doses of LT4 may have some inconsistency in their serum
TSH levels, varying within and slightly outside the normal range (303). We some-
times see patients with highly variable or “erratic” serum TSH levels (low, normal,
and high) while on the same doses of LT4. In these circumstances, poor or vari-
able adherence to the prescribed regimen must again be suspected. These patients
could also be taking their LT4 dose with iron, calcium, antacids, bile acid resins,
fiber supplements, or food without knowing the effects this may have on thyroid
medication absorption. Switching preparations between different brands can also
cause significant variations in serum TSH levels. Alternatively, minor differences
in serum TSH levels might simply result from having blood sampled at different
times of the day in relation to when the medication is taken. When serum TSH
levels vary significantly despite a seemingly constant LT4 dose, we recommend
that providers carefully review the patient’s medication administration habits and
instruct them to take their medications at the same time each day, to separate
their LT4 dose time by at least four to eight hours from food and drugs that may
interfere with LT4 absorption, and to consistently have their blood drawn for TSH
monitoring before their daily LT4 dose.

Patients Who Miss Their LT4 Dose


Patients sometimes inadvertently miss one or more daily doses of LT4. Some may
become quite concerned about potential adverse effects of this omission. While
missing doses should be discouraged, patients can be assured that LT4 has a long
half-life in the serum (seven days) and that there is virtually no harm resulting from
the occasional missed dose. However, the long serum half-life and slow absorption
of LT4 also allows providers the option of advising patients that taking an extra
“catch-up” pill the next day is permissible, particularly for those who appear to be
sensitive to even minor drops in their thyroid hormone levels.

Patients Who Are Treated With Desiccated Thyroid


Desiccated thyroid or “natural” thyroid preparations contain a higher LT3/LT4
ratio (1:4) than that produced by normal endogenous human thyroid secretion
176 McDermott and Ridgway

Table 11 Conversion from a Desiccated Thyroid Preparation to


Levothyroxine Therapy

Desiccated thyroid Levothyroxine


Equivalent
Dose T4 Content T3 Content Initial Dose

1.0 grain 38 ␮g 9 ␮g 100 ␮g


1.5 grains 57 ␮g 13.5 ␮g 150 ␮g
2.0 grains 76 ␮g 18 ␮g 200 ␮g
2.5 grains 95 ␮g 22.5 ␮g 250 ␮g
3.0 grains 114 ␮g 27 ␮g 300 ␮g
3.5 grains 133 ␮g 31.5 ␮g 350 ␮g
4.0 grains 152 ␮g 36 ␮g 400 ␮g

(approximately 1:10–20) (288,289). The excess LT3 content of these products


and the rapid absorption of T3 from the gastrointestinal tract often produce sup-
raphysiological serum T3 levels two to six hours following their ingestion (290).
Moreover, LT4 alone given in doses sufficient to normalize serum TSH levels has
been shown to normalize T3 levels also (292). For these reasons, most experts dis-
courage the use of desiccated thyroid and recommend that patients who are taking
these preparations be switched to brand-name LT4 products. Calculations that
take into account both T4 and T3 contents and rates of T4 to T3 conversion have
estimated that one grain (60 mg) of desiccated thyroid is approximately equal to
100 ␮g of LT4 (29,288,289). Accordingly, the recommended LT4 doses to be used
in patients being switch from desiccated thyroid products are shown in Table 11.
Once these changes are made, LT4 doses should be titrated by 12.5 to 25 ␮g every
six weeks until the serum TSH concentration is in the desired range. A significant
proportion of patients who have been converted from desiccated thyroid to LT4
complain of feeling worse, even when their TSH levels are normal and serum T4
levels high normal; this likely results from the lower serum T3 concentrations that
are present on LT4 therapy. Most patients, however, report that their symptoms
resolve over time. For those with persistent symptoms, deleting 25 to 50 ␮g of
LT4 and substituting 5 to 10 ␮g of LT3 might provide symptomatic relief. Caution
must still be exercised with such a regimen in order to avoid supraphysiological
T3 levels following medication ingestion, particularly in elderly subjects. As
discussed earlier, firm recommendations regarding LT4/LT3 combinations must
await the future development and careful testing of slow release T3 preparations.

Patients Who Are Sensitive to Thyroid Hormone Tablets


Patients occasionally complain of being “allergic” to thyroid hormone prepara-
tions. Provided the dose is correct, it is unlikely that a patient would have a true
allergic reaction to a hormone that is normally present in their body. However, a
patient may have sensitivity to a component of the pill such as coloring dye or
Hypothyroidism 177

a filler substance. When this is suspected, the 50 ␮g size of LT4, which in most
preparations is white with no added dye, can be given in quantities that add up
to the total intended dose. If reactions to the medication persist, changing to a
different preparation may be helpful.

Patients Who Are Temporarily Unable to Take LT4 Orally


Hypothyroid patients on LT4 replacement may be unable to take oral medications
for several days following a variety of general surgical procedures or during a
significant medical illness. If the anticipated duration of abstinence from LT4 is
no more than three days, no LT4 replacement during this brief interval is necessary.
However, if the period of abstinence is anticipated to be longer than three days,
the patient should be treated with intravenous LT4 in a daily dose that is 80% of
their usual oral dose; this amount is based on the determination that approximately
80% of an oral dose is absorbed into the circulation (301–303).

Surgery in Patients With Hypothyroidism


Surgery of various types may sometimes be needed in hypothyroid patients who
have been recently diagnosed but are not yet on adequate LT4 replacement. Most
studies indicate that hypothyroid patients tolerate surgery well and heal appro-
priately (381–383). However, in patients with more severe degrees of untreated
hypothyroidism, there is greater potential for some complications such as periop-
erative heart failure, ileus, absence of fever when infected, and neuropsychiatric
symptoms (384). It is advisable therefore to postpone elective surgical procedures
until patients are on LT4 replacement and have TSH levels that are in or near the
normal range (TSH ⬍10 mU/L). Patients requiring emergency surgery, including
coronary artery revascularization procedures (382,383,385), can proceed directly
to surgery without LT4 replacement until the postoperative period, at which time
dose schedules should be initiated, as discussed earlier, according to the patient’s
age, severity of hypothyroidism, and underlying general health.

Treatment of Nonthyroid Conditions With Thyroid Hormone


Obesity
Hypothyroidism is commonly associated with mild weight gain while thyrotoxi-
cosis usually causes weight loss. These observations prompted the use of high dose
thyroid hormone therapy to induce weight loss in the past (386,387), although this
practice was later abandoned because of the catabolic and potentially dangerous
consequences of iatrogenic thyrotoxicosis. Nonetheless, studies of obese subjects
placed on very low-calorie diets have shown that serum T3 concentrations and
resting metabolic rates decline in parallel, suggesting that these changes may con-
tribute to the plateau of body weight that is often seen after initial weight loss
occurs (388–391). High-dose T3 given during a very low-calorie diet has been
shown to prevent the drop in resting metabolic rate and to promote further weight
loss; however, this occurs at the expense of increased nitrogen loss suggesting
178 McDermott and Ridgway

protein catabolism (392–394). In contrast, lower-dose T3 therapy (40 ␮g/day)


has been shown in two small studies to promote weight loss in subjects on very
low-calorie diets without negatively affecting nitrogen balance (395,396).
At this time, there can be no justification for the use of thyroid hormone
preparations to assist with weight loss, except in hypothyroid patients in whom
only replacement LT4 doses are indicated. Nevertheless, the encouraging results
mentioned earlier will hopefully stimulate further study into the potential use of
new thyroid hormone analogs as an adjunct in the management of obesity.

Depression
Depression has been clearly identified as a symptom that may result from both
overt (139–141) and subclinical (235–239) hypothyroidism and that improves or
resolves with LT4 replacement therapy (141,338). Moreover, depression in appar-
ently euthyroid individuals has been shown to be associated with abnormalities of
the hypothalamic-pituitary-thyroid axis including abnormal TSH responses to the
administration of intravenous TRH (141). It is not surprising then that thyroid hor-
mone therapy has been used in the treatment of depression, even in patients with
normal thyroid function (397,398). LT4 has been used more commonly in bipolar
affective disorders, while LT3 has been given more often in depression (398).
While there is no substantial evidence that thyroid hormone alone has significant
antidepressant activity in euthyroid individuals, existing data do suggest that thy-
roid hormone supplementation may enhance the effect of standard antidepressant
medications in patients with refractory depression (397–400).

Premenstrual Syndrome
Disorders of the thyroid axis have been postulated to play a role in at least a subset
of patients with premenstrual syndrome (401). Controlled studies, however, have
failed to reveal any significant abnormalities of thyroid function or beneficial
responses to thyroid hormone therapy in patients with this disorder (402–404).
Therefore, there does not appear to be a current role for thyroid hormone therapy
in the management of premenstrual syndrome.
Consequences of Excess Thyroid Hormone Replacement
While our emphasis has been on the use of adequate doses of LT4 to alleviate the
symptoms of hypothyroidism, a caution regarding overtreatment must be raised
as well. If some is good, more is not necessarily better. Excess LT4 administration
most commonly occurs when serum TSH is not monitored regularly or when
patients request or self-administer higher LT4 doses to improve their energy level
or to help with weight loss. LT4 excess must be avoided, however, because of the
potentially deleterious effects on the cardiovascular system and the skeleton.
Low TSH levels in elderly subjects have clearly been demonstrated to be
associated with an increased risk of developing atrial fibrillation (231,405,406).
More subtle abnormalities of cardiac function, including increased heart rate,
atrial premature beats, increased left ventricular mass, increased left ventricular
Hypothyroidism 179

contractility, diastolic dysfunction, and impaired cardiac reserve have been demon-
strated in some studies during periods of even mild thyroid hormone excess (407–
410). Thyrotoxicosis has also been shown to stimulate osteoclastic bone resorption,
resulting in loss of bone mass. Although not a universal finding, several studies
have demonstrated that even mild thyroid hormone excess, as occurs in patients
taking excessive thyroid hormone doses, is associated with an increased risk of
bone loss and fractures, particularly in postmenopausal women (411–416). Thus,
there is ample evidence that even mild degrees of thyroid hormone excess should
be avoided in hypothyroid patents on chronic thyroid hormone replacement ther-
apy. Unfortunately, it has been shown that up to 20% of patients taking thyroid
hormone are over replaced (18).

SUMMARY
Hypothyroidism is a common disorder having multiple etiologies. Primary
hypothyroidism occurs far more frequently than does central hypothyroidism.
Thyroid hormone deficiency affects many tissues and organ systems, resulting
in a wide spectrum of clinical manifestations that include multiple symptoms,
signs, and laboratory abnormalities. Subclinical hypothyroidism is the mildest
form of this condition that may have physiologically relevant consequences and
beneficial responses to thyroid hormone replacement therapy. Myxedema coma,
the most severe form of hypothyroidism, is an extremely serious disorder with a
potentially high mortality rate without the prompt initiation of thyroid hormone
administration and appropriate treatment of precipitating causes.
Sensitive and accurate measurements of serum TSH levels have greatly
simplified the diagnosis of primary hypothyroidism. The diagnosis of central
hypothyroidism requires measurement of TSH along with free T4 or total T4 and
an assessment of other anterior pituitary functions. Measurement of antithyroid
antibodies helps to confirm the diagnosis of lymphocytic thyroiditis and aids in the
prediction of which patients with mild thyroid failure will progress on to develop
overt hypothyroidism.
The cornerstone of treatment of hypothyroidism is thyroid hormone replace-
ment. Treatment schedules and goals should be tailored to the individual patient
according to their age, severity of hypothyroidism, and underlying health status.
Serum TSH levels should be used to monitor LT4 therapy in patients with primary
hypothyroidism while free T4 levels are best when managing central hypothy-
roidism. The goals of treatment in most cases are to relieve the symptoms of
hypothyroidism and to avoid even subtle degrees of biochemical thyroid hormone
excess or deficiency.

REFERENCES
1. Larsen PR, Berry MJ. Type I iodothyronine deiodinase: Unexpected complexities in
a simple deiodination reaction. Thyroid 1994; 4:357–362.
180 McDermott and Ridgway

2. Braverman LE, Ingbar SH, Sterling K. Conversion of thyroxine to triiodothyronine


in athyrotic human subjects. J Clin Invest 1970; 49:855–864.
3. Pittman CS, Chambers JB, Read VH. The extrathyroidal conversion rate of thyroxine
to triiodothyronine in normal man. J Clin Invest 1971; 50:1187–1196.
4. Chin WW. Molecular mechanisms of thyroid hormone action. Thyroid 1994; 4:389–
393.
5. Brent GA. The molecular basis of thyroid hormone action. Mechanisms of Disease
1994; 331:847–853.
6. Oppenheimer JH, Schwartz HL, Sterling K. Thyroid hormone action 1994: The plot
thickens. Eur J Endocrinol 1994; 130:15–24.
7. Motomura K, Brent GA. Mechanisms of thyroid hormone action: Implications
for the clinical manifestation of thyrotoxicosis. Endocrinol Metab Clin NA 1998;
27:1–23.
8. Vanderpump MP, Tunbridge WM, French JM, et al. The incidence of thyroid dis-
orders in the community: A twenty-year follow-up of the Whickham survey. Clin
Endocrinol (Oxf) 1995; 43:55–68.
9. Flynn RW, MacDonald TM, Morris AD, et al. The thyroid epidemiology, audit, and
research study: Thyroid dysfunction in the general population. J Clin Endocrinol
Metab 2004; 89:3879–3884.
10. Tunbridge F, Evered DC, Hall R, et al. The spectrum of thyroid disease in a com-
munity: The Whickham survey. Clin Endocrinol (Oxf) 1977; 7:481–493.
11. Helfand M, Redfern CC. Clinical Guideline Part 2: Screening for thyroid disease:
An update. Ann Intern Med 1998; 129:144–158.
12. Ross DS. Subclinical hypothyroidism. In: Braverman LE, Utiger RD, eds. Werner
and Ingbar’s The Thyroid, 7th ed. Philadelphia, PA: Lippencott-Raven, 1996:1010–
1015.
13. Bagachi N, Brown TR, Parish RF. Thyroid dysfunction in adults over age 55 years.
A study in an urban U.S. community. Arch Intern Med 1990; 150:785–787.
14. Guel KW, van Sluisveld IL, Grobbee DE, et al. The importance of thyroid microso-
mal antibodies in the development of elevated serum TSH in middle-aged women:
Associations with serum lipids. Clin Endocrinol (Oxf) 1993; 39:275–280.
15. Sawin CT, Chopra D, Azizi F, et al. The aging thyroid. Increased prevalence of
elevated serum thyrotropin levels in the elderly. JAMA 1979; 242:247–250.
16. Rosenthal MJ, Hunt WC, Garry PJ, et al. Thyroid failure in the elderly. Microsomal
antibodies as discriminant for therapy. JAMA 1987; 258:209–213.
17. Parle JV, Franklyn JA, Cross KW, et al. Prevalence and follow-up of abnormal
thyrotrophin (TSH) concentrations in the elderly in the United Kingdom. Clin
Endocrinol (Oxf) 1991; 34:77–83.
18. Canaris GJ, Manowitz NR, Mayor G, et al. The Colorado thyroid disease prevalence
study. Arch Intern Med 2000; 160:526–534.
19. Hollowell JG, Staehling NW, Flanders WD, et al. Serum TSH, T4, and thyroid anti-
bodies in the United States population (1988 to 1994): National Health and Nutrition
Examination Survey (NHANES III). J Clin Endocrinol Metab 2002; 87:489–499.
20. Surks MI, Hollowell JG. Age-specific distribution of serum thyrotropin and antithy-
roid antibodies in the US population: Implications for the prevalence of subclinical
hypothyroidism. J Clin Endocrinol Metab 2007; 92:4575–4582.
21. Tunbridge F, Brewis M, French JM, et al. Natural history of autoimmune thyroiditis.
Br Med J 1981; 282:258–262.
Hypothyroidism 181

22. Kabadi UM. “Subclinical hypothyroidism.” Natural course of the syndrome during
a prolonged follow-up study. Arch Intern Med 1993; 153:957–961.
23. Huber G, Staub JJ, Meier C, et al. Prospective study of the spontaneous course of
subclinical hypothyroidism: Prognostic value of thyrotropin, thyroid reserve, and
thyroid antibodies. J Clin Endocrinol Metab 2002; 87:3221–3226.
24. Diez JJ, Iglesias P. Spontaneous subclinical hypothyroidism in patients older than
55 years: An analysis of natural course and risk factors for the development of overt
thyroid failure. J Clin Endocrinol Metab 2004; 89:4890–4897.
25. Diez JJ, Iglesias P, Burman KD. Spontaneous normalization of thyrotropin concen-
trations in patients with subclinical hypothyroidism. J Clin Endocrinol Metab 2005;
90:4124–4127.
26. Meyerovitch J, Rotman-Pikielny P, Sherf M, et al. Serum thyrotropin measurements
in the community: Five year follow-up in a large network of primary care physicians.
Arch Intern Med 2007; 167:1533–1538.
27. Samuels MH, Ridgway EC. Central hypothyroidism. Endocrinol Metab Clin NA
1992; 21:903–919.
28. Martino E, Bartalena L, Faglia G, et al. Central hypothyroidism. In: Braverman
LE, Utiger RD, eds. Werner and Ingbar’s The Thyroid, 7th ed. Philadelphia, PA:
Lippincott-Raven, 1996:779–791.
29. Larsen PR, Davies TF, Hay IL. Thyroid hormone deficiency. In: Wilson JD, Foster
DW, Kronenberg HM, Larsen PR, eds. William’s Textbook of Endocrinology, 9th
ed. Philadelphia, PA:W. B. Saunders, 1998:460–475.
30. Dayan CM, Daniels GH. Chronic autoimmune thyroiditis. N Engl J Med 1996;
335:99–107.
31. Nogueira CR, Nguyen LQ, Coelho-Neto JR, et al. Structural analysis of the thy-
rotropin receptor in four patients with congenital hypothyroidism due to thyroid
hypoplasia. Thyroid 1999; 9:523–529.
32. Kosugi S, Bhayana S, Dean HJ. A novel mutation in the sodium/iodide symporter
gene in the largest family with iodide transport defect. J Clin Endocrinol Metab
1999; 84:3248–3253.
33. Pannain S, Weiss RE, Jackson CE, et al. Two different mutations in the thyroid
peroxidase gene of a large inbred Amish kindred: Power and limits of homozygosity
mapping. J Clin Endocrinol Metab 1999; 84:1061–1071.
34. Targovnik HM, Frechtel GD, Mendive FM, et al. Evidence for the segregation of
three different mutated alleles of the thyroglobulin gene in a Brazilian family with
congenital goiter and hypothyroidism. Thyroid 1998; 8:291–297.
35. Medeiros-Neto G, Herodotou DT, Rajan S, et al. A circulating biologically inactive
thyrotropin caused by a mutation in the beta subunit gene. J Clin Invest 1996;
97:1250–1256.
36. Collu R, Tang J, Castagne J, et al. A novel mechanism for isolated central hypothy-
roidism: Inactivating mutations in the thyrotropin-releasing hormone receptor gene.
J Clin Endocrinol Metab 1997; 82:1561–1565.
37. Cohen LE, Wondisford FE, Salvatoni A, et al. A “hot spot” in the Pit-1 gene respon-
sible for combined pituitary hormone deficiency: Clinical and molecular correlates.
J Clin Endocrinol Metab 1995; 80:679–684.
38. Deladoey J, Fluck C, Buyukgebiz A, et al. “Hot spot” in the PROP1 gene responsible
for combined pituitary hormone deficiency. J Clin Endocrinol Metab 1999; 84:1645–
1650.
182 McDermott and Ridgway

39. McDermott MT, Haugen BR, Black JN, et al. Congenital isolated central hypothy-
roidism causes by a “hot spot” mutation in the thyrotropin-beta gene. Thyroid 2002;
12:1141–1146.
40. Surks MI, Sievert R. Drugs and thyroid function. N Engl J Med 1995; 333:1688–
1694.
41. Harjai KJ, Licata AA. Effects of amiodarone on thyroid function. Ann Intern Med
1997; 126:63–73.
42. Figge J, Dluhy RG. Amiodarone-induced elevation of thyroid stimulating hormone
in patients receiving levothyroxine for primary hypothyroidism. Ann Intern Med
1990; 113:553–555.
43. Berens SC, Berstein RS, Robbins J, et al. Antithyroid effects of lithium. J Clin Invest
1970; 49:1357–1367.
44. Deniker P, Eyquem A, Bernheim R, et al. Thyroid autoantibody levels during lithium
therapy. Neuropsychobiology 1978; 4:270–275.
45. Gelfand EW, Dosch HM, Hastings B, et al. Lithium: A modulator of cyclic AMP-
dependent events in lymphocytes? Science 1979; 203:365–367.
46. Dosch HM, Matheson D, Schuurman RK, et al. Anti-suppressor cell effects of lithium
in vitro and in vivo. Adv Exp Med Biol 1980; 127:447–462.
47. Roti E, Minelli R, Giuberti T, et al. Multiple changes in thyroid function in patients
with chronic active HCV hepatitis treated with recombinant interferon-alpha. Am J
Med 1996; 172:482–487.
48. Schuppert F, Rambusch E, Kirchner H, et al. Patients treated with interferon-alpha,
interferon-beta, and interleukin-2 have a different thyroid autoantibody pattern than
patients suffering from endogenous autoimmune thyroid disease. Thyroid 1997;
7:837–842.
49. Koh L, Greenspan FS, Yeo P. Interferon-alpha induced thyroid dysfunction: Three
clinical presentations and a review of the literature. Thyroid 1997; 7:891–896.
50. Atkins MB, Mier JW, Parkinson DR, et al. Hypothyroidism after treatment with
interleukin-2 and lymphokine-activated killer cells. N Engl J Med 1988; 318:1557–
1563.
51. Desai J, Yassa L, Marquesee E, et al. Hypothyroidism after sunitinib treatment for
patients with gastrointestinal stromal tumors. Ann Intern Med 2006; 145:660–664.
52. Rini BI, Tamaskar I, Shaheen P, et al. Hypothyroidism in patients with metastatic
renal cell carcinoma treated with sunitinib. J Natl Cancer Inst 2007; 99:81–83.
53. Wong E, Rosen LS, Mulay M, et al. Sunitinib induces hypothyroidism in advanced
cancer patients and may inhibit thyroid peroxidase activity. Thyroid 2007; 17:351–
355.
54. Mannavola D, Coco P, Vannucchi G, et al. A novel tyrosine kinase selective
inhibitor, sunitinib, induces transient hypothyroidism by blocking iodine uptake.
J Clin Endocrinol Metab 2007; 92:3531–3534.
55. Tamaskar I, Bukowski R, Elson P, et al. Thyroid function test abnormalities in
patients with metastatic renal cell carcinoma treated with sorafenib. Ann Oncol
2008; 19:265–268.
56. Sherman SI, Gopal J, Haugen BR, et al. Central hypothyroidism associated with
retinoid X receptor-selective ligands. N Engl J Med 1999; 340:1075–1079.
57. Golden WM, Weber KB, Hernandez TL, et al. Single-dose rexinoid rapidly and
specifically suppresses serum thyrotropin in normal subjects. J Clin Endocrinol
Metab 2007; 92:124–130.
Hypothyroidism 183

58. Billewicz W, Chapman RS, Crooks J, et al. Statistical methods applied to the diag-
nosis of hypothyroidism. Q J Med 1969; 38:255–266.
59. Oddie TH, Boyd CM, Fisher DA, et al. Incidence of signs and symptoms in thyroid
disease. Med J Aust 1972; 2:981–986.
60. Means JH. Relative frequency of the several symptoms and signs of myxedema. In:
Means JH, ed. The Thyroid and Its Diseases, 2nd ed. Philadelphia, PA: JB Lippencott
Company, 1948:232–234.
61. Zulewski HK, Muller B, Exer P, et al. Estimation of tissue hypothyroidism by a new
clinical score: Evaluation of patients with various grades of hypothyroidism and
controls. J Clin Endocrinol Metab 1997; 82:771–776.
62. Canaris GJ, Steiner JF, Ridgway EC. Do traditional symptoms of hypothyroidism
correlate with biochemical disease? J Gen Intern Med 1997; 12:544–550.
63. Doucet J, Trivalle C, Chassagne P, et al. Does age play a role in the clinical presen-
tation of hypothyroidism? J Am Geriatr Soc 1994; 42:984–986.
64. Wilson WR, Bedell GN. The pulmonary abnormalities in myxedema. J Clin Invest
1960; 39:42–55.
65. Nordqvist P, Dhuner KG, Stenberg K, et al. Myxedema coma and CO2 retention.
Acta Med Scand 1960; 166:189–194.
66. Massumi RA, Winnacker JL. Severe depression of the respiratory center in
myxedema. Am J Med 1964; 36:876–882.
67. Weg JG, Calverly JR, Johnson C. Hypothyroidism and alveolar hypoventilation.
Arch Intern Med 1965; 115:302–306.
68. Burack R, Edwards RH, Green M, et al. The response to exercise before and after
treatment of myxedema with thyroxine. J Pharmacol Exp Ther 1971; 176:212–219.
69. Laroche CM, Cairns T, Moxham J, et al. Hypothyroidism presenting with respiratory
muscle weakness. Am Rev Respir Dis 1988; 138:472–474.
70. Martinez FJ, Bermudez-Gomez M, Celli BR. Hypothyroidism. A reversible cause
of diaphragmatic dysfunction. Chest 1989; 96:1059–1063.
71. Siafakas NM, Salesiotou V, Filaditaki V, et al. Respiratory muscle strength in
hypothyroidism. Chest 1992; 102:189–194.
72. Zondek H, Michael M, Kaatz A. The capillaries in myxedema. Am J Med Sci 1941;
202:435–440.
73. Lange K. Capillary permeability in myxedema. Am J Med Sci 1944; 208:5–15.
74. Sachdev Y, Hall R. Effusions into body cavities in hypothyroidism. Lancet 1975;
1:564–566.
75. Gottehrer A, Roa J, Stanford GG, et al. Hypothyroidism and pleural effusions. Chest
1990; 98:1130–1132.
76. Zwillich CW, Pierson DJ, Hofeldt FD, et al. Ventilatory control in myxedema and
hypothyroidism. N Engl J Med 1975; 292:662–665.
77. Ladenson PW, Goldenheim PD, Ridgway EC. Prediction and reversal of blunted
ventilatory responsiveness in patients with hypothyroidism. Am J Med 1988; 84:877–
883.
78. Duranti R, Gheri RG, Gorini M, et al. Control of breathing in patients with severe
hypothyroidism. Am J Med 1993; 95:29–37.
79. Orr WC, Males JL, Imes NK. Myxedema and obstructive sleep apnea. Am J Med
1981; 70:1061–1066.
80. Skatrud J, Iber C, Ewart R, et al. Disordered breathing during sleep in hypothy-
roidism. Am Rev Respir Dis 1981; 124:325–329.
184 McDermott and Ridgway

81. Rajagopal KR, Abbrecht PH, Derderian SS, et al. Obstructive sleep apnea in hypothy-
roidism. Ann Intern Med 1984; 101:491–494.
82. Pelttari L, Rauhala E, Polo O, et al. Upper airway obstruction in hypothyroidism.
J Intern Med 1994; 236:177–181.
83. Millman RP, Bevilacqua J, Peterson DD, et al. Central sleep apnea in hypothyroidism.
Am Rev Respir Dis 1983; 127:504–507.
84. Pandya K, Lal C, Scheinhorn D, et al. Hypothyroidism and ventilator dependency.
Arch Intern Med 1989; 149:2115–2116.
85. Klein I, Levey GS. Unusual manifestations of hypothyroidism. Arch Intern Med
1984; 144:123–128.
86. Kabadi UM, Kumar SP. Pericardial effusion in primary hypothyroidism. Am Heart
J 1990; 120:1393–1395.
87. Kerber RE, Sherman B. Echocardiographic evaluation of pericardial effusion in
myxedema. Incidence and biochemical and clinical correlations. Circulation 1975;
52:823–827.
88. Manolis AS, Varriale P, Ostrowski RM. Hypothyroid cardiac tamponade. Arch Intern
Med 1987; 147:1167–1169.
89. Klein I, Ojamaa K. Clinical review 36: Cardiovascular manifestations of endocrine
disease. J Clin Endocrinol Metab 1992; 75:339–342.
90. Polikar R, Burger AG, Scherrer U, et al. The thyroid and the heart. Circulation 1993;
87:1435–1441.
91. Amidi M, Leon DF, DeGroot WJ, et al. Effect of the thyroid state on myocardial
contractility and ventricular ejection rate in man. Circulation 1968; 38:229–239.
92. Crowley WF Jr, Ridgway EC, Bough EW, et al. Noninvasive evaluation of cardiac
function in hypothyroidism. Response to gradual thyroxine replacement. N Engl J
Med 1977; 296:1–6.
93. Tseng KH, Walfish PG, Persaud JA, et al. Concurrent aortic and mitral valve echocar-
diography permits measurement of systolic time intervals as an index of peripheral
tissue thyroid functional status. J Clin Endocrinol Metab 1989; 69:633–638.
94. Klein I, Ojamaa K. Thyroid hormone and the cardiovascular system: From theory to
practice. J Clin Endocrinol Metab 1994; 78:1026–1027.
95. Wieshammer S, Keck FS, Waitsinger J, et al. Acute hypothyroidism slows the
rate of left ventricular diastolic relaxation. Can J Physiol Pharmacol 1989; 67:
1007–1010.
96. Vora J, O’Malley BP, Petersen S, et al. Reversible abnormalities of myocardial
relaxation in hypothyroidism. J Clin Endocrinol Metab 1985; 61:269–272.
97. Tielens ET, Pillay M, Storm C, et al. Cardiac function at rest in hypothyroidism
evaluated by equilibrium radionuclide angiography. Clin Endocrinol 1999; 50:497–
502.
98. Santos AD, Miller RP, Mathew PK, et al. Echocardiographic characterization of the
reversible cardiomyopathy of hypothyroidism. Am J Med 1980; 68:675–682.
99. Anthonisen P, Holst E, Thomsen AA. Determination of cardiac output and other
hemodynamic data in patients with hyper- and hypothyroidism, using dye dilution
technique. Scand J Clin Lab Invest 1960; 12:472–480.
100. Klein I, Ojamaa K. The cardiovascular system in hypothyroidism. In: Braverman
LE, Utiger RD, eds. Werner and Ingbar’s The Thyroid, 7th ed. Philadelphia, PA:
Lippincott-Raven, 1996:799–804.
Hypothyroidism 185

101. Ladenson PW, Sherman SI, Baughman KL, et al. Reversible alterations in myocardial
gene expression in a young man with dilated cardiomyopathy and hypothyroidism.
Proc Natl Acad Sci USA 1992; 89:5251–5255.
102. Tunbridge WM, Evered DC, Hall R, et al. Lipid profiles and cardiovascular disease
in the Whickham area with particular reference to thyroid failure. Clin Endocrinol
(Oxf) 1977; 7:495–508.
103. Streeten DHP, Anderson GH Jr, Howland T, et al. Effects of thyroid function on blood
pressure: Recognition of hypothyroid hypertension. Hypertension 1988; 11:78–83.
104. Bernstein R, Muller C, Midtbo K, et al. Silent myocardial ischemia in hypothy-
roidism. Thyroid 1995; 5:443–447.
105. Oboubie K, Smith J, Evans LM, et al. Increased central arterial stiffness in hypothy-
roidism. J Clin Endocrinol Metab 2002; 87:4662–4666.
106. Nagasaki T, Inaba M, Henmi Y, et al. Decrease in carotid intima-media thickness in
hypothyroid patients after normalization of thyroid function. Clin Endocrinol (Oxf)
2003; 59:607–612.
107. del Palacio A, Trueba JL, Cabello A, et al. Hypothyroid myopathy. Clinicopathologic
study of 20 cases. An Med Interna 1990; 7:115–119.
108. Burnett JR, Crooke MJ, Delahunt JW, et al. Serum enzymes in hypothyroidism.
N Z Med J 1994; 107:355–356.
109. Khaleeli AA, Griffith DG, Edwards RH. The clinical presentation of hypothyroid
myopathy and its relationship to abnormalities in structure and function of skeletal
muscle. Clin Endocrinol (Oxf) 1983; 19:365–376.
110. Beyer IW, Karmali R, DeMeester-Mirkine N, et al. Serum creatine kinase levels in
overt and subclinical hypothyroidism. Thyroid 1998; 8:1029–1031.
111. Miyamoto T, Nagasaka A, Kato K, et al. Immunoreactive creatine kinase-MB and
creatine kinase isozyme concentrations during treatment of hypothyroid patients.
Eur J Clin Chem Clin Biochem 1994; 32:589–593.
112. LeMar HJ Jr, West SG, Garrett CR, et al. Covert hypothyroidism presenting as a
cardiovascular event. Am J Med 1991; 91:549–552.
113. Contreras P, Generini G, Michelsen H, et al. Hyperprolactinemia and galactor-
rhea: Spontaneous versus iatrogenic hypothyroidism. J Clin Endocrinol Metab 1981;
53:1036–1039.
114. Honbo KS, van Herle AJ, Kellett KA. Serum prolactin levels in untreated primary
hypothyroidism. Am J Med 1978; 64:782–787.
115. Groff TR, Shulkin BL, Utiger RD, et al. Amenorrhea-galactorrhea, hyperprolactine-
mia, and suprasellar pituitary enlargement as presenting features of primary hypothy-
roidism. Obstet Gynecol 1984; 63:86S–89S.
116. Grubb MR, Chakeres D, Malarkey WB. Patients with primary hypothyroidism pre-
senting as prolactinomas. Am J Med 1987; 83:765–769.
117. Poretsky L, Garber J, Kleefield J. Primary amenorrhea and pseudoprolactinoma in a
patient with primary hypothyroidism. Am J Med 1986; 81:180–182.
118. Eisenbarth GS, Gottlieb PA. Autoimmune polyendocrine syndromes. N Engl J Med
2004; 350:2068–2079.
119. Means JH, Hertz S, Lerman J. The pituitary type of myxedema or Simmonds’ disease
masquerading as myxedema. Trans Assoc Am Physicians 1940; 55:32–53.
120. Peterson RE. The influence of the thyroid on adrenal cortical function. J Clin Invest
1958; 37:736–743.
186 McDermott and Ridgway

121. Peterson RE. The miscible pool and turnover rate of adrenocortical steroids in man.
Recent Prog Horm Res 1959; 15:231–274.
122. Iranmanesh A, Lizarralde G, Johnson ML, et al. Dynamics of 24-hour endogenous
cortisol secretion and clearance in primary hypothyroidism assessed before and after
partial thyroid hormone replacement. J Clin Endocrinol Metab 1990; 70:155–161.
123. Root AW, Rosenfield RL, Bongiovanni AM, et al. The plasma growth hormone
response to insulin-induced hypoglycemia in children with retardation of growth.
Pediatrics 1999; 39:844–852.
124. MacGillivray MH, Aceto T Jr, Frohman LA. Plasma growth hormone responses and
growth retardation of hypothyroidism. Am J Dis Child 1968; 115:273–276.
125. Katz HP, Youlton R, Kaplan SL, et al. Growth and growth hormone III. Growth
hormone release in children with primary hypothyroidism and thyrotoxicosis. J Clin
Endocrinol Metab 1969; 29:346–351.
126. Rivkees SA, Bode HH, Crawford JD. Long-term growth in juvenile acquired
hypothyroidism. N Engl J Med 1988; 318:599–602.
127. Hayles AB, Cloutier MD. Clinical hypothyroidism in the young—a second look.
Med Clin North Am 1972; 56:871–884.
128. Thomas R, Reid RL. Thyroid disease and reproductive dysfunction: A review. Obstet
Gynecol 1987; 70:789–798.
129. Kendle F. Case of precocious puberty in a female cretin. Br Med J 1905; 1:246.
130. Van Wyk J, Grumbach MM. Syndrome of precocious menstruation and galactorrhea
in juvenile hypothyroidism: An example of hormonal overlap pituitary feedback.
J Pediatr 1960; 57:416–435.
131. Wortsman J, Rosner W, Dufau ML. Abnormal testicular function in men with primary
hypothyroidism. Am J Med 1987; 82:207–212.
132. De La Balze FA. Male hypogonadism in hypothyroidism: A study of six cases.
J Clin Endocrinol Metab 1962; 22:212–222.
133. Krassas GE, Pontikides N. Male reproductive function in relation with thyroid alter-
ations. Best Pract Res Clin Endocrinol Metab 2004; 18:183–195.
134. Gull WW. On a cretinoid state supervening in adult life in women. Trans Clin Soc
(Lond) 1873; 7:180–185.
135. Inglis T. Two cases of myxoedema. Lancet 1880; 2:496–497.
136. Report of a committee of the Clinical Society of London. Report on myxedema.
Trans Clin Soc (Lond)(Suppl) 1888; 21:1–215.
137. Savage GH. Myxoedema and its nervous symptoms. J Ment Sci 1880; 25:417.
138. Asher R. Myxoedematous madness. Br Med J 1949; 2:555–562.
139. Whybrow PC. Behavioral and psychiatric aspects of hypothyroidism. In: Braverman
LE, ed. Werner and Ingbar’s The Thyroid, 7th ed. Philadelphia, PA: Lippencott-
Raven, 1996:866–870.
140. Pop VJ, Maartens LH, Leusink G, et al. Are autoimmune thyroid dysfunction and
depression related? J Clin Endocrinol Metab 1998; 83:3194–3197.
141. Jackson I. The thyroid axis and depression. Thyroid 1998; 8:951–956.
142. Oppenheimer JH, Schwartz HL. Molecular basis of thyroid hormone-dependent
brain development. Endocr Rev 1997; 18:462–475.
143. Haddow JE, Palomaki GE, Allan WC, et al. Maternal thyroid deficiency during
pregnancy and subsequent neuropsychological development of the child. N Engl J
Med 1999; 341:549–555.
Hypothyroidism 187

144. Man EB, Brown JF, Serunian SA. Maternal hypothyroxinemia: Psychoneurological
deficits of progeny. Ann Clin Lab Sci 1991; 21:227–239.
145. Pop VJ, Kuijpens JL, van Baar AL, et al. Low maternal free thyroxine concentrations
during early pregnancy are associated with impaired psychomotor development in
infancy. Clin Endocrinol 1999; 50:149–155.
146. Klein RZ, Mitchell ML. Maternal hypothyroidism and cognitive development of the
offspring. Curr Opin Pediatr 2002; 14:443–446.
147. Casey BM, Dashe JS, Wells CE, et al. Subclinical hypothyroidism and pregnancy
outcomes. Obstet Gynecol 2005; 105:239–245.
148. Casey BM, Dashe JS, Spong CY, et al. Perinatal significance of isolated maternal
hypothyroxinemia identified in the first half of pregnancy. Obstet Gynecol 2007;
109:1129–1135.
149. LaFranchi SH, Haddow JE, Hollowell JG. Is thyroid inadequacy during gestation
a risk factor for adverse pregnancy and developmental outcomes? Thyroid 2005;
15:60–71.
150. Phalen GS. The carpal-tunnel syndrome. Seventeen years’ experience in diagnosis
and treatment of six hundred fifty-four hands. J Bone Joint Surg Am 1966; 48:211–
228.
151. Earll JM, Kolb FO. Facial paralysis occurring with hypothyroidism. A report of two
cases. Calif Med 1967; 106:56–58.
152. Moosa A, Dubowitz V. Slow nerve conduction velocity in cretins. Arch Dis Child
1971; 46:852–854.
153. Beghi E, Delodovici ML, Bogliun G, et al. Hypothyroidism and polyneuropathy.
J Neurol Neurosurg Psychiatry 1989; 52:1420–1423.
154. Barnard RO, Campbell MJ, McDonald WI. Pathological findings in a case of
hypothyroidism with ataxia. J Neurol Neurosurg Psychiatry 1971; 34:755–760.
155. Khaleeli AA, Gohil K, McPhail G, et al. Muscle morphology and metabolism
in hypothyroid myopathy: Effects of treatment. J Clin Pathol 1983; 36:519–
526.
156. Mastaglia FL, Ojeda VJ, Sarnat HB, et al. Myopathies associated with hypothy-
roidism: A review based upon 13 cases. Aust N Z J Med 1988; 18:799–806.
157. DeLong GR. The neuromuscular system and brain in hypothyroidism. In: Braverman
LE, ed. Werner and Ingbar’s The Thyroid, 7th ed. Philadelphia, PA: Lippencott-
Raven, 1996:826–835.
158. Rodolico C, Toscano A, Benvenga S, et al. Myopathy as the persistently isolated
symptomatology of primary autoimmune hypothyroidism. Thyroid 1998; 8:1033–
1038.
159. Riggs JE. Acute exertional rhabdomyolysis in hypothyroidism: The result of a
reversible defect in glycogenolysis. Mil Med 1990; 155:171–172.
160. Hoffman J. Weiteger beitrag zur lehr von tetanie. Deutsch Z Nervenheilk 1897;
9:278.
161. Hetzel BS, Potter BJ, Dulberg EM. The iodine deficiency disorders: Nature, patho-
genesis and epidemiology. World Rev Nutr Diet 1990; 62:59–119.
162. Debre R, Semelaigne G. Syndrome of diffuse muscular hypertrophy in infants caus-
ing athletic appearance. Am J Dis Child 1935; 50:1351–1361.
163. Gupta OP, Bhatia PL, Agarwal MK, et al. Nasal, pharyngeal, and laryngeal mani-
festations of hypothyroidism. Ear Nose Throat J 1977; 56:349–356.
188 McDermott and Ridgway

164. Moley JR, Ohkawa M, Chaudry IH, et al. Hypothyroidism abolishes the hyper-
dynamic phase and increases susceptibility to sepsis. J Surg Res 1984; 36:265–
273.
165. Schoenfeld PS, Myers JW, Myers L, et al. Suppression of cell-mediated immunity
in hypothyroidism. South Med J 1995; 88:347–349.
166. Cooper DS, Halpern R, Wood LC, et al. L-thyroxine therapy in subclinical hypothy-
roidism. Ann Intern Med 1984; 101:18–24.
167. Nystrom E, Caidahl K, Fager G, et al. A double-blind cross-over 12-month study of
L-thyroxine treatment of women with “subclinical” hypothyroidism. Clin Endocrinol
1988; 29:63–76.
168. Staub JJ, Althaus BU, Engler H, et al. Spectrum of subclinical and overt hypothy-
roidism: Effect on thyrotropin, prolactin, and thyroid reserve, and metabolic impact
on peripheral target issues. Am J Med 1992; 92:631–642.
169. Kahaly GJ. Cardiovascular and atherogenic aspects of subclinical hypothyroidism.
Thyroid 2000; 10:665–670.
170. Meier C, Staub JJ, Roth CB, et al. TSH-controlled L-thyroxine therapy reduces
cholesterol levels and clinical symptoms in subclinical hypothyroidism: A double
blind, placebo-controlled trial (Basel Thyroid Study). J Clin Endocrinol Metab 2001;
86:4860–4866.
171. Biondi B, Palmieri EA, Lombardi G, et al. Effects of subclinical thyroid dysfunction
on the heart. Ann Intern Med 2002; 137:904–914.
172. Bianchi GP, Zaccheroni V, Solaroli E, et al. Health-related quality of life in patients
with thyroid disorders. Qual Life Res 2004; 13:45–54.
173. Carani C, Isidori AM, Granata A, et al. Multicenter study on the prevalence of sexual
symptoms in male hypo- and hyperthyroid patients. J Clin Endocrinol Metab 2005;
90:6472–6479.
174. Samuels MH, Schuff KG, Carlson NE, et al. Health status, mood, and cognition in
experimentally induced subclinical hypothyroidism. J Clin Endocrinol Metab 2007;
92:2545–2551.
175. Rasvi S, Ingoe L, Keeka G, et al. The beneficial effect of L-thyroxine on cardiovascu-
lar risk factors, endothelial function, and quality of life in subclinical hypothyroidism:
Randomized, crossover trial. J Clin Endocrinol Metab 2007; 92:1715–1723.
176. Cooper DS. Subclinical hypothyroidism. N Engl J Med 2001; 345:260–265.
177. Biondi B, Cooper DS. The clinical significance of subclinical thyroid dysfunction.
Endocr Rev 2008; 29:76–131.
178. Surks MI, Ortiz E, Daniels GH, et al. Subclinical thyroid disease. Scientific review
and guidelines for diagnosis and management. JAMA 2004; 291:228–238.
179. Col NF, Surks MI, Daniels MI. Subclinical thyroid disease. Clinical applications.
JAMA 2004; 291:239–243.
180. Ooi TC, Whitlock RM, Frengley PA, et al. Systolic time intervals and ankle reflex
time in patients with minimal serum TSH elevation: Response to triiodothyronine
therapy. Clin Endocrinol (Oxf) 1999; 13:621–627.
181. Ridgway EC, Cooper DS, Walker H, et al. Peripheral responses to thyroid hormone
before and after L-thyroxine therapy in patients with subclinical hypothyroidism.
J Clin Endocrinol Metab 1981; 53:1238–1242.
182. Bell GM, Todd WT, Forfar JC, et al. End-organ responses to thyroxine therapy in
subclinical hypothyroidism. Clin Endocrinol (Oxf) 1985; 22:83–89.
Hypothyroidism 189

183. Forfar JC, Wathen CG, Todd WT, et al. Left ventricular performance in subclinical
hypothyroidism. Q J Med 1985; 224:857–865.
184. Foldes J, Istvanfy M, Halmagyi M, et al. Hypothyroidism and the heart. Examination
of left ventricular function in subclinical hypothyroidism. Acta Med Hung 1987;
44:337–347.
185. Arem R, Rokey R, Kiefe C, et al. Cardiac systolic and diastolic function at rest and
exercise in subclinical hypothyroidism: Effect of thyroid hormone therapy. Thyroid
1996; 6:397–402.
186. Biondi B, Fazio S, Palmieri EA, et al. Left ventricular diastolic dysfunction in
patients with subclinical hypothyroidism. J Clin Endocrinol Metab 1999; 84:2064–
2067.
187. Monzani F, Di Bello V, Caraccio N, et al. Effect of levothyroxine on cardiac function
and structure in subclinical hypothyroidism: A double blind, placebo-controlled
study. J Clin Endocrinol Metab 2001; 86:1110–1115.
188. Vitale G, Galderisi M, Lupoli GA, et al. Left ventricular myocardial impairment in
subclinical hypothyroidism assessed by a new ultrasound tool: Pulsed tissue Doppler.
J Clin Endocrinol Metab 2002; 87:4350–4355.
189. Biondi B, Palmieri EA, Lombardi G, et al. Subclinical hypothyroidism and cardiac
function. Thyroid 2002; 12:505–510.
190. Brenta G, Mutti LA, Schnitman M, et al. Assessment of left ventricular diastolic
function by radionuclide ventriculography at rest and exercise in subclinical hypothy-
roidism, and its response to L-thyroxine therapy. Am J Cardiol 2003; 91:1327–1330.
191. Yazici M, Gorgulu S, Sertbas Y, et al. Effects of thyroxine therapy on cardiac function
in patients with subclinical hypothyroidism: Index of myocardial performance in the
evaluation of left ventricular function. International J Cardiol 2004; 95:135–143.
192. Turhan S, Tulunay C, Cin MO, et al. Effects of thyroxine therapy on right ventricular
systolic and diastolic function in patients with subclinical hypothyroidism: A study
by pulsed wave tissue Doppler imaging. J Clin Endocrinol Metab 2006; 91:3490–
3493.
193. Rodondi N, Newman AB, Vittinghoff E, et al. Subclinical hypothyroidism and the
risk of heart failure, other cardiovascular events, and death. Arch Intern Med 2005;
165:2460–2466.
194. Tunbridge WM, Evered DC, Hall R, et al. Lipid profiles and cardiovascular disease
in the Whickham area with particular reference to thyroid failure. Clin Endocrinol
(Oxf) 1977; 7:495–508.
195. Althaus BU, Staub JJ, Ryff-De Leche A, et al. LDL/HDL-changes in subclinical
hypothyroidism: Possible link factors for coronary heart disease. Clin Endocrinol
1988; 28:157–163.
196. Caron P, Calazel C, Parra HJ, et al. Decreased HDL cholesterol in subclinical
hypothyroidism: The effect of L-thyroxine therapy. Clin Endocrinol 1990; 33:519–
523.
197. Tanis BC, Westendorp GJ, Smelt AHM. Effect of thyroid substitution on hyperc-
holesterolaemia in patients with subclinical hypothyroidism: A reanalysis of inter-
vention studies. Clin Endocrinol 1996; 44:643–649.
198. Danese MD, Ladenson PW, Meinert CL, et al. Effect of thyroxine therapy on serum
lipoproteins in patients with mild thyroid failure: A quantitative review of the liter-
ature. J Clin Endocrinol Metab 2000; 85:2993–3001.
190 McDermott and Ridgway

199. Caraccio N, Ferrannini E, Monzani F. Lipoprotein profile in subclinical hypothy-


roidism: Response to levothyroxine replacement, a randomized placebo-controlled
study. J Clin Endocrinol Metab 2002; 87:1533–1538.
200. Monzani F, Caraccio N, Kozakowa M, et al. Effect of levothyroxine replacement on
lipid profile and intima-media thickness in subclinical hypothyroidism: A double-
blind, placebo-controlled study. J Clin Endocrinol Metab 2004; 89:2099–2106.
201. Walsh JP, Bremner AP, Bulsara MK, et al. Thyroid dysfunction and serum lipids: A
community based study. Clin Endocrinol (Oxf) 2005; 63:670–675.
202. Iqbal A, Jorde R, Figenschau Y. Serum lipid levels in relation to serum thyroid-
stimulating hormone and the effect of thyroxine treatment on serum lipid levels in
subjects with subclinical hypothyroidism: The Tromso Study. J Intern Med 2006;
260:53–61.
203. Ito M, Arishima T, Kudo T, et al. Effect of levo-thyroxine replacement on non-high-
density lipoprotein cholesterol in hypothyroid patients. J Clin Endocrinol Metab.
2007; 92:608–611.
204. Ito M, Takamatsu J, Sasaki I, et al. Disturbed metabolism of remnant lipoproteins in
patients with subclinical hypothyroidism. Am J Med 2004; 117:696–699.
205. Pearce EN, Wilson PW, Yang Q, et al. Thyroid function and lipid subparticle sizes
in patients with short-term hypothyroidism and a population based cohort. J Clin
Endocrinol Metab 2008; 93:888–894.
206. Michalopoulou G, Alevizaki M, Piperingos G, et al. High serum cholesterol levels
in persons with “high normal” TSH levels: Should one extend the definition of
subclinical hypothyroidism. Eur J Endocrinol 1998; 138(2):141–145.
207. Bindels AJ, Westendorp RG, Frolich M, et al. The prevalence of subclinical hypothy-
roidism at different total plasma cholesterol levels in middle aged men and women:
A need for case-finding? Clin Endocrinol 1999; 50:217–220.
208. Bakker SJL, Ter Matten JC, Popp-Snijders C, et al. The relationship between thy-
rotropin and low density lipoprotein cholesterol is modified by insulin sensitivity in
healthy euthyroid subjects. J Clin Endocrinol Metab 2001; 86:1206–1211.
209. Luboshitzky R, Aviv A, Herer P, et al. Risk factors for cardiovascular disease in
women with subclinical hypothyroidism. Thyroid 2002; 12:421–425.
210. Kvetny J, Heldgaard PE, Bladbjerg EM, et al. Subclinical hypothyroidism is
associated with a low grade inflammation, increased triglyceride levels and pre-
dicts cardiovascular disease in males below 50 years. Clin Endocrinol 2004; 61:
232–238.
211. Roos A, Bakker SJL, Links TP, et al. Thyroid function is associated with components
of the metabolic syndrome in euthyroid subjects. J Clin Endocrinol Metab 2007;
92:491–496.
212. Gumieniak O, Perlstein TS, Hopkins PN, et al. Thyroid function and blood pressure
regulation in euthyroid subjects. J Clin Endocrinol Metab 2004; 89:3455–3461.
213. Asvold BO, Bjoro T, Nilsen TIL, et al. Association between blood pressure and
serum thyroid-stimulating hormone concentration within the reference range: A
population-based study. J Clin Endocrinol Metab 2007; 92:841–845.
214. Waterhouse DF, McLaughlin AM, Walsh CD, et al. An examination of the relation-
ship between normal range thyrotropin and cardiovascular risk parameters: A study
in healthy women. Thyroid. 2007; 17;243–248.
215. Mariotti S, Cambuli VM. Cardiovascular risk in elderly hypothyroid patients. Thy-
roid 2007; 17:1067–1073.
Hypothyroidism 191

216. Walsh JP, Bremner AP, Bulsara MK, et al. Subclinical thyroid dysfunction and blood
pressure: A community based study. Clin Endocrinol (Oxf) 2006; 65:486–491.
217. Knudsen N, Laurberg P, Rasmussen LB, et al. Small differences in thyroid func-
tion may be important for body mass index and the occurrence of obesity in the
population. J Clin Endocrinol Metab 2005; 90:4019–4024.
218. Nyrnes A, Jorde R, Sundsfjord J. Serum TSH is positively associated with BMI. Int
J Obes (Lond) 2006; 30:100–105.
219. Diekman T, Lansberg PJ, Kastelein JJ, et al. Prevalence and correction of hypothy-
roidism in a large cohort of patients referred for dyslipidemia. Arch Intern Med
1995; 155:1490–1495.
220. Nagasaki T, Inaba M, Kumeda Y, et al. Increased pulse wave velocity in subclinical
hypothyroidism. J Clin Endocrinol Metab 2006; 91:154–158.
221. Taddei S, Carracio N, Virdis A, et al. Impaired endothelium-dependent vasodilation
in subclinical hypothyroidism: Beneficial effect of levothyroxine therapy. J Clin
Endocrinol Metab 2003; 88:3731–3737.
222. Cikim AS, Oflaz H, Ozbey N, et al. Evaluation of endothelial function in subclinical
hypothyroidism and subclinical hyperthyroidism. Thyroid 2004; 14:605–609.
223. Owen PJD, Rajiv C, Vinereanu D, et al. Subclinical hypothyroidism, arterial stiffness,
and myocardial reserve. J Clin Endocrinol Metab 2006; 91:2126–2132.
224. Hak AE, Pols HA, Visser TJ, et al. Subclinical hypothyroidism is an independent
risk factor for atherosclerosis and myocardial infarction in elderly women: The
Rotterdam Study. Ann Intern Med 2000; 132:270–278.
225. Mya MM, Aronow WS. Subclinical hypothyroidism is associated with coronary
artery disease in older persons. J Gerontol A Biol Sci Med Sci 2002; 57:M658–
M659.
226. Imaizumi M, Akahoshi M, Ichimaru S, et al. Risk for ischemic heart disease and
all-cause mortality in subclinical hypothyroidism. J Clin Endocrinol Metab 2004;
89:3365–3370.
227. Walsh JP, Bremner AP, Bulsara MK, et al. Subclinical thyroid dysfunction as a risk
factor for cardiovascular disease. Arch Intern Med 2005; 165:2467–2472.
228. Rodondi N, Aujesky D, Vittinghoff E, et al. Subclinical hypothyroidism and the risk
of coronary heart disease: A meta-analysis. Am J Med 2006; 119:541–551.
229. Rodondi N, Newman AB, Cappola AR, et al. Association between subclinical
hypothyroidism and coronary heart disease in 2006. Int J Cardiol 2007; Nov 19
[epub ahead of print].
230. Lindeman RD, Romero LJ, Schade DS, et al. Impact of subclinical hypothyroidism
on serum total homocysteine concentrations, the prevalence of coronary heart disease
(CHD), and CHD risk factors in the New Mexico Elder Health Survey. Thyroid 2003;
13:595–600.
231. Cappola AR, Fried LP, Arnold AM, et al. Thyroid status, cardiovascular risk, and
mortality in older adults. JAMA 2006; 295:1033–1041.
232. Gussekloo J, van Exel E, de Craen AJ, et al. Thyroid status, disability and cognitive
function, and survival in old age. JAMA 2004; 292:2651–2654.
233. Perk M, O’Neill BJ. The effect of thyroid hormone therapy on angiographic coronary
artery disease progression. Can J Cardiol 1997; 13:273–276.
234. Flynn RW, MacDonald TM, Jung RT, et al. Mortality and vascular outcomes in
patients treated for thyroid dysfunction. J Clin Endocrinol Metab 2006; 91:2159–
2164.
192 McDermott and Ridgway

235. Monzani F, Del Guerra P, Caraccio N, et al. Subclinical hypothyroidism: Neurobe-


havioral features and beneficial effect of L-thyroxine treatment. Clin Invest 1993;
71:367–371.
236. Tappy L, Randin JP, Schwed P, et al. Prevalence of thyroid disorders in psychogeri-
atric inpatients. A possible relationship of hypothyroidism with neurotic depression
but not dementia. J Am Geriatr Soc 1987; 35:526–531.
237. Joffe RT, Levitt AJ. Major depression and subclinical (grade 2) hypothyroidism.
Psychoneuroendocrinology 1992; 17:215–221.
238. Haggerty JJ Jr, Stern RA, Mason GA, et al. Subclinical hypothyroidism: A modifiable
risk factor for depression? Am J Psychiatry 1993; 150:508–510.
239. Manciet G, Dartigues JF, Decamps A, et al. The PAQUID survey and correlates
of subclinical hypothyroidism in elderly community residents in the southwest of
France. Age Aging 1995; 24(3):235–241.
240. Jaeschke R, Guyatt G, Gerstein H, et al. Does treatment with L-thyroxine influence
health status in middle-aged and older adults with subclinical hypothyroidism?
J Gen Intern Med 1996; 11:744–749.
241. Baldini IM, Vita A, Maura MC, et al. Psychological and cognitive features in subclin-
ical hypothyroidism. Prog Neuropsychopharmacol Biol Psychiatry 1997; 21(6):925–
935.
242. Ganguli M, Burmeister LA, Seaberg EC, et al. Association between dementia
and elevated TSH: A community-based study. Biol Psychiatry 1996; 40(8):714–
725.
243. Monzani F, Caraccio N, Siciliano G, et al. Clinical and biochemical features of
muscle dysfunction in subclinical hypothyroidism. J Clin Endocrinol Metab 1997;
82:3315–3318.
244. Monzani F, Caraccio N, Del Guerra P, et al. Neuromuscular symptoms and dys-
function in subclinical hypothyroid patients: Beneficial effect of L-T4 replacement
therapy. Clin Endocrinol 1999; 51:237–242.
245. Roberts LM, Pattison H, Roalfe A, et al. Is subclinical thyroid dysfunction in the
elderly associated with depression or cognitive dysfunction? Ann Intern Med 2006;
145:573–581.
246. Jorde R, Waterloo K, Storhaug H, et al. Neuropsychological function and symptoms
in subjects with subclinical hypothyroidism and the effect of thyroxine treatment.
J Clin Endocrinol Metab 2006; 91:145–153.
247. Misiunas A, Ravera HN, Faraj G, et al. Peripheral neuropathy in subclinical hypothy-
roidism. Thyroid 1995; 5:283–286.
248. Goulis DG, Tsimpiris N, Delaroudis S, et al. Stapedial reflex: A biological index
found to be abnormal in clinical and subclinical hypothyroidism. Thyroid 1998;
8:583–587.
249. Tutuncu NB, Karatas M, Sozay S. Prolonged P300 latency in thyroid failure: A
paradox. P300 latency recovers later in mild hypothyroidism than in severe hypothy-
roidism. Thyroid 2004; 14:622–627.
250. Krausz Y, Freedman N, Lester H, et al. Regional cerebral blood flow in patients with
mild hypothyroidism. J Nucl Med 2004; 45:1712–1715.
251. Nicoloff JT, LoPresti JS. Myxedema coma. A form of decompensated hypothy-
roidism. Endocrinol Metab Clin NA 1993; 22:279–290.
252. Wartofsky L. Myxedema coma. In: Braverman LE, Utiger RD, eds. Werner and
Ingbar’s The Thyroid, 7th ed. Philadelphia, PA: Lippencott-Raven, 1996:871–877.
Hypothyroidism 193

253. Fliers E, Wiersinga WM. Myxedema coma. Rev Endo Metab Dis 2003; 4:137–141.
254. Sarlis NJ, Gourgiotis L. Thyroid emergencies. Rev Endo Metab Dis 2003; 4:129–36.
255. Yamamoto T, Fukuyama J, Fujiyoshi A. Factors associated with mortality of
myxedema coma: Report of eight cases and literature survey. Thyroid 1999; 9:1167–
1174.
256. Rodriguez I, Fluiters E, Perez-Mendez LF, et al. Severe mental impairment and poor
physiologic status are associated with mortality in myxedema coma. J Endocrinol
2004; 180:347–350.
257. Lum SM, Nicoloff JT, Spencer CA. Peripheral tissue mechanism for maintenance
of serum triiodothyronine values in a thyroxine-deficient state in man. J Clin Invest
1984; 73:570–575.
258. Croteau W, Davey JC, Galton VA, et al. Cloning of the mammalian type II iodothy-
ronine deiodinase. A selenoprotein differentially expressed and regulated in human
and rat brain and other tissues. J Clin Invest 1996; 98:405–417.
259. Salvatore D, Tu H, Harney JW, Larsen PR. Type 2 iodothyronine deiodinase is highly
expressed in human thyroid. J Clin Invest 1996; 98:962–968.
260. Ferretti E, Persani L, Jaffrain-Rea M-L, et al. Evaluation of the adequacy of
levothyroxine replacement therapy in patients with central hypothyroidism. J Clin
Endocrinol Metab 1999; 84:924–929.
261. Faglia G, Bitensky L, Pinchera A, et al. Thyrotropin secretion in patients with
central hypothyroidism: Evidence for reduced biological activity of immunoreactive
thyrotropin. J Clin Endocrinol Metab 1979; 48:989–998.
262. Beck-Peccoz P, Amr S, Menezes-Ferreira MM, et al. Decreased receptor binding of
biologically inactive thyrotropin in central hypothyroidism. Effect of treatment with
thyrotropin-releasing hormone. N Engl J Med 1985; 312:1085–1090.
263. Uy HL, Reasner CA, Samuels MH. Pattern of recovery of hypothalamic-pituitary-
thyroid axis following radioactive iodine therapy in patients with Graves’ disease.
Am J Med 1995; 99:173–179.
264. Sawers JSA, Toft AD, Irvine WJ, et al. Transient hypothyroidism after iodine-131
treatment of thyrotoxicosis. J Clin Endocrinol Metab 1980; 50:226–229.
265. Wartofsky L, Burman KD. Alterations in thyroid function in patients with systemic
illness: The “euthyroid sick syndrome”. Endocr Rev 1982; 3:164–217.
266. Adler SM, Wartofsky L. The nonthyroidal illness syndrome. Endocrinol Metab Clin
NA 2007; 36:657–672.
267. Hangaard J, Andersen M, Grodum E, et al. Pulsatile thyrotropin secretion in patients
with Addison’s disease during variable glucocorticoid therapy. J Clin Endocrinol
Metab 1996; 81:2502–2507.
268. Tonacchera M, Perri A, De Marco G, et al. Low prevalence of thyrotropin receptor
mutations in a large series of subjects with sporadic and familial non-autoimmune
subclinical hypothyroidism. J Clin Endocrinol Metab 2004; 89:5787–5793.
269. Steel NR, Weightman DR, Taylor JJ, et al. Blocking activity to action of thyroid
stimulating hormone in serum from patients with primary hypothyroidism. Br Med
J 1984; 288:1559–1562.
270. Konishi J, Iida Y, Kasagi K, et al. Primary myxedema with thyrotrophin-binding
inhibitor immunoglobulins. Ann Intern Med 1985; 103:26–31.
271. Takasu N, Yamada T, Takasu M, et al. Disappearance of thyrotropin-blocking anti-
bodies and spontaneous recovery from hypothyroidism in autoimmune thyroiditis.
N Engl J Med 1992; 326:513–518.
194 McDermott and Ridgway

272. Arbelle JE, Porath A. Practice guidelines for the detection and management of thyroid
dysfunction. A comparative review of the recommendations. Clin Endocrinol (Oxf)
1999; 51:11–18.
273. Danese MD, Powe NR, Sawin CT, et al. Screening for mild thyroid failure at the
periodic health examination. JAMA 1996; 276:285–292.
274. Am Col Phys. Screening for thyroid disease. Ann Intern Med 1998; 129:141–143.
275. Helfand M. Screening for subclinical thyroid dysfunction in nonpregnant adults: A
summary of the evidence for the US Preventive Services Task Force. Ann Intern
Med 2004; 140:128–141.
276. Am Col Phys. Screening for thyroid disease: Recommendation statement. Ann Intern
Med 2004; 140:125–127.
277. Committee on Medicare Coverage of Routine Thyroid Screening: Medicare coverage
of routine screening for thyroid dysfunction. Stone MB and Wallace RB (editors),
2003.
278. Ladenson PW, Singer PA, Ain KB, et al. American Thyroid Association guidelines
for detection of thyroid dysfunction. Arch Intern Med 2000; 160:1573–1575.
279. Gharib H, Tuttle RM, Baskin J, et al. Consensus statement: Subclinical thyroid
dysfunction: A joint statement on management from the American Association of
Clinical Endocrinologists, the American Thyroid Association, and the Endocrine
Society. J Clin Endocrinol Metab 2005; 90:581–585.
280. Abalovich M, Amino N, Barbour LA, et al. Management of thyroid dysfunction
during pregnancy and postpartum; an Endocrine Society Practice Guideline. J Clin
Endocrinol Metab 2007; 92(8 suppl.):S1–S47.
281. Vaidya B, Anthony S, Bilous M, et al. Detection of thyroid dysfunction in early
pregnancy: Universal screening or targeted high-risk case finding? J Clin Endocrinol
Metab 2007; 92:203–207.
282. Murray GR. Note on the treatment of myxoedema by hypodermic injections of an
extract of the thyroid gland of a sheep. Br Med J 1891; 2:796–797.
283. MacKenzie HWG. A case of myxoedema treated with great benefit by feeding with
fresh thyroid glands. Br Med J 1892; 2:940–941.
284. Toft AD. Thyroxine therapy. N Engl J Med 1994; 331:174–181.
285. Oppenheimer JH, Braverman LE, Toft AD, et al. A therapeutic controversy. Thyroid
hormone treatment: When and what? J Clin Endocrinol Metab 1995; 80:2873–2882.
286. Helfand M, Crapo LM. Monitoring therapy in patients taking levothyroxine. Ann
Intern Med 1990; 113:450–454.
287. Mandel SJ, Brent GA, Larsen PR. Levothyroxine therapy in patients with thyroid
disease. Ann Intern Med 1993; 119:492–502.
288. Rees-Jones RW, Larsen PR. Triiodothyronine and thyroxine content of desiccated
thyroid tablets. Metabolism 1977; 26:1213–1218.
289. Rees-Jones RW, Rolla AR, Larsen PR. Hormonal content of thyroid replacement
preparations. JAMA 1980; 243:549–550.
290. LeBoff MS, Kaplan MM, Silva JE, et al. Bioavailability of thyroid hormones from
oral replacement preparations. Metabolism 1982; 31:900–905.
291. Singer PA, Cooper DS, Levy EG, Treatment guidelines for patients with hyperthy-
roidism and hypothyroidism. JAMA 1995; 273:808–812.
292. Jonklass J, Davidson B, Bhagat S, et al. Triiodothyronine levels in athyroeotic
individuals during levothyroxine therapy. JAMA 2008; 299:817–819.
Hypothyroidism 195

293. Stoffer SS, Szpunar WE. Potency of brand name and generic levothyroxine. JAMA
1980; 244:1704–1705.
294. Sawin CT, Surks MI, London M, et al. Oral thyroxine: Variation in biologic action
and tablet content. Ann Intern Med 1984; 100:641–645.
295. Hennessey JV, Burman KD, Wartofsky L. The equivalency of two L-thyroxine
preparations. Ann Intern Med 1985; 102:770–773.
296. Dong BJ, Young VR, Rapoport B. The nonequivalence of levothyroxine products.
Drug Intell Clin Pharm 1986; 20:77–78.
297. Dong BJ, Hauck WW, Gambertoglio JG, et al. Bioequivalence of generic and brand-
name levothyroxine products in the treatment of hypothyroidism. JAMA 1997;
277:1205–1213.
298. Rennie D. Thyroid storm. JAMA 1997; 277:1238–1243.
299. Spigelman MK. Bioequivalence of levothyroxine preparations for treatment of
hypothyroidism. JAMA 1997; 277:1199.
300. Eckert CH. Bioequivalence of levothyroxine preparations: Industry sponsorship and
academic freedom. JAMA 1997; 277:1200–1201.
301. Hays MT. Thyroid hormone and the gut. Endocr Res 1988; 14:203–224.
302. Hays MT. Localization of human thyroxine absorption. Thyroid 1991; 1:241–248.
303. Fish LH, Schwartz HL, Cavanaugh J, et al. Replacement dose, metabolism, and
bioavailability of levothyroxine in the treatment of hypothyroidism. N Engl J Med
1987; 316:764–770.
304. Wenzel KW, Kirschsieper HE. Aspects of the absorption of oral l-thyroxine in normal
man. Metabolism 1977; 26:1–8.
305. Bolk N, Visser TJ, Kalsbeek A, et al. Effects of evening vs morning thyroxine
ingestion on serum thyroid hormone profiles in hypothyroid patients. Clin Endocrinol
(Oxf) 2007; 66:43–48.
306. Greenspan SL, Klibanski A, Schoenfeld D, et al. Pulsatile secretion of thyrotropin
in man. J Clin Endocrinol Metab 1986; 63:661–668.
307. Ain KB, Pucino F, Shiver TM, et al. Thyroid hormone levels affected by time of
blood sampling in thyroxine-treated patients. Thyroid 1993; 3:81–85.
308. Wennlund A. Variation in serum levels of T3, T4, FT4 and TSH during thyroxine
replacement therapy. Acta Endocrinol (Copenh) 1999; 113:47–49.
309. Hennessey JV, Evaul JE, Tseng Y-C, et al. L-thyroxine dosage: A reevaluation of
therapy with contemporary preparations. Ann Intern Med 1986; 105:11–15.
310. Roos A, Linn-Rasker SP, van Domburg RT, et al. The starting dose of levothyroxine
in primary hypothyroidism treatment: A prospective, randomized, double-blind trial.
Arch Intern Med 2005; 165:1714–1720.
311. Santini F, Pinchera A, Marsili A, et al. Lean body mass is a major determinant of
levothyroxine dosage in the treatment of thyroid diseases. J Clin Endocrinol Metab
2005; 90:124–127.
312. Kabadi UM. Optimal daily levothyroxine dose in primary hypothyroidism. Arch
Intern Med 1989; 149:2209–2212.
313. McDermott MT, Haugen BR, Lezotte DC, et al. Management practices among
primary care physicians and thyroid specialists in the care of hypothyroid patients.
Thyroid 2001; 11:757–764.
314. Carr D, McLeod DT, Parry G, et al. Fine adjustment of thyroxine replacement dosage:
Comparison of the thyrotrophin releasing hormone test using a sensitive thyrotrophin
196 McDermott and Ridgway

assay with measurement of free thyroid hormones and clinical assessment. Clin
Endocrinol 1988; 28:325–333.
315. McDermott MT, Ridgway ECR. Subclinical hypothyroidism is mild thyroid failure
and should be treated. J Clin Endocrinol Metab 2001; 86:4585–4590.
316. Chu JW, Crapo LM. The treatment of subclinical hypothyroidism is seldom neces-
sary. J Clin Endocrinol Metab 2001; 86:4591–4599.
317. Villar HC, Saconato H, Valente O, et al. Thyroid hormone replacement for subclinical
hypothyroidism. Cochrane Database Syst Rev 2007; 18(3):CD003419.
318. Tibaldi J, Barzel US. Thyroxine supplementation. Method for the prevention of
clinical hypothyroidism. Am J Med 1985; 79:241–244.
319. Cooper DS. Thyroid disease in the oldest old: The exception to the rule. JAMA 2004;
292:2651–2654.
320. Tallstedt L, Lundell G, Blomgren H. Does early administration of thyroxine reduce
the development of Graves’ ophthalmopathy after radioiodine treatment? Eur J
Endocrinol 1994; 130:494–497.
321. Slawik M, Klawitter B, Meiser E, et al. Thyroid hormone replacement for central
hypothyroidism: A randomized controlled trial comparing two doses of thyroxine
(T4) with a combination of T4 and triiodothyronine. J Clin Endocrinol Metab 2007;
92:4115–4122.
322. Burmeister LA, Goumaz MO, Mariash CN, et al. Levothyroxine dose requirements
for thyrotropin suppression in the treatment of differentiated thyroid cancer. J Clin
Endocrinol Metab 1992; 75:344–350.
323. Wartofsky L. Use of sensitive TSH assay to determine optimal thyroid hormone
therapy and avoid osteoporosis. Annual Rev Med 1991; 42:341–345.
324. Spencer CA, LoPresti JS, Nicoloff JT, et al. Multiphasic thyrotropin responses to
thyroid hormone administration in man. J Clin Endocrinol Metab 1995; 80:854–
859.
325. Pujol P, Daures J-P, Nsakala N, et al. Degree of thyrotropin suppression as a prog-
nostic determinant in differentiated thyroid cancer. J Clin Endocrinol Metab 1996;
81:4318–4323.
326. Cooper DS, Doherty GM, Haugen BR, et al. The American Thyroid Association
Taskforce. Thyroid 2006; 16:109–142.
327. Othman S, Phillips DIW, Parkes AB, et al. A long-term follow-up of postpartum
thyroiditis. Clin Endocrinol 1990; 32:559–564.
328. Burrow GN. Thyroid dysfunction in the recently pregnant: Postpartum thyroiditis.
Thyroid 1994; 4:363–365.
329. Lazarus JH. Clinical manifestations of postpartum thyroid disease. Thyroid 1999;
9:685–689.
330. Wolff J, Chaikoff IL. Plasma inorganic iodide as a homeostatic regulator of thyroid
function. J Biol Chem 1948; 174:555–564.
331. Raben MS. The paradoxical effects of thiocyanate and of thyrotropin on the organic
binding of iodine by the thyroid in the presence of large amounts of iodide.
Endocrinology 1949; 45:296–304.
332. Holvey DN, Goodner CJ, Nicoloff JT, et al. Treatment of myxedema coma with
intravenous thyroxine. Arch Intern Med 1964; 113:89–96.
333. Ridgway EC, McCammon JA, Benotti J, et al. Acute metabolic responses in
myxedema to large doses of intravenous L-thyroxine. Ann Intern Med 1972; 77:549–
555.
Hypothyroidism 197

334. Chernow B, Burman KD, Johnson DL, et al. T3 may be a better agent than T4 in
the critically ill hypothyroid patient: Evaluation of transport across the blood-brain
barrier in a primate model. Crit Care Med 1983; 11:99–104.
335. Ladenson PW, Goldenheim PD, Ridgway EC. Rapid pituitary and peripheral tissue
responses to intravenous L-triiodothyronine in hypothyroidism. J Clin Endocrinol
Metab 1983; 56:1252–1259.
336. Saravanan P, Chau WF, Roberts N, et al. Psychological well-being in patients on
“adequate” doses of l-thyroxine: Results of a large, controlled community-based
questionnaire study. Clin Endocrinol (Oxf) 2002; 57:577–585.
337. Wekking EM, Appelhof BC, Fliers E, et al. Cognitive functioning and well-being
in euthyroid patients on thyroxine replacement therapy for primary hypothyroidism.
Eur J Endocrinol 2005; 153:747–753.
338. Saravanan P, Visser TJ, Dayan CM. Psychological well-being correlates with free
thyroxine but not free 3,5,3’- triiodothyronine levels in patients on thyroid hormone
replacement. J Clin Endocrinol Metab 2006; 91:3389–3393.
339. Samuels MH, Schuff KG, Carlson NE, et al. Health status, psychological symptoms,
mood, and cognition in L-thyroxine-treated hypothyroid subjects. Thyroid 2007;
17;249–258.
340. Bunevicius R, Kazanavicius G, Zalinkevicius R, et al. Effects of thyroxine as com-
pared with thyroxine plus triiodothyronine in patients with hypothyroidism. N Engl
J Med 1999; 340:424–429.
341. Toft AD. Thyroid hormone replacement—one hormone or two? N Engl J Med 1999;
340:469–470.
342. Walsh JP, Shiels L, Lim EM, et al. Combined thyroxine/liothyronine treatment does
not improve well-being, quality of life, or cognitive function compared to thyroxine
alone: A randomized controlled trial in patients with primary hypothyroidism. J Clin
Endocrinol Metab 2003; 88:4543–4550.
343. Sawka AM, Gerstein HC, Marriott MJ, et al. Does a combination regimen of thy-
roxine (T4) and 3,5,3’-triiodothyronine improve depressive symptoms better than
T4 alone in patients with hypothyroidism? Results of a double-blind, randomized,
controlled trial. J Clin Endocrinol Metab 2003; 88:4551–4555.
344. Clyde PW, Harari AE, Getka EJ, et al. Combined levothyroxine plus liothyronine
compared with levothyroxine alone in primary hypothyroidism: A randomized con-
trolled trial. JAMA 2003; 290:2952–2958.
345. Siegmund W, Spieker K, Weike AI, et al. Replacement therapy with levothyroxine
plus triiodothyronine (bioavailable molar ratio 14:1) is not superior to thyroxine
alone to improve well-being and cognitive performance in hypothyroidism. Clin
Endocrinol (Oxf) 2004; 60:750–757.
346. Escobar-Morreale HF, Botella-Carretero JI, Gomez-Bueno M, et al. Thyroid hor-
mone replacement in primary hypothyroidism: A randomized trial comparing L-
thyroxine plus liothyronine with L-thyroxine alone. Ann Intern Med 2005; 142:412–
424.
347. Saravanan P, Simmons DJ, Greenwood R, et al. Partial substitution of thyroxine
(T4) with triiodothyronine in patients on T4 replacement therapy: Results of a
large community-based randomized controlled trial. J Clin Endocrinol Metab 2005;
90:805–812.
348. Rodriguez T, Lavis VR, Meininger JC, et al. Substitution of liothyronine at a 1:5
ratio for a portion of levothyroxine: Effect on fatigue, symptoms of depression,
198 McDermott and Ridgway

and working memory versus treatment with levothyroxine alone. Endocrine Practice
2005; 11:223–233.
349. Appelhof BC, Fliers E, Wekking EM, et al. Combined therapy with levothyroxine
and liothyronine in two ratios, compared with levothyroxine monotherapy in pri-
mary hypothyroidism: A double-blind, randomized, controlled clinical trial. J Clin
Endocrinol Metab 2005; 90:2666–2674.
350. Regalbuto C, Maiorana R, Alagona C, et al. Effects of either LT4 monotherapy or
LT4/LT3 combined therapy in patients totally thyroidectomized for thyroid cancer.
Thyroid 2007; 17;323–3231.
351. Escobar-Morreale HF, Botella-Carretero JI, Escobar del Rey F, et al. Review: Treat-
ment of hypothyroidism with combinations of levothyroxine plus liothyronine.
J Clin Endocrinol Metab 2005; 90:4946–4954.
352. Grozinsky-Glasberg S, Fraser A, Nahshoni E, et al. Thyroxine-Triiodothyronine
combination therapy versus thyroxine monotherapy for clinical hypothyroidism:
Meta-analysis of randomized controlled trials. J Clin Endocrinol Metab 2006;
91:2592–2599.
353. Mandel SJ, Larsen PR, Seely EW, et al. Increased need for thyroxine during preg-
nancy in women with primary hypothyroidism. N Engl J Med 1990; 323:91–96.
354. Kaplan MM. Monitoring thyroxine treatment during pregnancy. Thyroid 1992;
2:147–152.
355. Alexander EK, Marqusee E, Lawrence J, et al. Timing and magnitude of increases
in levothyroxine requirements during pregnancy in women with hypothyroidism.
N Engl J Med 2004; 351:241–249.
356. Rotondi M, Mazziotti G, Sorvillo F, et al. Effects of increased thyroxine dosage
pre-conception on thyroid function during early pregnancy. Eur J Endocrinol 2004;
151:695–700.
357. Witzum JL, Jacobs LS, Schonfeld G. Thyroid hormone and thyrotropin levels in
patients placed on colestipol hydrochloride. J Clin Endocrinol Metab 1978; 46:838–
840.
358. Campbell NR, Hasinoff BB, Stalts H, et al. Ferrous sulfate reduces thyroxine efficacy
in patients with hypothyroidism. Ann Intern Med 1992; 117:1010–1013.
359. Liel Y, Sperber AD, Shany S. Nonspecific intestinal adsorption of levothyroxine by
aluminum hydroxide. Am J Med 1994; 97:363–365.
360. Schneyer CR. Calcium carbonate and reduction of levothyroxine efficacy. JAMA
1998; 279:750.
361. Chiu AC, Sherman SI. Effects of pharmacological fiber supplements on levothyrox-
ine absorption. Thyroid 1998; 8:667–671.
362. Bell DS, Ovalle F. Use of soy protein supplement and resultant need for increased
dose of levothyroxine. Endocr Pract 2001; 7:193–194.
363. Siraj ES, Gupta MK, Reddy SS. Raloxifene causing malabsorption of levothyroxine.
Arch Intern Med 2003; 163:1367–1370.
364. John-Kalarickal J, Pearlman G, Carlson HE. New medications which decrease
levothyroxine absorption. Thyroid 2007; 17:763–765.
365. Diskin CJ, Stokes TJ, Dansby LM, et al. The effect of phosphate binders upon TSH
and L-thyroxine dose in patients on thyroid replacement. Int Urol Nephrol 2007;
39:599–602.
366. Centanni M, Gargano L, Canettieri G, et al. Thyroxine in goiter, helicobacter pylori
infection, and chronic gastritis. N Engl J Med 2006; 354:1787–1795.
Hypothyroidism 199

367. Checchi S, Montanaro A, Pasqui L, et al. L-thyroxine requirement in patients with


autoimmune hypothyroidism and parietal cell antibodies. J Clin Endocrinol Metab
2008; 93:465–469.
368. McDermott JH, Coss A, Walsh CH. Celiac disease presenting as resistant hypothy-
roidism. Thyroid 2005; 15:386–388.
369. Munoz-Torres, Varsavsky M, Alonso G. Lactose intolerance revealed by severe
resistance to treatment with levothyroxine. Thyroid 2006; 16:1171–1176.
370. Junglee NA, Scanlon MF, Rees DA. Increasing thyroxine requirements in
primary hypothyroidism: Don’t forget the urinalysis. J Postgrad Med 2006; 52:201–
203.
371. Blackshear JL, Schultz AL, Napier JS, et al. Thyroxine replacement requirements
in hypothyroid patients receiving phenytoin. Ann Intern Med 1983; 99:341–342.
372. Arafah BM. Increased need for thyroxine in women with hypothyroidism during
estrogen therapy. N Engl J Med 2001; 344:1743–1749.
373. McCowen KC, Garber JR, Spark R. Elevated serum thyrotropin in thyroxine
treated patients with hypothyroidism given sertraline. N Engl J Med 1997; 337:1010–
1011.
374. Ain KB, Refetoff S, Fein HG, et al. Pseudomalabsorption of levothyroxine. JAMA
1991; 266:2118–2120.
375. Livadariu E, Valdes-Socin H, Burlacu MC, et al. Pseudomalabsorption of thyroid
hormones: Case report and review of the literature. Ann Endocrinol (Paris) 2007;
68:460–463.
376. Grebe S, Cooke RR, Ford HC, et al. Treatment of hypothyroidism with once weekly
thyroxine. J Clin Endocrinol Metab 1997; 82:870–875.
377. Sawin CT, Herman T, Moltich ME, et al. Aging and the thyroid: Decreased require-
ment for thyroid hormone in older hypothyroid patients. Am J Med 1983; 75:206–
209.
378. Robuschi G, Safran M, Braverman LE, et al. Hypothyroidism in the elderly. Endocr
Rev 1987; 8:142–153.
379. Arafah BM. Decreased levothyroxine requirement in women with hypothyroidism
during androgen therapy for breast cancer. Ann Intern Med 1994; 121:247–251.
380. Vigersky RA, Filmore-Nassar A, Glass AR. Thyrotropin suppression by metformin.
J Clin Endocrinol Metab 2006; 91:225–227.
381. Weinberg AD, Brennan MD, Gorman CA, et al. Outcome of anesthesia and surgery
in hypothyroid patients. Arch Intern Med 1983; 143:893–897.
382. Drucker DJ, Burrows GN. Cardiovascular surgery in the hypothyroid patient. Arch
Intern Med 1985; 145:1585–1587.
383. Becker C. Hypothyroidism and atherosclerotic heart disease: Pathogenesis, medical
management, and role of coronary artery bypass surgery. Endocr Rev 1985; 6:432–
440.
384. Ladenson PW, Levin AA, Ridgway EC, et al. Complications of surgery in hypothy-
roid patients. Am J Med 1984; 77:261–266.
385. Sherman SI, Ladenson PW. Percutaneous transluminal coronary angioplasty in
hypothyroidism. Am J Med 1991; 90:367–370.
386. Gwinup G, Poucher R. A controlled study of thyroid analogs in the therapy of
obesity. Am J Med Sci 1967; 254:416–420.
387. Rivilin RS. Drug therapy: Therapy of obesity with hormones. N Engl J Med 1975;
292:26–29.
200 McDermott and Ridgway

388. Sorbris R, Petersson BG, Nilsson-Ehle P. The variability of weight reduction during
fasting: Predictive value of thyroid hormone measurements. Int J Obes 1982; 6:101–
111.
389. Stokholm KH, Hansen MS. Lowering of serum total T3 during a conventional
slimming regime. Int J Obes 1983; 7:195–199.
390. Welle SL, Amatruda JM, Forbes GB, et al. Resting metabolic rates of obese women
after rapid weight loss. J Clin Endocrinol Metab 1984; 59:41–44.
391. Cavallo E, Armellini F, Zamboni M, et al. Resting metabolic rate, body composition,
and thyroid hormones. Short term effects of very low calorie diet. Horm Metab Res
1990; 22:632–635.
392. Moore R, Grant AM, Howard AN, et al. Treatment of obesity with triiodothyronine
and a very-low-calorie liquid-formula diet. Lancet 1980; 1:223–226.
393. Wilson JH, Lamberts SW. The effect of triiodothyronine on weight loss and nitrogen
balance of obese patients on a very-low-calorie liquid-formula diet. Int J Obes 1981;
5:279–282.
394. Wolman SI, Sheppard H, Fern M, et al. The effect of triiodothyronine (T3) on protein
turnover and metabolic rate. Int J Obes 1985; 9:459–463.
395. Pasquali R, Baraldi G, Biso P, et al. Effect of “physiological” doses of triiodothy-
ronine replacement on the hormonal and metabolic adaptation to short-term semi-
starvation and to low-calorie diet in obese patients. Clin Endocrinol (Oxf) 1984;
21:357–367.
396. Rosen R, Abraham G, Falcou R, et al. Effects of a “physiological” dose of tri-
iodothyronine on obese subjects during a protein-sparing diet. Int J Obes 1986;
10:303–312.
397. Joffe RT, Sokolov S, Singer W. Thyroid hormone treatment of depression. Thyroid
1995; 5:235–239.
398. Prange AJ Jr. Novel uses of thyroid hormones in patients with affective disorders.
Thyroid 1996; 6:537–543.
399. Lojko D, Rybakowski JK. L-thyroxine augmentation of serotonergic antidepressants
in female patients with refractory depression. J Affect Disord 2007; 103:253–256.
400. Cooper-Kazaz R, Lerer B. Efficacy and safety of triiodothyronine supplementation
in patients with major depressive disorder treated with specific serotonin reuptake
inhibitors. Int J Neuropsychopharmacol 2007; 30:1–15.
401. Girdler SS, Pedersen CA, Light KC. Thyroid axis function during the menstrual
cycle in women with premenstrual syndrome. Psychoneuroendocrinology 1995;
20:395–403.
402. Casper RF, Patel-Christopher A, Powell AM. Thyrotropin and prolactin responses to
thyrotropin-releasing hormone in premenstrual syndrome. J Clin Endocrinol Metab
1989; 68:608–612.
403. Nikolai TF, Molligan GM, Gribble RK, et al. Thyroid function and treatment in
premenstrual syndrome. J Clin Endocrinol Metab 1990; 70:1108–1113.
404. Schmidt PJ, Grover GN, Roy-Byrne PP, et al. Thyroid function in women with
premenstrual syndrome. J Clin Endocrinol Metab 1993; 76:671–674.
405. Sawin CT, Geller A, Wolf PA, et al. Low serum thyrotropin concentrations as a risk
factor for atrial fibrillation in older persons. N Engl J Med 1994; 331:1249–1252.
406. Auer J, Scheiber P, Mische T, et al. Subclinical hyperthyroidism as a risk factor for
atrial fibrillation. Am Heart J 2001; 142:838–842.
Hypothyroidism 201

407. Bell GM, Sawers JSA, Forfar JC, et al. The effect of minor increments in plasma
thyroxine on heart rate and urinary sodium excretion. Clin Endocrinol 1983; 18:511–
516.
408. Biondi B, Fazio S, Carella C, et al. Cardiac effects of long-term thyrotropin-
suppressive therapy with levothyroxine. J Clin Endocrinol Metab 1993; 77:334–338.
409. Fazio S, Biondi B, Carella C, et al. Diastolic dysfunction in patients on thyroid-
stimulating hormone suppressive therapy with levothyroxine: Beneficial effect of
beta-blockade. J Clin Endocrinol Metab 1995; 80:2222–2226.
410. Biondi B, Fazio S, Cuocolo A, et al. Impaired cardiac reserve and exercise capacity
in patients receiving long-term thyrotropin suppressive therapy with levothyroxine.
J Clin Endocrinol Metab 1996; 81:4224–4228.
411. Stall GM, Harris S, Sokoll LJ, et al. Accelerated bone loss in hypothyroid patients
overtreated with L-thyroxine. Ann Intern Med 1990; 113:265–269.
412. Adlin EV, Maurer AH, Marks AD, et al. Bone mineral density in postmenopausal
women treated with L-thyroxine. Am J Med 1991; 90:360–366.
413. Schneider DL, Barrett-Connor EL, Morton DJ. Thyroid hormone use and bone
mineral density in elderly women. JAMA 1994; 271:1245–1249.
414. Faber J, Galloe AM. Changes in bone mass during prolonged subclinical hyper-
thyroidism due to L-thyroxine treatment: A meta-analysis. Eur J Endocrinol 1994;
130:350–356.
415. Greenspan SL, Greenspan FS. The effect of thyroid hormone on skeletal integrity.
Ann Intern Med 1999; 130:750–758.
416. Bauer DC. Ettinger B, Nevitt MC, et al. Study of Osteoporosis Fractures Study
Group. Ann Intern Med 2001; 134:561–568.
5
Thyroid Nodules and Multinodular
Goiter

Hossein Gharib
College of Medicine, Mayo Clinic, Rochester, Minnesota,
U.S.A.

INTRODUCTION
Nodular thyroid disease, the presence of single or multiple nodules within the thy-
roid gland, is a common clinical problem. Most clinicians—particularly primary
care physicians, pediatricians, internists, endocrinologists, and general surgeons—
should regularly evaluate patients with thyroid nodules and consequently must
make diagnostic and management decisions. Solitary nodules may be benign or
malignant; multinodular glands may be asymptomatic, toxic, nontoxic, benign, or
malignant. Each patient is evaluated for structural and functional abnormalities.
It is difficult to overstate the influence of recent technologic developments
on examination and treatment of thyroid nodules. The introduction of sensi-
tive thyrotropin [i.e., thyroid-stimulating hormone (TSH)] assays, the universal
application of fine-needle aspiration (FNA) biopsy, and the widespread use of
high-resolution ultrasonography (US) has markedly improved the management of
thyroid nodule. Furthermore, collected data on the efficacy of levothyroxine (T4 )
suppressive therapy have been inconsistent at best and disappointing at worst,
resulting in a decline in the routine use of T4 in the treatment of thyroid nodules.
Improved technology has not been without unexpected and sometimes undesirable
consequences. For example, increased sensitivity of current imaging equipment
has resulted in the frequent discovery of subclinical nodules in the thyroid gland,
creating difficult treatment decisions for both the clinician and the patient.

203
204 Gharib

Table 1 Classification of Diffuse Goiter


Sporadic (nonendemic)
Iodine deficiency (endemic)
Autoimmune
Neoplasms
Genetic
Goitrogens

This chapter on nodular thyroid disease includes a discussion on the classi-


fication and prevalence of thyroid nodules; the pathogenesis of thyroid nodules;
laboratory tests; a review on the uses and limitations of FNA biopsy, thyroid
scanning, and US; management of clinically solitary nodules and multinodular
glands; and the problem of thyroid incidentalomas. It also includes recommen-
dations from recent practice management guidelines published by the American
Association of Clinical Endocrinologists in collaboration with the Associazione
Medici Endocrinologi (1), by the American Thyroid Association (2), and by the
European Thyroid Association (3) in the past 2 years.

CLASSIFICATION AND PREVALENCE


Thyroid gland enlargement (goiter) is one of the most common endocrine prob-
lems in clinical practice. Goiter is classified as “diffuse” or “nodular”; it may be
either “toxic” or “nontoxic.” The common conditions considered in the differential
diagnosis of diffuse goiters are listed in Table 1. “Nontoxic goiter” refers to thy-
roid enlargement associated with normal serum levels of TSH and without clinical
hyperthyroidism. In some geographic regions, iodine deficiency or environmental
goitrogens affect more than 10% of the local population, causing “endemic goi-
ter.” Worldwide, endemic goiter probably represents the most common endocrine
disorder. In the United States and Great Britain, nonendemic, or sporadic, goiter
affects more than 5% of the adult population (4–6). With declining iodine intake,
goiter incidence may increase. Additionally, nonendemic goiter may be caused by
excessive iodine intake (e.g., from amiodarone or kelp).
Solitary palpable thyroid nodules are found in 4% to 7% of the adult pop-
ulation in North America, 6.4% in women and 1.5% in men from 30 to 59 years
old (4,5,7,8). The prevalence of nodules increases throughout life, in women, and
in persons exposed to ionizing radiation in infancy or childhood. The data pro-
vided by the Framingham, Massachusetts, population study suggest an incidence
of 0.1% per year, or approximately 300,000 new nodules in the United States in
2007, and a 10% lifetime expectancy of nodule development (4,5,8).
However, these figures on palpable nodules substantially underestimate the
problem. For example, in an autopsy series, Mortensen and colleagues (9) reported
that in patients whose glands appeared clinically normal, one or more thyroid
Thyroid Nodules and Multinodular Goiter 205

nodules were detected in approximately 50%. Furthermore, 35% of patients had


nodules greater than 2 cm that had escaped clinical detection. More recent US
data support earlier autopsy results. Brander and associates (10) found that 30%
of asymptomatic adults had occult thyroid nodules detected with US, and Hor-
locker and colleagues (11) reported that 41% of 1000 patients with primary hyper-
parathyroidism had one or more thyroid nodules detected ultrasonographically
and later confirmed by surgery. Overall, most US studies suggest that unsus-
pected thyroid nodules are present in 20% to 50% of adult women and in 17%
to 30% of adult men. Thus, one can conclude that more than 100 million peo-
ple in the United States have asymptomatic, incidental, or subclinical thyroid
nodules (4,7).
In a multinodular thyroid, multiple small nodules are often identified by US
examination, whereas multiple nodules in an enlarged gland suggest a multinodu-
lar goiter (MNG). Recent imaging data have shown that in patients with clini-
cal solitary thyroid nodules, high-resolution US identifies one or more nodules
in approximately 50% (12). This finding suggests that the distinction between
clinically multinodular glands and clinically solitary nodules is not as sharp as
previously considered, and glands with an apparent single nodule likely have
many other nodules. Furthermore, the frequency of malignancy in patients with
single nodules and patients with multinodular glands seems to be almost identical.
In a study of 5637 patients, Belfiore and colleagues (13) reported thyroid can-
cer in 4.1% of clinically solitary nodules versus 4.7% of clinically multinodular
glands. However, another report showed that although cancer rates were similar
for patients with glands with one nodule or at least two nodules (14.8% vs. 14.9%,
respectively), the rate of malignancy was 14.8% for solitary nodules and 8.1% for
nonsolitary nodules (14).

PATHOGENESIS
Currently, the molecular mechanisms that stimulate the formation and growth
of only a few cells within a thyroid follicle and lead to nodule formation and
growth are poorly understood (15–17). TSH is a known stimulator and regulator
of differentiated thyroid follicular cell function. TSH interacts with follicular cell
receptors to activate adenylate cyclase and generate cyclic adenosine monophos-
phate; this in turn activates protein kinase A and leads to biochemical events that
stimulate the growth of follicular cells (17). Extrathyroidal factors may also act
on the intrinsic and abnormal growth potential of thyroid follicular cells, resulting
in accelerated nodular growth and development. Another possible mechanism is
that fibrous tissue, formed because of follicular necrosis and hemorrhage, may
promote thyroid nodularity in MNG. Nodule formation has also been attributed
to somatic mutations (15). It is now believed that benign and malignant thy-
roid tumors are monoclonal neoplasms arising from a single precursor cell that
206 Gharib

presumably gained a growth advantage through somatic mutation of genes critical


for growth regulation.
The most widely held theory about nodule formation and growth is that
chronic TSH stimulation initially causes development of a diffuse enlargement,
which in turn leads to thyroid nodule development, and ends with the formation
of multiple nodules, typically an MNG. However, it is important to emphasize
that the exact role of TSH as a thyroid growth factor remains controversial (18).
In recent years, non-TSH growth factors have been identified that stimulate thy-
roid cell growth: growth-stimulating immunoglobulin(s), epidermal growth factor,
insulin-like growth factors, interleukin-1, interferon-␥ , and transforming growth
factor-␤ (19). It is also noteworthy that the growth-promoting effects of TSH in
vitro depend largely on the presence of insulin-like growth factors. Additionally,
somatic mutations known to occur in thyroid follicular cells include ras onco-
genes, G proteins, and the TSH receptor gene, which result in hyperfunctioning
adenomas (15).
In summary, the exact cause(s) or molecular mechanism(s) that stimulate
growth of a single cell within a thyroid follicle is not known. It is likely that both
intra- and extrathyroidal factors are important in the development and growth of
nodules. These factors include follicular cell mutations, stimulation by TSH and
other growth factors, and development of follicular necrosis, hemorrhage, and
fibrous tissue.

HISTORY AND EXAMINATION


Thyroid nodules are usually asymptomatic and are often discovered by a patient
or a physician after careful palpation of the neck (Fig. 1). MNGs are also often
asymptomatic, but they may be associated with hoarseness, neck pressure or pain,
cough, dyspnea, or dysphagia. Features that increase the likelihood of thyroid
malignancy include a family history of thyroid cancer, age younger than 20 or
older than 60 years, and a history of head or neck irradiation (7,20,21). Nodules
are more common in women but are more likely to be malignant in men (1,7).
On physical examination, common and different causes of thyroid nodules
should be kept in mind (Table 2). The characteristics of thyroid nodules, includ-
ing location, consistency, dimensions, and number, need to be carefully recorded.
Physical findings suggestive of malignancy include a hard, nontender nodule, fixa-
tion to adjacent tissue, and the presence of regional lymphadenopathy. However, it
should be emphasized that history and physical examination are often insufficient
for diagnosing thyroid cancer in most patients.

LABORATORY AND RADIOLOGIC DIAGNOSIS


Serum TSH should be measured in all patients by using a sensitive, third-generation
assay. If TSH levels are less than 0.5 mIU/L, levels of free T4 and triiodothyronine
(T3 ) in the serum should be measured. If serum TSH is greater than 5.0 mIU/L,
Thyroid Nodules and Multinodular Goiter 207

Figure 1 An 81-year-old woman with a recently discovered 5-cm right thyroid nodule.
She reported that nodule size increased during the preceding 10 years. Serum level of
thyroid-stimulating hormone was 1.4 mIU/L; findings on fine-needle aspiration biopsy
were benign (colloid nodule). Thyroidectomy revealed a large, benign, colloid goiter.

measurement of free T4 and thyroid peroxidase antibodies are necessary (1).


Routine measurement of serum thyroglobulin is not recommended (1,2).
Currently, the issue of routine serum calcitonin (CT) measurement in patients
with thyroid nodules is controversial (1–4). The prevalence of sporadic medullary
thyroid carcinoma (MTC) in patients with thyroid nodules ranges from 0.3% to
1.3% (22–27). In these reports, some patients received the correct diagnosis of
MTC on the basis of serum CT measurement and not by FNA biopsy. Thus, CT
measurement may detect unsuspected MTC, and hence, the recommendation by

Table 2 Common Causes of Thyroid Nodules


Benign
Colloid
Cyst
Thyroiditis
Hürthle cell lesion
Follicular cell lesion
Malignant
Primary—papillary, follicular, medullary, anaplastic cancer, lymphoma
Secondary—breast, lung, kidney (3 most common sources)
208 Gharib

Table 3 Indications for Ultrasonography


Palpable nodule(s)
Nodule detected incidentally by imaginga
Difficult thyroid palpation
Unexplained cervical adenopathy
History of neck irradiation
Family history of papillary or medullary thyroid carcinoma, or multiple endocrine
neoplasia type 2
Follow-up of thyroid cancer
a Imaging methods include positron emission tomography, computed tomography, and
magnetic resonance imaging.

some experts is that CT assays should be performed routinely in all patients with
thyroid nodules (3,28).
According to some European groups (3), if basal CT levels exceed 10 pg/mL,
pentagastrin stimulation should be performed, and if peak CT levels exceed 100
pg/mL, thyroidectomy should follow. For values between 10 and 100 pg/mL,
medical follow-up is suggested (28,29). However, in the United States, rou-
tine CT measurement was not endorsed either in a recent expert opinion (29)
or in practice guidelines (1), which stated that routine testing of serum CT in
all patients with unselected thyroid nodules does not seem to be cost-effective
and is not recommended. It is also important to note that pentagastrin is cur-
rently not available in the United States, and this issue limits the use of CT
measurement.

Ultrasonography
Indications
Three recently published guidelines recommend thyroid US for all patients with
one or more suspected nodules (1–3). US is recommended to confirm the presence
of a nodule. One study showed that 18% of palpable abnormalities were not thyroid
nodules when examined by US (30). This method can also be used to identify any
suspicious sonographic features in the nodule and to document the presence of
other potentially more clinically significant nodules. Additional indications for
US are detailed in Table 3.
Technique
High-resolution US uses frequencies between 5 and 10 MHz, which allow mea-
surement of the volume of the gland and the number, size, and characteristics
of the nodules within it (1,31). Sound waves are produced and received by the
transducer. Images are obtained in the longitudinal and transverse planes. Cur-
rently, transducers can identify cystic or solid lesions as small as 1 to 2 mm in
the gland. High-resolution US equipment has become less expensive and more
Thyroid Nodules and Multinodular Goiter 209

A B

C D

Figure 2 Evaluation of thyroid with high-resolution ultrasonography. (A) Transverse view


of right thyroid nodule containing both cystic and solid components; nodule was benign
on biopsy. (B) Longitudinal view of solid thyroid nodule with a “halo” at nodule periph-
ery. Fine-needle aspiration (FNA) biopsy showed this was a benign follicular adenoma.
(C) Nodule with hypoechoic pattern, irregular margins, and punctate microcalcifications;
FNA biopsy showed papillary thyroid carcinoma, later confirmed at surgery. (D) Transverse
view of right thyroid lobe showing a solid hypoechoic nodule with scattered calcifications
suggestive of carcinoma; FNA biopsy suggested medullary thyroid carcinoma, later con-
firmed with thyroidectomy. Abbreviations: C, carotid artery; T, trachea.

user friendly; many endocrinologists use it in the office for careful evaluation of a
broad spectrum of thyroid abnormalities.

Results
High-resolution US patterns of the thyroid include consistency, echogenicity, pat-
terns of calcification, and color Doppler flow (1,32–34). On the basis of consis-
tency, thyroid nodules are divided into three categories: solid, cystic, and mixed
solid and cystic. Simple or pure cystic lesions are extremely rare, and most cystic
lesions are considered mixed (or “complex”) lesions [Fig. 2(A)]. Benign nodules
are hyperechoic and may have a sonolucent rim (“halo”) surrounding the nodule
[Fig. 2(B)]. A malignant thyroid nodule is typically an irregular, solid, hypoechoic
210 Gharib

Table 4 Value of US Features Predicting Thyroid Malignancy


Positive Negative
Sensitivity Specificity predictive predictive Relative
US feature (%) (%) value (%) value (%) risk

Microcalcifications 26.1–59.1 85.8–95.0 24.3–70.7 41.8–94.2 4.97


Hypoechogenicity 26.5–87.1 43.4–94.3 11.4–68.4 73.5–93.8 1.92
Irregular margins 17.4–77.5 38.9–85.0 9.3–60.0 38.9–97.8 16.83
or no halo
Solid 69.0–75.0 52.5–55.9 15.6–27.0 88.0–92.1 4.2a
Intranodule 54.3–74.2 78.6–80.8 24.0–41.9 85.7–97.4 14.29
vascularityb
More tall than 32.7 92.5 66.7 74.8 10.5a
wide
a Unpublished data from a series of 400 patients undergoing surgery for thyroid nodular disease.

Regina Apostolorum Hospital, Albano, Rome. Courtesy of Papini E. and Guglielmi R.


b Patterns of nodule color Doppler flow and malignancy were described previously (1,4).

Abbreviation: US, ultrasonography.


Source: From Ref. 33.

mass, although follicular cancers can be hyperechoic. Calcification occurs in 13%


of nodules. Types of calcifications are important—peripheral eggshell calcifica-
tions are seen in benign, degenerating adenomas, whereas increased blood flow on
Doppler imaging and microcalcifications (seen as punctate deposits and indicative
of psammoma bodies) suggest papillary carcinoma [Fig. 2(C)]. MTC is suspected
when a solid nodule has scattered calcification [Fig. 2(D)].
No single ultrasonographic feature unequivocally differentiates benign from
malignant nodules. Although US results are highly operator dependent, in expe-
rienced hands, US is useful for predicting the risk of malignancy (Table 4). For
example, microcalcifications have 86% to 95% specificity for cancer, whereas
hypoechogenicity has a lower value of 43% to 94%. Overall, identification of at
least two clinically suspicious sonographic criteria is 85% to 95% accurate for
diagnosing malignancy (1,4).

Scintigraphy
Indications
With increasing use of FNA biopsy in the diagnosis of thyroid nodules, the role of
thyroid scintigraphy has declined progressively. Indications for radioisotope scan-
ning are summarized in Table 5. If a patient has a palpable nodule and suppressed
serum TSH levels, the next appropriate test is a thyroid scan to determine whether
the nodule is functional (1,2,4). Thyroid scanning is also useful for determining
goiter size because extent of the goiter may influence clinical management. Fur-
thermore, radioisotope scans can evaluate nodule function in MNGs, determine the
Thyroid Nodules and Multinodular Goiter 211

Table 5 Indications for Radioisotope Scan


Diagnose hot nodule
Determine goiter size
Assess nodule function in multinodular goiter
Evaluate substernal goiter
Identify ectopic thyroid tissue (sublingual thyroid, struma ovarii)

extent and size of substernal goiter, identify ectopic thyroid tissue, and determine
function in a nodule of a patient with Graves’ disease.

Technique
A gamma scintillation camera with a pinhole collimator is generally used for
thyroid scanning; it has replaced the rectilinear scanner. The most commonly used
radioisotope is radioiodine (123 I); technetium (99m Tc) is being used less frequently
(4,35,36). Both isotopes are transported into thyroid follicular cells, but only
iodine is organified, that is, incorporated into hormone production. Theoretically,
a rare nodule may exhibit discordance between 123 I and 99m Tc scan findings.
Because 99m Tc is trapped but not organified, a nodule may appear functional (hot)
on pertechnetate imaging and nonfunctional (cold) on 123 I imaging. In practice,
this has not been an important problem. For studies with 99m Tc, 1 to 10 mCi is
administered intravenously and imaging is performed 20 minutes later.

Results
The radionuclide scan may show several different patterns of function (Fig. 3).
The normal thyroid has a butterfly shape, the two “wings” of which are con-
nected by an isthmus that crosses the trachea anteriorly, below the level of the
cricoid cartilage. In a healthy subject, the radioisotope signal appears homoge-
neously distributed [Fig. 3(A)]. Characteristically, salivary glands are visualized
with 99m Tc, whereas they are not visualized with 123 I because images are obtained
6 to 24 hours later (the isotope has cleared from the salivary glands by then).
In hyperthyroidism with increased uptake (Graves’ disease), an enlarged gland
has intense, diffuse, and homogeneous uptake [Fig. 3(B)]. A hypofunctioning
(cold) nodule means no or subnormal isotope concentration in the nodular tis-
sue [Fig. 3(C)]. The likelihood of carcinoma in cold thyroid nodules varies from
approximately 5% to 15% (1,35,36). Because approximately 85% of nodules are
cold, surgical treatment based on radionuclide scan results is not cost-effective.
Another scan pattern is that of a hyperfunctioning (hot) nodule, with the absence
or near absence of isotope in the rest of the thyroid gland due to suppression
of endogenous serum TSH [Fig. 3(D)]. Some small, hot nodules make insuffi-
cient amounts of thyroid hormone to cause TSH suppression, so the remaining
normal thyroid tissue is still visible. In MNG, the pattern has an inhomoge-
neous, irregular, or patchy appearance, with functioning and nonfunctioning areas
212 Gharib

A B

C D

Figure 3 Four different 99m Tc scan patterns. (A) Normal thyroid, showing function in
both lobes connected by the isthmus. (B) A 38-year-old man with hyperthyroid Graves’
disease, thyroid-stimulating hormone (TSH) of 0.006 mIU/L, and radioiodine uptake of
92%. Note that the scan shows enlarged thyroid gland with intense and diffuse uptake.
(C) A 38-year-old woman with a palpable, 2-cm cold nodule in the right thyroid lobe. The
nodule was benign on biopsy. (D) A 39-year-old man with a palpable 3-cm right thyroid
nodule, hyperfunctioning on scan, with completely suppressed uptake in the rest of the
gland. Serum level of TSH was 0.05 mIU/L and radioiodine uptake was 22%.

within an often enlarged gland. Patchy uptake may also be seen in Hashimoto
thyroiditis.

Other Imaging Techniques


Magnetic resonance imaging or computed tomography is used occasionally to
evaluate the size and extent of an MNG (Fig. 4). These imaging techniques
can be particularly helpful when evaluating substernal goiters and defining the
Thyroid Nodules and Multinodular Goiter 213

Figure 4 Computed tomography scan of the neck showing a large left thyroid mass with
tracheal deviation in a 56-year-old woman with long-standing and progressively enlarging
nodular goiter. Thyroidectomy uncovered a large benign colloid goiter.

relationship of the goiter to surrounding structures, especially to exclude tracheal


compression. Neither technique separates benign from malignant thyroid growths,
and because of their relatively high cost, they are used less frequently than US and
scintigraphy (1,37).

FNA Biopsy
FNA biopsy—now clearly considered the most accurate test for the diagnosis
of thyroid nodules—has emerged as a safe, accurate, and cost-effective test (1–
4,7,37–52).

Indications
Biopsies may be performed with direct palpation for clinically solitary thyroid
nodules or with US guidance. Ultrasonographically guided FNA (US-FNA) biopsy
should be performed on patients with small or impalpable cystic nodules. Patients
with diffuse goiters, for example, Hashimoto thyroiditis, subacute (granulomatous)
thyroiditis, or amyloid goiter, can undergo diagnostic FNA biopsy. Primary or
metastatic malignancy of the thyroid and adenopathy due to thyroid malignancy
can be diagnosed reliably by FNA biopsy (39).
214 Gharib

Technique
Biopsies can be performed on the examining table in the office or on a hospital bed;
the patient is supine and the neck is hyperextended by placing a pillow under the
shoulders. This position allows maximal exposure (37,38,45). The area for biopsy
is identified clearly and the skin cleansed with alcohol or betadine. Although some
physicians use 1% lidocaine for local anesthesia (47) or apply ice cubes contained
in a plastic bag (45), some argue that biopsies can be performed without any such
preparation (43). A 27- or 25-gauge needle attached to a 10-mL disposable plastic
syringe is used; the syringe can be attached to a mechanical syringe holder. After
the free hand of the operator locates the nodule, the needle is inserted rapidly into
the nodule and mild suction is applied immediately. Too much suction may dilute
the specimen with blood; this should be avoided. The needle is moved back and
forth several times; after a few seconds, the suction is released and the needle
withdrawn. The needle is then dislodged from the syringe, air is drawn into the
syringe, the needle is replaced in the syringe, and the aspirated material is forced
onto glass slides. (Alternatively, 1 cc of air can be drawn into the syringe before
aspiration.) A drop of aspirate is placed on each of several slides, and smears are
prepared with an additional glass slide, in a manner similar to that of preparing
blood smears. For wet-fixed smears, glass slides are placed immediately in 95%
ethyl alcohol. Air-dried slides are left unfixed. Usually, 8 to 12 slides from 2 to 4
aspirations are prepared per nodule; some physicians advocate at least 6 separate
aspirations per nodule. Many clinicians also wash the syringe with cytology fluid,
centrifuge the suspension, and microscopically examine the cell pellet. For larger
nodules, aspirates are obtained from the center and circumferentially from the
periphery. Slides are then taken to the laboratory; wet-fixed slides are stained with
a modified Papanicolaou stain and air-dried smears are stained with Diff-Quik or
other stains (1,39,46).
After the procedure is complete, local pressure is applied to the biopsy sites.
The patient is observed for a few minutes and is then allowed to leave.

Results
Results of FNA usually are categorized into four diagnostic groups: benign (neg-
ative), clinically suspicious (indeterminate), malignant (positive), and unsatisfac-
tory (nondiagnostic) (1,4,40,51,53). The most common benign cytodiagnosis is
that of adenomatoid (or “colloid”) or benign thyroid nodule. The colloid nodule
shows abundant colloid, normal follicular cells, and some foam cells. Other benign
diagnoses include cystic lesions (usually adenomatoid nodules with cystic degen-
eration), lymphocytic thyroiditis, and subacute (granulomatous) thyroiditis. The
indeterminate category consists of specimens with features that are suggestive of
but not diagnostic for malignancy. This group includes cellular specimens (e.g.,
follicular neoplasms, Hürthle cell neoplasms, and other specimens with varying
degrees of cellular atypia) that require histologic evaluation for a conclusive diag-
nosis (38,44,45,54,55). Characteristically, these nodules produce hypercellular
Thyroid Nodules and Multinodular Goiter 215

aspirates with microfollicular patterns, or Hürthle cell changes with scant or no


colloid. The malignant (positive) group includes primary or metastatic carcino-
mas and thyroid lymphoma. Papillary carcinomas account for more than 80% of
malignant lesions and are usually diagnosed confidently with FNA (28,43,44).
Common cytologic findings of the thyroid are shown in Figure 5.
Hypocellular smears, which account for 5% to 15% of specimens, are con-
sidered nondiagnostic or unsatisfactory (49,53,56–58). The criteria for judging
specimen adequacy are neither well defined nor standardized. The variability
between laboratories and clinics regarding the definition of an “adequate” speci-
men is considerable. Most cytologists believe that a satisfactory smear must contain
at least six clusters of well-preserved cells, and each group must be composed of
at least 10 cells from separate aspirates (44). Unsatisfactory smears usually result
from poor biopsy technique and, less often, from cystic lesions yielding fluid and
foam cells, vascular lesions yielding too much blood, excessive air-drying, or poor
smear preparation.
Published FNA series from many centers in various countries confirm its use-
fulness and accuracy (38,41,49,50,52). For example, a 1995 review (42) described
combined results from more than 16,500 specimens from two institutions—the
diagnostic results were as follows: benign, 69%; malignant, 4%; and suspicious
or nondiagnostic, 27%. Analysis of published data (Table 6) shows that FNA has
a sensitivity of 65% to 98% (mean, 83%), a specificity of 72% to 100% (mean,
92%), and a diagnostic accuracy of 85% to 100% (mean, 95%). The predictive
value of a positive or suspicious cytologic result is 75% (range, 50–96%). The
false-negative rate may be as low as 1% and as high as 11% (mean, 5%), and
false-positive rates range from 0% to 7% (mean, 5%) (38,49).
Limitations of FNA biopsy include unsatisfactory or nondiagnostic results,
suspicious or indeterminate cytologic findings, and false-negative diagnoses
(37,46–52,57–60). Although nondiagnostic results can be as high as 20%, a repeat
biopsy yields satisfactory results in half of the cases. US-FNA further reduces

Table 6 Summary Characteristics for Thyroid FNA Biopsies: Results of a Literature


Survey

Feature Mean Range Definition

Sensitivity (%) 83 65–98 Likelihood that a patient with disease has


positive test results
Specificity (%) 92 72–100 Likelihood that a patient without disease
has negative test result
Positive predictive 75 50–96 Fraction of patients with positive test
value (%) results and disease
False-negative rate (%) 5 1–11 FNA negative; histology positive for cancer
False-positive rate (%) 5 0–7 FNA positive; histology negative for cancer

Abbreviation: FNA, fine-needle aspiration.


Source: From Ref. 1.
216 Gharib

A B

C D

E F

Figure 5 Thyroid cytology. (A) Nondiagnostic smear. Degenerative foam cells with-
out follicular cells (PAP; ×60). (B) Colloid nodule. Cohesive group of thyroid cells in
a patient with multinodular goiter (PAP; ×50). (C) Hashimoto thyroiditis. Lymphocytes
and Hürthle cells showing abundant granular cytoplasm (PAP; ×250). (D). Follicular neo-
plasm. Hypercellular aspirate with microfollicular pattern lacking colloid is indeterminate
(PAP; ×205). Nodule was a benign follicular adenoma at surgery (E) Papillary carcinoma.
Cellular specimen showing tumor cells with irregular, enlarged nuclei. Note lack of colloid
(PAP; ×100). (F) Medullary carcinoma. Loosely cohesive neoplastic cells with elongated
nuclei. (MGG stain; ×400). Abbreviations: MGG, May–Grunwald–Giemsa stain; PAP,
Papanicolaou stain.
Thyroid Nodules and Multinodular Goiter 217

nondiagnostic rates, but there may be a residual 5% to 10% frequency of non-


diagnostic results. Surgical excision should be considered in patients with solid
nodules and persistently nondiagnostic results because the probability of malig-
nancy is not negligible (1,2,56,61). Follicular and Hürthle cell neoplasms pose a
difficult problem because benign nodules cannot be distinguished from malignant
lesions by FNA cytology alone, and they usually require histologic examination
of the tumor capsule to identify capsular or vascular invasion (or both). Because
approximately 20% of cytologically suspicious nodules are malignant, the current
recommendation is to excise these nodules surgically (1–4). Before surgery, a
radioiodine scan should be considered, even if the serum TSH level is normal,
to exclude the admittedly unlikely possibility of an autonomously functioning
thyroid nodule. False-negative results mean missed malignancy and are typically
the result of sampling errors and errors of interpretation (37,47,61). The adequacy
of sampling can be increased by carefully sampling different portions of a nodule,
using US-FNA for nodules less than 1 cm in diameter and obtaining multiple FNA
samples from large tumors (62–66).
Cytologic examination has proved to be both safe and accurate and is now
recommended as the primary diagnostic method for benign and malignant thyroid
lesions. Complications of FNA biopsy are minor, transient, and very rare. Bruising
or hematoma is infrequent and mild, and seeding in the needle tract is extremely
rare with FNA. Performance of probably 5 to 10 biopsies with supervision and
another 10 biopsies are necessary to acquire adequate experience, and at least 20
FNAs should be performed annually to maintain and upgrade biopsy technique.
There is evidence that greater experience with aspiration directly affects (reduces)
the insufficiency rate (60). Published reports document the high sensitivity and
specificity of the procedure. However, the advantages and limitations of FNA
biopsy should be recognized, and the procedure should be applied knowledgeably
to the evaluation of nodular thyroid disease.
US-FNA
In recent years, use of US-FNA in clinical practice has increased (1,4,34,62–65).
The indications for US-FNA include evaluation of (1) nodules that are impalpable
or difficult to palpate, (2) multiple thyroid nodules, (3) unexplained adenopathy
or palpable lymph nodes in patients with a history of thyroid cancer, (4) patients
with high risk of cancer such as MTC or multiple endocrine neoplasia type 2, (5)
patients whose initial biopsy results were nondiagnostic, and (6) nodules of any
size with US features suggestive of malignancy (Table 7). Several studies have
shown that US-FNA biopsies have a significantly lower percentage of inadequate
samples compared with direct (palpation-guided) FNA (64,65). With US-FNA,
the biopsy sites can be precisely selected and the needle correctly positioned to
allow sampling of the cyst walls or solid components. As a result, the rate of
satisfactory aspirates has increased. Color Doppler imaging may also be used to
obtain adequate aspirates (4). The application of US-FNA has resulted in fewer
thyroidectomies because of the increased yield of satisfactory aspirates (65,66).
218 Gharib

Table 7 Indications for Ultrasonographically Guided Fine-Needle Aspiration Biopsy


Impalpable or small (⬍1.5 cm) nodule with clinically suspicious ultrasonographic
features
Multiple nodules
Unexplained adenopathy or adenopathy in patients with history of cancer
Any size nodule in a patient with a history of neck radiation, medullary thyroid
carcinoma, or multiple endocrine neoplasia type 2
Initial biopsy was nondiagnostic
Any size nodule with ultrasonographic features suggestive of malignancy

Source: From Ref. 1.

Neither nodule number nor nodule size is predictive of malignancy (1,4,32).


When multiple nodules are present, selection for FNA should be made on the
basis of US features rather than size (1,4). When a patient has multiple, identi-
cally appearing spongiform nodules with little to no intervening normal thyroid
parenchyma, typical of a MNG, only the largest nodules require FNA. Some sug-
gest selecting only nodules larger than 1.0 or 1.5 cm in diameter for FNA (33),
citing the usual nonaggressive behavior of small (⬍1.5 cm) thyroid microcarcino-
mas. This issue is currently a matter of debate and controversy (34).

Conclusions
Recent treatment guidelines recommend routine serum TSH measurement, US
examination, and FNA biopsy in all patients with nodular thyroid disease (1–3).
US examination increasingly is used and has value in predicting malignancy.
However, the crux of nodule evaluation is thyroid cytology, and management
decisions should be made on the basis of FNA biopsy results. In experienced
centers, the probability of false-negative FNA rates (missed malignancy) is less
than 2%. The limitations of FNA include indeterminate (suspicious) cytology,
nondiagnostic smears, and false-negative FNA findings.

MANAGEMENT OF THYROID NODULES AFTER FNA


Solitary Nodules
A cytologically benign nodule can be monitored with or without additional aspi-
ration and with or without T4 therapy (Table 8). FNA-benign nodules account for
80% of satisfactory aspirates, require no further evaluation, and can be managed
expectantly. Considering 1% to 3% rate of false-negative diagnoses in this group,
some have suggested follow-up aspiration of benign nodules (67), whereas others
do not believe that additional biopsy is necessary (68). Generally, another aspira-
tion is required if nodules increase in size (1,2,4). Table 9 lists other indications
for performing additional biopsies.
Thyroid Nodules and Multinodular Goiter 219

Table 8 Management Options for Patients with FNA-Benign Nodules


Follow-up evaluation with palpation only
Follow-up evaluation with palpation and ultrasonography
Repeat FNA biopsy in 1 yr
T4 therapy

Abbreviation: FNA, fine-needle aspiration.

For patients with malignant FNA biopsy findings, surgical treatment is


indicated. The extent of thyroid surgery depends on the histologic type of malignant
disease: near-total thyroidectomy for papillary thyroid carcinoma or follicular
thyroid carcinoma and total thyroidectomy for MTC. Central compartment node
examination should be considered for patients with a diagnosis of papillary thyroid
carcinoma or MTC (1–3).
Approximately 10% of patients have a cytologic diagnosis that is clinically
suspicious for malignancy or otherwise indeterminate, which creates a difficult
dilemma for the clinician (Fig. 6). Although about 20% of these nodules are malig-
nant overall, if a cytologic study shows atypical features of papillary cancer, the
risk of malignancy is 60%, and if it shows follicular neoplasm, the risk is 15%
(4,49,53). Consequently, most clinicians believe that surgical excision of a nodule
with clinically suspicious cytologic features is appropriate. Intraoperative frozen
section may help the surgeon decide whether to perform only lobectomy and
isthmectomy when frozen sections show benign findings or perform a near-total
thyroidectomy if a specimen is malignant (34,69,70). Attempts have been made
to stratify patients with suspicious cytologic findings. For example, Schlinkert
and colleagues (71) studied 219 patients with suspicious follicular neoplasms and
reported that younger age, nodule size greater than 4 cm, solitary nodule, and
fixed primary nodule were predictive of malignancy. Similarly, Tuttle et al. (72)
reported a 21% incidence of malignancy in 103 patients with follicular neoplasia;
the risk of malignancy was significantly higher in males, nodules greater than
4 cm, and solitary nodules (shown by palpation). As noted above, it is suggested
to have radioisotope scanning for a cytologically suspicious nodule, with sur-
gical treatment for cold nodules and medical treatment for functioning nodules
(2,35,43).

Table 9 Reasons to Repeat FNA Biopsy


Enlarging, cytologically benign nodule
Recurrent cyst
Large nodule (⬎4 cm)a
Initial FNA biopsy was nondiagnostic
aTo minimize risk of missed malignancy in a large nodule.
Abbreviation: FNA, fine-needle aspiration.
220 Gharib

A B

C D

Figure 6 Suspicious thyroid cytology in a 46-year-old woman with a right thyroid nodule
(3 × 2 cm) and a history of childhood neck irradiation for acne. (A) Fine-needle aspiration
(FNA) biopsy showed a hypercellular specimen with microfollicular pattern, suspicious for
follicular neoplasm (PAP; ×25). (B) Thyroid scan showed that right thyroid nodule was
hypofunctioning. (C) Thyroidectomy showed area of focal hemorrhage corresponding to
FNA biopsy sites (arrow). (D) Histological examination showed benign follicular adenoma
with intact capsule (Hematoxylin and eosin; ×50).

Immunohistochemical markers have been studied in an effort to separate


benign from malignant nodules when FNA results show cellular smears that
are suspicious for follicular or Hürthle cell neoplasm (73–77). HBME-1 (78),
galectin-3 (79), and peroxidase and telomerase (76) have been repeatedly helpful
for identifying benign follicular cell lesions. However, none of these markers have
sufficiently high sensitivity and specificity for routine clinical use, and current
guidelines from the American Association of Clinical Endocrinologists and Asso-
ciazione Medici Endocrinologi, American Thyroid Association, and European
Thyroid Association, as well as expert opinions, do not endorse their clinical use
(1–4,34).
Approximately 15% of nodules are nondiagnostic on initial biopsy
(4,38,44,56,80,81). A nondiagnostic result should prompt a second aspiration,
which may be satisfactory for half of the cases; however, US-FNA may be neces-
sary for diagnosis of some cases. The routine removal of all nondiagnostic nodules
is not recommended; however, it seems reasonable to surgically remove recurrent
Thyroid Nodules and Multinodular Goiter 221

cysts larger than 4 cm in diameter, repeatedly nondiagnostic solid nodules, or other


lesions meeting clinical criteria that increase the risk of malignancy (1,2).

Thyroid Hormone Therapy


For the past 60 years, thyroid hormones have been used to suppress growth of
thyroid nodules. The rationale for suppressive therapy is based on the assumption
that because TSH stimulates nodule growth, its suppression should shrink nodules
or at least arrest the growth (19). By definition, suppressive therapy requires a
sufficiently high T4 dosage to suppress pituitary TSH secretion. The practice of
T4 suppressive therapy for nodular thyroid disease is controversial and debatable
(4,7,19,34).
During the past two decades, a number of randomized, controlled trials
have shown that few nodules shrink with thyroxine suppressive therapy (82–93).
Because each study included only a small number of subjects, several groups have
performed a meta-analysis of the data (94–96). Richter et al. (96) performed a
meta-analysis of nine studies (596 patients) and concluded that T4 treatment was
associated with decreased nodular volume (defined as ≥50% when measured by
US) in only 20% of patients (pooled relative risk, 1.83; 95% confidence interval,
0.9–3.73). Data analysis suggested that T4 therapy did not reach target effective-
ness (relative risk of 2.0) (Fig. 7).
In 1998, Gharib and Mazzaferri (97) reviewed data on suppressive trials,
evolving concepts, and controversies on suppressive therapy and offered their
recommendations. They concluded that nodules shrink in response to T4 therapy
in less than 20% of patients; the nodules that responded were usually the smallest
(⬍2.5 cm in diameter). The data did not show that T4 therapy prevented further
growth of existing nodules or emergence of new nodules. Furthermore, T4
therapy in dosages sufficient to suppress TSH may have adverse effects. It is
now established that long-term T4 suppressive therapy causes iatrogenic sub-
clinical hyperthyroidism that may be associated with osteopenia or osteoporosis
and altered cardiac function, which is especially notable in elderly persons
(4,7,8,98,99). A low serum level of TSH in persons 60 years or older is associated
with a 3-fold increase in the risk of atrial fibrillation (98). In addition, suppressive
therapy may be associated with decreased bone mineral density in postmenopausal
women (99).
Very little information is available about the outcome of untreated, asymp-
tomatic, benign, thyroid nodules. In reports from Japan, Kuma and associates
(100,101) examined the long-term outcome of untreated, cytologically benign,
thyroid nodules in 134 patients who received follow-up care for 10 to 30 years
(average, 15 years) and observed the following: nodules decreased in size in 53%
of patients, remained the same in 34%, and increased in size in only 13%. A
striking finding was that nodules were no longer palpable at the end of follow-up
in 30% of patients and had decreased in size in another 13%. Of the nodules that
completely disappeared, US showed that they were predominantly cystic.
222 Gharib

T4 suppressive therapy
change in nodule size

Cheung et al. (83)


Gharib et al. (82)

Papini et al. (93) pooled RR of 1.83


(95% CI, 0.90–3.73)
Reverter et al. (84)
Zelmanovitz et al. (94)

Figure 7 Reduction of nodule volume of at least 50% (random effects model). The right
side indicates improvement in reduction. The size of the filled diamond at the middle of
the central line (arrow 1) represents the sample size of each study. The box (arrow 2)
represents the 95% confidence interval (CI) of the relative risk (RR; marked with a line in
the box). The unfilled diamond with a central line (arrow 3) denotes the pooled risk ratio
itself. Abbreviation: T4 , levothyroxine. Source: From Ref. 96.

Recent guidelines from the American Association of Clinical Endocrinol-


ogists and Associazione Medici Endocrinologi (1) and the American Thyroid
Association (2) suggest that patients with cytologically benign nodules prefer-
entially should be monitored by palpation, with US whenever indicated, and
without routine T4 suppression. Adverse effects of suppressive therapy include
osteoporosis and cardiac arrhythmias, which can be clinically significant risks in
postmenopausal women or elderly patients.
Percutaneous Ethanol Injection
In Europe, percutaneous ethanol injection (PEI) has been used successfully in the
treatment of solid and cystic thyroid nodules with normal function and of hyper-
functioning nodules (102–113). Currently, it is recommended only for treatment
of cystic nodules and shows a size reduction of 50% or more in almost 90% of
cases treated (1,106).
The usual treatment protocol is a single bolus of ethanol administered with
a 20- to 22-gauge needle using ultrasonographic guidance. The procedure is per-
formed in an outpatient setting. A total of 2 to 50 mL may be injected in 2 to
12 weekly sessions, with 1 to 10 mL of ethanol per injection. For cystic lesions,
Thyroid Nodules and Multinodular Goiter 223

ethanol is injected slowly after cyst fluid is removed; the volume of ethanol injected
is decided on the basis of the aspirated fluid volume. Generally, PEI is given once
or twice weekly, and treatment is usually completed in 4 to 8 procedures. Com-
plications include transient dysphonia, local pain, hematoma, and mild fever. In
experienced hands, adverse effects are transient and minimal, and treatment is
generally well accepted by most patients (1,103,106).
The recurrence rate of cysts after PEI is low; one injection frequently
is curative, and nodule shrinkage persists long after treatment ends (106,109).
Although PEI is not used for benign cysts in the United States, Italian physicians
have reported large, successful treatment groups and remain enthusiastic about
the technique (1,106–108). PEI treatment seems to be appropriate for recurrent,
symptomatic, benign thyroid cysts.

Percutaneous Laser Thermal Ablation


Percutaneous laser thermal ablation is another medical procedure used to treat
thyroid nodules. This minimally invasive procedure directs laser beams to nodules,
reducing their size and the local symptoms (114–118). The technique requires use
of US, local anesthesia, and special training. One adverse effect is local pain that is
often transient; other problems have not been reported. Percutaneous laser thermal
ablation is currently available in some European centers but is not used widely or
routinely in clinical practice; also, it has never been used in the United States.

Conclusions
Cytologically benign nodules should be monitored without T4 therapy. Nodule
shrinkage for its own sake is an outcome that may not be of clinical value to either
the patient or physician. The potential risks of long-term T4 therapy outweigh the
potential benefits in most patients, particularly postmenopausal women (4,7,97).
One study (119) suggested benefit from T4 therapy in previously irradiated patients
who had undergone subtotal thyroidectomy for benign nodules. Treatment of this
group of patients with T4 seems reasonable.

Multinodular Goiter
Evaluation of patients with MNG includes determination of thyroid function, esti-
mation of goiter size, exclusion of malignancy, and assessment of local symptoms
(120). Evaluation should begin with a US examination and serum TSH determi-
nation; if TSH is suppressed, serum levels of thyroid hormone (free T4 and T3 )
are determined and a radioiodine uptake (RAIU) test is performed. Toxic MNG is
treated with radioiodine or surgery; subclinical hyperthyroidism may be observed
for a period of time and then reassessed. A normal serum concentration of TSH
indicates a nontoxic MNG. Cytologic evaluation by US-FNA biopsy helps deter-
mine the management strategy (Fig. 8). In patients with benign goiters, periodic
evaluation—including thyroid palpation, determination of serum levels of TSH,
224 Gharib

Figure 8 Management of patient with a multinodular goiter (MNG). Evaluation begins


by determining thyroid-stimulating hormone (TSH) levels; suppressed TSH (⬍0.1 mIU/L)
suggests subclinical or clinical hyperthyroidism and the patient is treated accordingly.
Most often, when TSH is normal (nontoxic goiter), fine-needle aspiration (FNA) biopsy
results decide management. Benign and/or small goiters are followed without thyroxine
therapy. Symptomatic, large MNGs are treated with either surgery or radioiodine (131 I).
Malignant goiters are surgically excised. Abbreviations: FT4 , free thyroxine; N, normal;
RAIU, radioiodine uptake; Rx, therapy; T3 , triiodothyronine; US, ultrasound.

and cross-sectional imaging studies—is helpful in management. Cytologically


suspicious or malignant MNGs should be treated surgically.
Surgical excision is preferred in patients with nontoxic MNG and local
compression symptoms or cosmetic concerns. Bilateral subtotal thyroidectomy
is standard therapy for patients with MNG. The most frequent complications
of thyroidectomy include injury of the recurrent laryngeal nerve, hypoparathy-
roidism, hypothyroidism, and postoperative bleeding. However, with experienced
surgeons, complication rates are quite low. Postoperatively, T4 replacement is
frequently administered to prevent recurrence of goiter, but recent studies have
suggested that this therapy is probably ineffective; hence routine T4 therapy is no
longer advised (7,97). In patients with primary lobectomy, T4 therapy is advised
only if hypothyroidism develops.
Radioiodine (131 I) has been used successfully to treat toxic and nontoxic
MNGs (120–126). Radioiodine therapy results in amelioration of hyperthyroidism,
Thyroid Nodules and Multinodular Goiter 225

reduction of goiter size, and decreased local pressure or pain. Huysmans and col-
leagues (122,123) described patients with nontoxic goiters who were treated with
radioiodine and reported symptomatic improvement in 71%, decrease in tracheal
deviation in 20%, and increased tracheal lumen in 36%. Therapeutic doses of
radioiodine have ranged from 25 to 150 mCi, but the dosage depends on RAIU
and the mass of thyroid tissue. Adverse effects include posttherapy hyperthy-
roidism, hypothyroidism, and thyroiditis. Also, the possibility of malignant change
in residual thyroid tissue is always a concern (7,125–127).
Currently, radioiodine generally is considered an alternative and effective
treatment to surgical thyroidectomy; it is a good choice for treatment of small
(volume, ⬍100 mL) and nontoxic MNG, for patients previously treated with
thyroidectomy, or for elderly patients with high risk of surgical intervention (124).
In areas of high iodine intake, RAIU may be low or low-normal, thus reducing
the efficacy of radioiodine treatment and requiring increased doses of radioiodine
(120). Recent reports suggest that recombinant human TSH (rhTSH) may increase
RAIU and benefit patients who are candidates for this nonsurgical treatment
(128,129). The administration of small doses (0.10–0.30 mg) of rhTSH results
in 4-fold increased uptake, 24 to 72 hours after injection (127–131). Thus, the
131
I becomes more effective for goiter volume reduction, but rhTSH may increase
risk of respiratory problems (attributable to acute goiter enlargement), transient
posttherapy hyperthyroidism, and permanent hypothyroidism (132). Smaller doses
of rhTSH (0.01 mg) may be as effective and may have fewer serious adverse
effects.
In conclusion, serum TSH level, FNA biopsy, and imaging with radioiso-
tope scanning, US, computed tomography, or magnetic resonance help delineate
function, morphology, and extent of an MNG. Indications for treatment include
tracheal compression, cosmesis, and concern about malignancy on the basis of
growth or cytologic findings (or both). Thyroxine therapy reduces nodule or goiter
size in a minority of patients, and its routine use is no longer endorsed (1,2,7,8,97).
However, T4 should never be used in patients who already have low-normal or
suppressed serum TSH levels for fear of toxicity in an already autonomous gland.
Surgery is standard therapy for nontoxic MNGs, and radioiodine, in large doses,
is an attractive alternative therapy in elderly patients and those considered to have
high risk for surgery.

INCIDENTALOMA
Nonpalpable thyroid nodules discovered incidentally on imaging examination are
described as “incidentalomas.” They are usually less than 1.5 cm in diameter, are
a common clinical problem, and constitute a management dilemma for clinicians
(12,42,133). Most of them are discovered when high-resolution US is used for
parathyroid evaluation, carotid disease, or other nonthyroid diseases of the neck.
Incidentalomas are discovered in 30% to 50% of the normal population without
226 Gharib

thyroid disease who undergo neck US. The prevalence of incidentalomas appears
higher in the elderly and in persons with iodine deficiency or radiation exposure
(133). The results of an older autopsy study (9) and more recent US-FNA study
(66) suggest that fewer than 5% of asymptomatic nodules may be malignant.
Schneider and coworkers (21) studied the results of US in patients with a history
of upper-body irradiation. They reported that 87% of patients had one or more
discrete nodules on US and that 75% of nodules were less than 1 cm in diameter.
The authors concluded that thyroid US is more sensitive than physical examination
and radioisotope scanning. Patients with no history of radiation exposure should
not undergo a US examination when thyroid palpation is normal. Those with
radiation exposure may undergo periodic US and US-FNA whenever indicated.
However, because US is so sensitive, great caution must be used when interpreting
the results.
We previously have reviewed the clinical importance of thyroid inci-
dentalomas and proposed practical management guidelines (Fig. 9) (133). If

Figure 9 Diagnostic approach to patient with incidentaloma. The algorithm separates


patients into high-risk and low-risk groups on the basis of clinical assessment. Nodule size
and appearance on high-resolution ultrasonography (US) also influence management. For
low-risk patients with benign-appearing incidentalomas less than 1.5 cm in diameter, obser-
vation and follow-up palpation are sufficient. Abbreviation: FNA, fine-needle aspiration.
Source: From Ref. 133.
Thyroid Nodules and Multinodular Goiter 227

incidentalomas do not have sonographic features of malignancy and are smaller


than 1.0 to 1.5 cm in diameter, and if the clinical history is not suggestive of
increased risk for thyroid cancer, follow-up neck palpation and US at 6 months and
annually thereafter is a practical, cost-effective approach. However, for patients
with nodules larger than 1.0 to 1.5 cm in diameter, with previous neck irradia-
tion, or with imaging characteristics suspicious for malignancy, US-FNA should
be performed. Suspicious imaging features include a solid, hypoechoic nodule
with irregular borders, sometimes containing punctate microcalcifications and
increased blood flow on Doppler imaging. Nonpalpable nodules that are predom-
inately cystic are most likely benign.
In conclusion, incidental thyroid micronodules are common, recognized
with increasing frequency, and are commonly benign. US-FNA should be per-
formed when clinical history or US examination findings are suspicious for
malignancy.

CONCLUSIONS
Thyroid nodules are very common and have a malignancy risk of less than 5%. This
risk seems independent of either nodule size or number. Initial evaluation should
include a serum TSH measurement, US examination, and FNA biopsy. FNA is
now established as a safe and accurate diagnostic test. US-FNA biopsy should
be performed if the nodule is small (1.0–1.5 cm), impalpable, or if a previous
FNA biopsy result was nondiagnostic. Clinically suspicious or malignant nodules
should be treated surgically.
FNA-benign nodules should have careful follow-up that includes palpation
and possibly US in 12 months. Although some recommend routine, additional
aspiration of benign nodules, this is not universally accepted. T4 suppression is
no longer recommended because few nodules decrease in size with suppressive
therapy, and important adverse effects may accompany treatment. Most benign
nodules remain stable in size; enlarging nodules should be evaluated by reaspira-
tion.
MNGs can be toxic or nontoxic, and patients in either group may require
treatment. Standard treatment is surgical thyroidectomy, although radioiodine
increasingly is used for patients who refuse surgery or are not candidates for
surgical treatment. When RAIU is low, rhTSH can be used to stimulate uptake
and improve results of treatment.
Certain management issues remain controversial, including routine CT mea-
surement; the application of immunohistochemical markers in nodules with clin-
ically suspicious biopsy findings; how many and which nodule(s) should undergo
FNA biopsy when patients have multiple nodules; extent of initial surgery for
nodule with clinically suspicious biopsy findings; and the role of radioiodine in
the treatment of MNG.
228 Gharib

REFERENCES

1. AACE/AME Task Force on Thyroid Nodules. American Association of Clinical


Endocrinologists and Associazione Medici Endocrinologi medical guidelines for
clinical practice for the diagnosis and management of thyroid nodules. Endocr Pract
2006; 12(1):63–102.
2. Cooper DS, Doherty GM, Haugen BR, et al. The American Thyroid Association
Guidelines Task Force. Management guidelines for patients with thyroid nodules
and differentiated thyroid cancer. Thyroid 2006; 16(2):109–142.
3. Pacini F, Schlumberger M, Dralle H, et al. European Thyroid Cancer Task Force.
European consensus for the management of patients with differentiated thyroid
carcinoma of the follicular epithelium. Eur J Endocrinol 2006; 154(6):787–803
[Erratum: Eur J Endocrinol 2006; 155(2):385].
4. Gharib H, Papini E. Thyroid nodules: Clinical importance, assessment, and treat-
ment. Endocrinol Metab Clin North Am 2007; 36(3):707–735.
5. Vander JB, Gaston EA, Dawber TR. The significance of nontoxic thyroid nodules:
Final report of a 15-year study of the incidence of thyroid malignancy. Ann Intern
Med 1968; 69(3):537–540.
6. Tunbridge WM, Evered DC, Hall R, et al. The spectrum of thyroid disease in a
community: The Whickham survey. Clin Endocrinol (Oxf) 1977; 7(6):481–493.
7. Hegedus L. Clinical practice: The thyroid nodule. N Engl J Med 2004; 351(17):1764–
1771.
8. Mazzaferri EL. Management of a solitary thyroid nodule. N Engl J Med 1993;
328(8):553–559.
9. Mortensen JD, Woolner LB, Bennett WA. Gross and microscopic findings in clini-
cally normal thyroid glands. J Clin Endocrinol Metab 1955; 15(10):1270–1280.
10. Brander A, Viikinkoski P, Nickels J, et al. Thyroid gland: US screening in a random
adult population. Radiology 1991; 181(3):683–687.
11. Horlocker TT, Hay JE, James EM, et al. Prevalence of incidental nodular thyroid
disease detected during high-resolution parathyroid ultrasonography. In: Medeiros-
Neto G, Gaitan E, eds. Frontiers in Thyroidology, Vol. 2. New York, NY: Plenum
Medical Book Co., 1986:1309–1312.
12. Tan GH, Gharib H, Reading CC. Solitary thyroid nodule: Comparison between
palpation and ultrasonography. Arch Intern Med 1995; 155(22):2418–2423.
13. Belfiore A, La Rosa GL, La Porta GA, et al. Cancer risk in patients with cold thyroid
nodules: Relevance of iodine intake, sex, age, and multinodularity. Am J Med 1992;
93(4):363–369.
14. Frates MC, Benson CB, Doubilet PM, et al. Prevalence and distribution of carci-
noma in patients with solitary and multiple thyroid nodules on sonography. J Clin
Endocrinol Metab 2006; 91(9):3411–3417.
15. Fagin JA, ed. Thyroid Cancer. Boston, MA: Kluwer Academic, 1998:59–83.
16. Derwahl M, Studes H. Pathogenesis and treatment of multinodular goiter. In: Fagin
JA, ed. Thyroid Cancer. Boston, MA: Kluwer Academic, 1998:155–186.
17. Fagin JA. Molecular pathogenesis of human thyroid neoplasm. Thyroid Today 1994;
17(3):1–7.
18. Derwahl M, Broecker M, Kraiem Z. Clinical review 101: Thyrotropin may not be the
dominant growth factor in benign and malignant thyroid tumors. J Clin Endocrinol
Metab 1999; 84(3):829–834.
Thyroid Nodules and Multinodular Goiter 229

19. Smith SA, Gharib H. Thyroid nodule suppression. Adv Endocrinol Metab 1991;
2:107–124.
20. Burch HB. Evaluation and management of the solid thyroid nodule. Endocrinol
Metab Clin North Am 1995; 24(4):663–710.
21. Schneider AB, Bekerman C, Leland J, et al. Thyroid nodules in the follow-up of irra-
diated individuals: Comparison of thyroid ultrasound with scanning and palpation.
J Clin Endocrinol Metab 1997; 82(12):4020–4027.
22. Pacini F, Fontanelli M, Fugazzola L, et al. Routine measurement of serum calcitonin
in nodular thyroid diseases allows the preoperative diagnosis of unsuspected sporadic
medullary thyroid carcinoma. J Clin Endocrinol Metab 1994; 78(4):826–829.
23. Rieu M, Lame MC, Richard A, et al. Prevalence of sporadic medullary thyroid carci-
noma: The importance of routine measurement of serum calcitonin in the diagnostic
evaluation of thyroid nodules. Clin Endocrinol (Oxf) 1995; 42(5):453–460.
24. Vierhapper H, Raber W, Bieglmayer C, et al. Routine measurement of plasma cal-
citonin in nodular thyroid diseases. J Clin Endocrinol Metab 1997; 82(5):1589–
1593.
25. Henry JF, Denizot A, Puccini M, et al. Latent subclinical medullary thyroid carci-
noma: Diagnosis and treatment. World J Surg 1998; 22(7):752–756.
26. Nicolli P, Wion-Barbot N, Caron P, et al. The French Medullary Study Group.
Interest of routine measurement of serum calcitonin: Study in a large series of
thyroidectomized patients. J Clin Endocrinol Metab 1997; 82(2):338–341.
27. Niederle B, Scheuba C. Invited commentary. World J Surg 1998; 22:756–757.
28. Borget I, De Pouvourville G, Schlumberger M. Editorial: Calcitonin determination
in patients with nodular thyroid disease. J Clin Endocrinol Metab 2007; 92(2):425–
427.
29. Hodak SP, Burman KD. The calcitonin conundrum: Is it time for routine measure-
ment of serum calcitonin in patients with thyroid nodules? J Clin Endocrinol Metab
2004; 89(2):511–514.
30. Marqusee E, Benson CB, Frates MC, et al. Usefulness of ultrasonography in the
management of nodular thyroid disease. Ann Intern Med 2000; 133(9):696–700.
31. Blum M, Yee J. Advances in thyroid imaging: Thyroid sonography: When and how
should it be used? Thyroid Today 1997; 20(3):1–13.
32. Papini E, Guglielmi R, Bianchini A, et al. Risk of malignancy in nonpalpable thyroid
nodules: Predictive value of ultrasound and color-Doppler features. J Clin Endocrinol
Metab 2002; 87(5):1941–1946.
33. Frates MC, Benson CB, Charboneau JW, et al. Society of Radiologists in Ultra-
sound. Management of thyroid nodules detected at US: Society of Radiologists in
Ultrasound consensus conference statement. Radiology 2005; 237(3):794–800.
34. Castro MR, Gharib H. Continuing controversies in the management of thyroid nod-
ules. Ann Intern Med 2005; 142(11):926–931.
35. Gharib H. Changing concepts in the diagnosis and management of thyroid nodules.
Endocrinol Metab Clin North Am 1997; 26(4):777–800.
36. Rojeski MT, Gharib H. Nodular thyroid disease: Evaluation and management.
N Engl J Med 1985; 313(7):428–436.
37. Gharib H, Goellner JR. Evaluation of nodular thyroid disease. Endocrinol Metab
Clin North Am 1988; 17(3):511–526.
38. Gharib H. Fine-needle aspiration biopsy of thyroid nodules: Advantages, limitations,
and effect. Mayo Clin Proc 1994; 69(1):44–49.
230 Gharib

39. Atkinson BF. Fine needle aspiration of the thyroid. Monogr Pathol 1993; 35:166–
199.
40. Gharib H. Diagnosis of thyroid nodules by fine-needle aspiration biopsy. Curr Opin
Endocrinol 1996; 3:433–438.
41. Gharib H. Management of thyroid nodules: Another look. Thyroid Today 1997;
20(1):1–11.
42. Giuffrida D, Gharib H. Controversies in the management of cold, hot, and occult
thyroid nodules. Am J Med 1995; 99(6):642–650.
43. Singer PA. Evaluation and management of the solitary thyroid nodule. Otolaryngol
Clin North Am 1996; 29(4):577–591.
44. Gharib H, Goellner JR. Fine-needle aspiration biopsy of thyroid nodules. Endocr
Pract 1995; 1(6):410–417.
45. Oertel YC. Fine-needle aspiration of the thyroid: Technique and terminology.
Endocrinol Metab Clin North Am 2007; 36(3):737–751.
46. Jeffrey PB, Miller TR. Fine-needle aspiration cytology of the thyroid. Pathology
(Phila) 1996; 4(2):319–335.
47. Hamburger JI. Diagnosis of thyroid nodules by fine needle biopsy: Use and abuse.
J Clin Endocrinol Metab 1994; 79(2):335–339.
48. Gharib H, Goellner JR, Johnson DA. Fine-needle aspiration cytology of the thyroid:
A 12-year experience with 11,000 biopsies. Clin Lab Med 1993; 13(3):699–709.
49. Gharib H, Goellner JR. Fine-needle aspiration biopsy of the thyroid: An appraisal.
Ann Intern Med 1993; 118(4):282–289.
50. Bisi H, de Camargo RY, Longatto Filho A. Role of fine-needle aspiration cytology in
the management of thyroid nodules: Review of experience with 1,925 cases. Diagn
Cytopathol 1992; 8(5):504–510.
51. Goellner JR, Gharib H, Grant CS, et al. Fine needle aspiration cytology of the
thyroid, 1980–1986. Acta Cytol 1987; 31(5):587–590.
52. Caruso D, Mazzaferri EL. Fine needle aspiration biopsy in the management of
thyroid nodules. Endocrinologist 1991; 1:194–202.
53. Castro MR, Gharib H. Thyroid fine-needle aspiration biopsy: Progress, practice, and
pitfalls. Endocr Pract 2003; 9(2):128–136.
54. Caraway NP, Sneige N, Samaan NA. Diagnostic pitfalls in thyroid fine-needle aspi-
ration: A review of 394 cases. Diagn Cytopathol 1993; 9(3):345–350.
55. Cersosimo E, Gharib H, Suman VJ, et al. Suspicious thyroid cytologic findings:
Outcome in patients without immediate surgical treatment. Mayo Clin Proc 1993;
68(4):343–348.
56. MacDonald L, Yazdi HM. Nondiagnostic fine needle aspiration biopsy of the thyroid
gland: A diagnostic dilemma. Acta Cytol 1996; 40(3):423–428.
57. Powers CN, Frable WJ. Fine Needle Aspiration Biopsy of the Head and Neck.
Boston, MA: Butterworth-Heinemann, 1996.
58. Schmidt T, Riggs MW, Speights VO Jr. Significance of nondiagnostic fine-needle
aspiration of the thyroid. South Med J 1997; 90(12):1183–1186.
59. Musgrave YM, Davey DD, Weeks JA, et al. Assessment of fine-needle aspiration
sampling technique in thyroid nodules. Diagn Cytopathol 1998; 18(1):76–80.
60. Burch HB, Burman KD, Reed HL, et al. Fine needle aspiration of thyroid nodules:
Determinants of insufficiency rate and malignancy yield at thyroidectomy. Acta
Cytol 1996; 40(6):1176–1183.
Thyroid Nodules and Multinodular Goiter 231

61. Hall TL, Layfield LJ, Philippe A, et al. Sources of diagnostic error in fine needle
aspiration of the thyroid. Cancer 1989; 63(4):718–725.
62. Cochand-Priollot B, Guillausseau PJ, Chagnon S, et al. The diagnostic value of fine-
needle aspiration biopsy under ultrasonography in nonfunctional thyroid nodules:
A prospective study comparing cytologic and histologic findings. Am J Med 1994;
97(2):152–157 [Erratum: Am J Med 1994; 97(3):311].
63. Carmeci C, Jeffrey RB, McDougall IR, et al. Ultrasound-guided fine-needle aspira-
tion biopsy of thyroid masses. Thyroid 1998; 8(4):283–289.
64. Danese D, Sciacchitano S, Farsetti A, et al. Diagnostic accuracy of conventional
versus sonography-guided fine-needle aspiration biopsy of thyroid nodules. Thyroid
1998; 8(1):15–21.
65. Izquierdo R, Arekat MR, Knudson PE, et al. Comparison of palpation-guided versus
ultrasound-guided fine-needle aspiration biopsies of thyroid nodules in an outpatient
endocrinology practice. Endocr Pract 2006; 12(6):609–614.
66. Hagag P, Strauss S, Weiss M. Role of ultrasound-guided fine-needle aspiration biopsy
in evaluation of nonpalpable thyroid nodules. Thyroid 1998; 8(11):989–995.
67. Chehade JM, Silverberg AB, Kim J, et al. Role of repeated fine-needle aspiration
of thyroid nodules with benign cytologic features. Endocr Pract 2001; 7(4):237–
243.
68. Lucas A, Llatjos M, Salinas I, et al. Fine-needle aspiration cytology of benign nodular
thyroid disease: Value of re-aspiration. Eur J Endocrinol 1995; 132(6):677–680.
69. Lee TI, Yang HJ, Lin SY, et al. The accuracy of fine-needle aspiration biopsy and
frozen section in patients with thyroid cancer. Thyroid 2002; 12(7):619–626.
70. Paphavasit A, Thompson GB, Hay ID, et al. Follicular and Hürthle cell thyroid
neoplasms. Is frozen-section evaluation worthwhile? Arch Surg 1997; 132(6):674–
678.
71. Schlinkert RT, van Heerden JA, Goellner JR, et al. Factors that predict malignant
thyroid lesions when fine-needle aspiration is “suspicious for follicular neoplasm”.
Mayo Clin Proc 1997; 72(10):913–916.
72. Tuttle RM, Lemar H, Burch HB. Clinical features associated with an increased risk
of thyroid malignancy in patients with follicular neoplasia by fine-needle aspiration.
Thyroid 1998; 8(5):377–383.
73. Komatsu M, Kuroda T. Flow cytometric nuclear DNA analysis of follicular thyroid
nodules. Res Surg 1995; 7:36–41.
74. Karmakar T, Dey P. Role of AgNOR in diagnosis of thyroid follicular neoplasms on
fine-needle aspiration smears. Diagn Cytopathol 1995; 12(2):148–149.
75. Delbridge L, Lean CL, Russell P, et al. Proton magnetic resonance and human thyroid
neoplasia. II: Potential avoidance of surgery for benign follicular neoplasms. World
J Surg 1994; 18(4):512–516.
76. Yamashita H, Noguchi S, Murakami N, et al. Immunohistological differentiation
of benign thyroid follicular cell tumors from malignant ones: Usefulness of anti-
peroxidase and JT-95 antibodies. Acta Pathol Jpn 1993; 43(11):670–673.
77. Segev DL, Clark DP, Zeiger MA, et al. Beyond the suspicious thyroid fine needle
aspirate: A review. Acta Cytol 2003; 47(5):709–722.
78. Miettinen M, Karkkainen P. Differential reactivity of HBME-1 and CD15 antibodies
in benign and malignant thyroid tumours: Preferential reactivity with malignant
tumours. Virchows Arch 1996; 429(4–5):213–219.
232 Gharib

79. Bartolazzi A, Gasbarri A, Papotti M, et al. Thyroid Cancer Study Group. Application
of an immunodiagnostic method for improving preoperative diagnosis of nodular
thyroid lesions. Lancet 2001; 357(9269):1644–1650.
80. McHenry CR, Walfish PG, Rosen IB. Non-diagnostic fine needle aspiration biopsy:
A dilemma in management of nodular thyroid disease. Am Surg 1993; 59(7):415–
419.
81. Baloch ZW, Sack MJ, Yu GH, et al. Fine-needle aspiration of thyroid: An institutional
experience. Thyroid 1998; 8(7):565–569.
82. Gharib H, James EM, Charboneau JW, et al. Suppressive therapy with levothyroxine
for solitary thyroid nodules: A double-blind controlled clinical study. N Engl J Med
1987; 317(2):70–75.
83. Cheung PS, Lee JM, Boey JH. Thyroxine suppressive therapy of benign solitary
thyroid nodules: A prospective randomized study. World J Surg 1989; 13(6):818–
821.
84. Reverter JL, Lucas A, Salinas I, et al. Suppressive therapy with levothyroxine for
solitary thyroid nodules. Clin Endocrinol (Oxf) 1992; 36(1):25–28.
85. Berghout A, Wiersinga WM, Drexhage HA, et al. Comparison of placebo with
l-thyroxine alone or with carbimazole for treatment of sporadic non-toxic goitre.
Lancet 1990; 336(8709):193–197.
86. Celani MF. Levothyroxine suppressive therapy in the medical management of
nontoxic benign multinodular goiter. Exp Clin Endocrinol 1993; 101(5):326–
332.
87. Celani MR, Mariani M, Mariani G. On the usefulness of levothyroxine suppressive
therapy in the medical treatment of benign solitary, solid or predominantly solid,
thyroid nodules. Acta Endocrinol (Copenh) 1990; 123(6):603–608.
88. Diacinti D, Salabe GB, Olivieri A, et al. Efficacy of L-thyroxine (L-T4 ) therapy on
the volume of the thyroid gland and nodules in patients with euthyroid nodular goiter
(ENG) [Italian]. Minerva Med 1992; 83(11):745–751.
89. La Rosa GL, Ippolito AM, Lupo L, et al. Cold thyroid nodule reduction with L-
thyroxine can be predicted by initial nodule volume and cytological characteristics.
J Clin Endocrinol Metab 1996; 81(12):4385–4387.
90. La Rosa GL, Lupo L, Giuffrida D, et al. Levothyroxine and potassium iodide are
both effective in treating benign solitary solid cold nodules of the thyroid. Ann Intern
Med 1995; 122(1):1–8.
91. Mainini E, Martinelli I, Morandi G, et al. Levothyroxine suppressive therapy for
solitary thyroid nodule. J Endocrinol Invest 1995; 18(10):796–799.
92. Wemeau JL, Caron P, Schvartz C, et al. Effects of thyroid-stimulating hormone sup-
pression with levothyroxine in reducing the volume of solitary thyroid nodules and
improving extranodular nonpalpable changes: A randomized, double-blind, placebo-
controlled trial by the French Thyroid Research Group. J Clin Endocrinol Metab
2002; 87(11):4928–4934.
93. Papini E, Bacci V, Panunzi C, et al. A prospective randomized trial of levothyrox-
ine suppressive therapy for solitary thyroid nodules. Clin Endocrinol (Oxf) 1993;
38(5):507–513.
94. Zelmanovitz F, Genro S, Gross JL. Suppressive therapy with levothyroxine for
solitary thyroid nodules: A double-blind controlled clinical study and cumulative
meta-analyses. J Clin Endocrinol Metab 1998; 83(11):3881–3885.
Thyroid Nodules and Multinodular Goiter 233

95. Castro MR, Caraballo PJ, Morris JC. Effectiveness of thyroid hormone suppressive
therapy in benign solitary thyroid nodules: A meta-analysis. J Clin Endocrinol Metab
2002; 87(9):4154–4159.
96. Richter B, Neises G, Clar C. Pharmacotherapy for thyroid nodules: A systematic
review and meta-analysis. Endocrinol Metab Clin North Am 2002; 31(3):699–722.
97. Gharib H, Mazzaferri EL. Thyroxine suppressive therapy in patients with nodular
thyroid disease. Ann Intern Med 1998; 128(5):386–394.
98. Faber J, Galle AM. Changes in bone mass during prolonged subclinical hyperthy-
roidism due to L-thyroxine treatment: A meta-analysis. Eur J Endocrinol 1994;
130(4):350–356.
99. Sawin CT, Geller A, Wolf PA, et al. Low serum thyrotropin concentrations as a risk
factor for atrial fibrillation in older persons. N Engl J Med 1994; 331(19):1249–1252.
100. Kuma K, Matsuzuka F, Kobayashi A, et al. Outcome of long standing solitary thyroid
nodules. World J Surg 1992; 16(4):583–587.
101. Kuma K, Matsuzuka F, Yokozawa T, et al. Fate of untreated benign thyroid nodules:
Results of long-term follow-up. World J Surg 1994; 18(4):495–498.
102. Livraghi T, Paracchi A, Ferrari C, et al. Treatment of autonomous thyroid nodules
with percutaneous ethanol injection: Preliminary results. Work in progress. Radiol-
ogy 1990; 175(3):827–829.
103. Verde G, Papini E, Pacella CM, et al. Ultrasound guided percutaneous ethanol
injection in the treatment of cystic thyroid nodules. Clin Endocrinol (Oxf) 1994;
41(6):719–724.
104. Zingrillo M, Torlontano M, Chiarella R, et al. Percutaneous ethanol injection may
be a definitive treatment for symptomatic thyroid cystic nodules not treatable by
surgery: Five-year follow-up study. Thyroid 1999; 9(8):763–767.
105. Kim JH, Lee HK, Lee JH, et al. Efficacy of sonographically guided percutaneous
ethanol injection for treatment of thyroid cysts versus solid thyroid nodules. AJR
Am J Roentgenol 2003; 180(6):1723–1726.
106. Valcavi R, Frasoldati A. Ultrasound-guided percutaneous ethanol injection therapy
in thyroid cystic nodules. Endocr Pract 2004; 10(3):269–275.
107. Guglielmi R, Pacella CM, Bianchini A, et al. Percutaneous ethanol injection treat-
ment in benign thyroid lesions: Role and efficacy. Thyroid 2004; 14(2):125–131.
108. Bennedbaek FN, Hegedus L. Treatment of recurrent thyroid cysts with ethanol:
A randomized double-blind controlled trial. J Clin Endocrinol Metab 2003;
88(12):5773–5777.
109. Papini E, Panunzi C, Pacella CM, et al. Percutaneous ultrasound-guided ethanol
injection: A new treatment of toxic autonomously functioning thyroid nodules?
J Clin Endocrinol Metab 1993; 76(2):411–416.
110. Bennedbaek FN, Nielsen LK, Hegedus L. Effect of percutaneous ethanol injection
therapy versus suppressive doses of L-thyroxine on benign solitary solid cold thyroid
nodules: A randomized trial. J Clin Endocrinol Metab 1998; 83(3):830–835.
111. Lippi F, Manetti L, Rago T. Percutaneous ultrasound-guided ethanol injection for
treatment of autonomous thyroid nodules: Results of a multicentric study abstract.
J Endcrinol Invest 1994; 17(Suppl. 2):71.
112. Zingrillo M, Collura D, Ghiggi MR, et al. Treatment of large cold benign thyroid nod-
ules not eligible for surgery with percutaneous ethanol injection. J Clin Endocrinol
Metab 1998; 83(11):3905–3907.
234 Gharib

113. Zingrillo M, Modoni S, Conte M, et al. Percutaneous ethanol injection plus radioio-
dine versus radioiodine alone in the treatment of large toxic thyroid nodules. J Nucl
Med 2003; 44(2):207–210.
114. Pacella CM, Bizzarri G, Guglielmi R, et al. Thyroid tissue: US-guided percutaneous
interstitial laser ablation: A feasibility study. Radiology 2000; 217(3):673–677.
115. Dossing H, Bennedbaek FN, Karstrup S, et al. Benign solitary solid cold thyroid
nodules: US-guided interstitial laser photocoagulation: Initial experience. Radiology
2002; 225(1):53–57.
116. Papini E, Guglielmi R, Bizzarri G, et al. Ultrasound-guided laser thermal ablation
for treatment of benign thyroid nodules. Endocr Pract 2004; 10(3):276–283.
117. Pacella CM, Bizzarri G, Spiezia S, et al. Thyroid tissue: US-guided percutaneous
laser thermal ablation. Radiology 2004; 232(1):272–280 [Epub May 20, 2004].
118. Dossing H, Bennedbaek FN, Hegedus L. Effect of ultrasound-guided interstitial laser
photocoagulation on benign solitary solid cold thyroid nodules: One versus three
treatments. Thyroid 2006; 16(8):763–768.
119. Fogelfeld L, Wiviott MB, Shore-Freedman E, et al. Recurrence of thyroid nodules
after surgical removal in patients irradiated in childhood for benign conditions.
N Engl J Med 1989; 320(13):835–840.
120. Hurley DL, Gharib H. Evaluation and management of multinodular goiter. Otolaryn-
gol Clin North Am 1996; 29(4):527–540.
121. Huysmans DA, Buijs WC, van de Ven MT, et al. Dosimetry and risk estimates of
radioiodine therapy for large, multinodular goiters. J Nucl Med 1996; 37(12):2072–
2079.
122. Huysmans DA, Hermus AR, Corstens FH, et al. Large, compressive goiters treated
with radioiodine. Ann Intern Med 1994; 121(10):757–762.
123. Huysmans AK, Hermus RM, Edelbroek MA, et al. Autoimmune hyperthyroidism
occurring late after radioiodine treatment for volume reduction of large multinodular
goiters. Thyroid 1997; 7(4):535–539.
124. Huysmans D, Hermus A, Edelbroek M, et al. Radioiodine for nontoxic multinodular
goiter. Thyroid 1997; 7(2):235–239.
125. Wesche MF, Tiel-v-Buul MM, Smits NJ, et al. Reduction in goiter size by 131 I therapy
in patients with non-toxic multinodular goiter. Eur J Endocrinol 1995; 132(1):86–87.
126. Hermus AR, Huysmans DA. Treatment of benign nodular thyroid disease. N Engl J
Med 1998; 338(20):1438–1447.
127. Silva MN, Rubio IG, Knobel M, et al. Treatment of multinodular goiters in elderly
patients with therapeutic doses of radioiodine preceded by stimulation with human
recombinant TSH. Endocr J 2000; 47(Suppl.):144.
128. Duick DS, Baskin HJ. Significance of radioiodine uptake at 72 hours versus
24 hours after pretreatment with recombinant human thyrotropin for enhancement
of radioiodine therapy in patients with symptomatic nontoxic or toxic multinodular
goiter. Endocr Pract 2004; 10(3):253–260.
129. Albino CC, Mesa CC Jr, Olandoski M, et al. Recombinant human thyrotropin as
adjuvant in the treatment of multinodular goiters with radioiodine. J Clin Endocrinol
Metab 2005; 90(5):2775–2780 [Epub Feb 15, 2005].
130. Duick DS, Baskin HJ. Utility of recombinant human thyrotropin for augmentation of
radioiodine uptake and treatment of nontoxic and toxic multinodular goiters. Endocr
Pract 2003; 9(3):204–209.
Thyroid Nodules and Multinodular Goiter 235

131. Nielsen VE, Bonnema SJ, Hegedus L. Transient goiter enlargement after adminis-
tration of 0.3 mg of recombinant human thyrotropin in patients with benign nontoxic
nodular goiter: A randomized, double-blind, crossover trial. J Clin Endocrinol Metab
2006; 91(4):1317–1322 [Epub Jan 24, 2006].
132. Nielsen VE, Bonnema SJ, Boel-Jorgensen H, et al. Stimulation with 0.3-mg recom-
binant human thyrotropin prior to iodine 131 therapy to improve the size reduction of
benign nontoxic nodular goiter: A prospective randomized double-blind trial. Arch
Intern Med 2006; 24(14):1476–1482.
133. Tan GH, Gharib H. Thyroid incidentalomas: Management approaches to nonpal-
pable nodules discovered incidentally on thyroid imaging. Ann Intern Med 1997;
126(3):226–231.
6
Differentiated Thyroid Carcinoma

Jennifer A. Sipos
The Ohio State University, Columbus, Ohio, U.S.A.

Ernest L. Mazzaferri
University of Florida, Gainesville, Florida, U.S.A.

INTRODUCTION
Thyroid carcinoma comprises a spectrum of malignancies ranging from rarely
lethal, slow-growing neoplasms to among the most deadly aggressive cancers to
afflict humanity. Fortunately, most cases are well-differentiated tumors that can
be treated successfully. Still, therapy remains controversial because prospective
clinical studies are confounded by low disease incidence, prolonged disease
time course, and relatively good outcome in the majority of patients. Further,
while staging systems predict outcome reasonably well, they remain inexact.
Thus, while the majority of patients can be reassured that they are likely to do well,
this outcome cannot be absolutely guaranteed. As a result, therapy has remained
relatively dogmatic, as the ideal strategy that prevents unnecessary overtreatment
while avoiding detrimental undertreatment is not a current reality, leading physi-
cians to essentially choose on which side to err. The goals of patient management
are to minimize morbidity and mortality from cancer (tumor recurrence, metas-
tases, and death) as well as from therapy [surgery, hypothyroidism, iodine-131
(131 I) therapy, thyroid-stimulating hormone (TSH) suppression, and cost]. This
chapter emphasizes features that predict outcome and current management
paradigms.

237
238 Sipos and Mazzaferri

Figure 1 Annual incidence of thyroid carcinoma (all types) in the United States according
to age at the time of diagnosis and patient gender. Source: Adapted from Ref. 2.

EPIDEMIOLOGY
Cancers of the thyroid are rare, comprising approximately 2% of all diagnosed
cancers but accounting for ⬎93% of all cancers of the endocrine system (1).
Approximately 30,180 new cases and 1500 thyroid cancer deaths occur each year
in the United States: (1) ranking it 9th among malignancies for women and 19th
for men (2). Thyroid cancer is nearly three times more common in women than
men (1). It may occur at any age, but is more common in middle-aged women and
in men over the age of 60 years (Fig. 1). Thyroid cancer death rates are ⬍10%,
but vary significantly among the various types of thyroid cancer (Table 1) (3).
Data from the National Cancer Institute’s Surveillance, Epidemiology, and
End Results (SEER) program, demonstrate that the incidence of thyroid cancer
has more than doubled over the past three decades, increasing from a rate of
3.6 per 100,000 in 1973 to 8.7 per 100,000 in 2002, a twofold increase that is
statistically significant (p ⬍ 0.001 for trend) (4). New papillary thyroid cancers
Differentiated Thyroid Carcinoma 239

Table 1 Incidence and 10-Year Relative Survival Rates of the


Major Histological Types of Thyroid Carcinoma Among 53,856
Cases Diagnosed Between 1985 and 1995 (3)a

Thyroid carcinoma type Frequency (%) Survival (%)

Papillary 80 93
Follicular 11 85
Hürthle 3 76
Medullary 4 75
Anaplastic 2 14

a Relative survival is death attributed to thyroid carcinoma after correcting


for death from other causes.
Source: Adapted from Ref. 3.

represent virtually all of this increased incidence, while the other three major forms
of thyroid cancer have remained constant (4). The rising incidence has occurred
mostly in women but thyroid cancer is also rising in men, albeit at half the rate
of that in women (5). Of major importance, nearly half (49%) the tumors are
papillary cancers ⬍1 cm [papillary thyroid microcarcinoma (PTMC)] and most
(87%) are papillary tumors ⬍2 cm in largest dimensions (4).
One study reported that in spite of this escalating frequency of small thy-
roid cancers, the overall mortality rate per 100,000 people in the population has
remained constant according to death certificate data from the National Vital
Statistics System (4). The authors of this study attribute the increasing incidence
of papillary thyroid carcinoma to overdiagnosis or “increased diagnostic scrutiny,”
which makes it difficult to know which patients need treatment, suggesting that
small asymptomatic thyroid nodules should be followed up for a period of time
without immediately initiating diagnostic studies.
Other data (6) indicate that thyroid cancer deaths have not remained stable
over the past several decades. The National Cancer Institute’s SEER database (7)
shows that among women, the five-year relative survival rates for thyroid cancer
from 1974 to 2001 increased significantly from 92.7% to 97.4% in 1974 to 2003
(Fig. 2, p ⬍ 0.5). During this same time, however, the rates of thyroid cancer
distant metastases in men at the time of diagnosis were more than twofold more
than those of women (9% vs. 4%), and during 1992 to 2000 the annual percent
change in thyroid cancer mortality significantly increased in men by 2.4% (Fig. 2,
p ⬍ 0.05), the largest increase of any type of cancer (7).

Classification
Thyroid carcinomas are classified into four major types, which, in decreasing
order of frequency, are papillary (PTC), follicular (FTC), medullary (MTC), and
240 Sipos and Mazzaferri

(A)

Figure 2 (A) Papillary thyroid carcinoma (PTC): tumor infiltrating thyroid and invading
thyroid capsule (top left); FNA cytology specimen showing the typical features of PTC,
including large nuclei with inclusion bodies (top right); histology showing typical papillary
fronds (lower left); psammoma bodies showing lamellate appearance, which are virtually
diagnostic of this tumor (lower right), (B). Follicular thyroid carcinoma (FTC): solid
encapsulated tumor with areas of necrosis and invasion of tumor capsule at 3 o’clock (top
left); microscopic tumor invasion of tumor capsule (top right); FNA cytology specimen
showing sheets of follicular cells without colloid that is suspicious of FTC (lower left);
Hürthle cells showing abundant oxyphilic cytoplasm (lower right), (C). Anaplastic thyroid
carcinoma (ATC) and medullary thyroid carcinoma (MTC): ATC with tumor infiltrating
the entire gland and invading the thyroid capsule at 4 o’clock (top left); histology of ATC
showing large, bizarre-appearing cells (top right); MTC with spindle cells and dense stroma
that stains positive for amyloid (lower right); spindle cell tumor that was a MTC, which
can mimic ATC (lower right).

anaplastic thyroid carcinomas (ATC) (Table 1). Although categorized as a separate


entity in some classifications, Oncocytic [Hürthle cell (HTC)] carcinomas are con-
sidered by the World Health Organization classification of tumors (8) as follicular
carcinomas composed exclusively or predominantly (⬎75%) of oncocytic cells,
the ultrastructure of which are filled with mitochondria producing a cytologic
Differentiated Thyroid Carcinoma 241

(B)

(C)

Figure 2 (Continued)
242 Sipos and Mazzaferri

appearance that is unique. PTC, FTC, and HTC—often termed differentiated thy-
roid carcinomas or DTC— arise from follicular cells, synthesize thyroglobulin
(Tg), and tend to be sporadic tumors, although occasionally PTC is familial.
MTC, which originates from thyroidal C cells that secrete calcitonin, may be spo-
radic or familial. ATC usually arises from well-differentiated thyroid carcinomas,
particularly PTC (9).

PAPILLARY AND FOLLICULAR THYROID CARCINOMAS


Papillary Thyroid Carcinoma
This tumor accounts for approximately 80% of all thyroid carcinomas in the United
States (3,5). It is three times more frequent in women than in men. Incidence is
highest in women in midlife, but it occurs at all ages. Often considered a separate
disease, occult microscopic PTC (≤1 cm) is found in 10% or more of autopsy
and surgical thyroid specimens obtained from men and women throughout adult
life (5).
A small group of nonmedullary thyroid cancers appear to be inherited, most
of which are PTC. Approximately 5% of PTCs are familial tumors inherited as
an autosomal dominant trait without other associated pathology. Familial PTC
is characterized by an earlier age of onset and more aggressive phenotype (10)
than sporadic PTC. One study of 258 cases (11) found that although patient
gender, age, and tumor histology were similar to that of sporadic PTC, familial
tumors were more likely to have intrathyroidal dissemination (41% vs. 29%) and
higher recurrence rates (16% vs. 10%) than sporadic tumors without displaying
significant differences in size, local invasion, or macroscopic metastasis.

Follicular Thyroid Carcinoma


This tumor occurs sporadically and accounts for approximately 10% of all thyroid
cancers in the United States, although it is more common in countries with low
dietary iodine (12). It usually occurs at a slightly older age than PTC (13,14), but
in some studies almost half of the patients are older than 40 years at diagnosis (15).
This tumor is rare in children, occurs infrequently after head and neck irradiation,
and is not commonly found incidentally.

RADIATION-INDUCED THYROID CARCINOMA


Epidemiology
Exposure to ionizing radiation during childhood is the best understood cause of
papillary, and less commonly, follicular thyroid carcinoma. Nevertheless, such a
history usually is now elicited in only approximately 5% of patients (16). Some
radiation exposure events have been associated with an increased risk of thyroid
cancer, while others have not. Possible explanations of this difference may include
Differentiated Thyroid Carcinoma 243

the various sources of radiation involved, which delivered different thyroid doses
and different thyroid dose rates. Treatment of benign conditions of the head and
neck, such as acne, tonsillitis, and sinusitis, with radiation therapy was common-
place in the 1940s and 1950s. Fortunately, such treatments are no longer used, but
their carcinogenic effects are still seen more than 40 years later (17). Similarly,
survivors of childhood malignancies treated with radiation to the head/neck and
chest are also at increased risk of thyroid carcinoma (18).
Factors that increase the risk of developing thyroid cancer after external
beam radiation therapy are female gender, radiation for childhood cancer (as
opposed to benign conditions), and family history of thyroid cancer (19).

PAPILLARY THYROID CARCINOMA


Pathology
Papillary thyroid carcinoma (PTC) is typically an unencapsulated invasive tumor
with ill-defined margins. In approximately 10% of cases, the thyroid capsule is
penetrated by tumor that may invade surrounding tissues, while another 10% are
fully encapsulated (20,21). The tumor is typically firm and solid but may develop
chronic hemorrhagic necrosis, making it soft, and yielding a thick brownish fluid
on needle biopsy that may cause it to be mistaken for a benign cyst (22). In
rare cases, the entire thyroid is effaced with large cystic PTC tumors (23). Small
tumors ≤1.0 cm in diameter often have a stellate appearance and are usually found
by serendipity. Although they generally pose no risk to the patient, PTMCs are
occasionally locally invasive and metastatic (24).
Most PTCs have a typical microscopic appearance with complex, branching
papillae, and a fibrovascular core covered by a single layer of tumor cells intermin-
gled with follicular structures, but some PTCs have a pure follicular appearance
that grossly resembles FTC (8). The term mixed papillary–follicular carcinoma
has no clinical value because the follicular component does not alter 131 I uptake
or prognosis, thus such tumors are considered to be classic PTC. Nuclear features
are more important than architectural appearance in establishing a PTC diagnosis.
Tumors with a pure microfollicular pattern with virtually no papillary structures
but with typical cellular features of PTC are termed follicular variant of PTC
(FVPC) (8).
The cellular features of PTC distinguish it from other tumors, regardless
of its architecture, permitting an accurate diagnosis by fine-needle aspiration
(FNA) cytology [Fig. 2(A)]. The large cells contain pink to amphophilic finely
granular cytoplasm and large pale nuclei with inclusion bodies, sometimes called
“orphan Annie eye” nuclei, and nuclear grooves that identify it as PTC [Fig. 2(A)].
Psammoma bodies—the “ghosts” of infarcted papillae that are virtually pathog-
nomonic of PTC—are calcified, concentric lamellated spheres found in about half
the cases [Fig. 2(A)] (25). Multiple histological PTC subtypes or variants have
been described (Table 2) (8).
244 Sipos and Mazzaferri

Table 2 Prognosis of Main Histological Variantsa of Papillary and Follicular Thyroid


Carcinoma

Better Worse Possibly worse

Papillary thyroid carcinoma


Encapsulated Tall cell
Cystic Columnar cell Solid
Microcarcinoma Diffuse sclerosis Oncocytic (Hürthle cell)
Macrocarcinoma Diffuse follicular Associated with Graves’
disease
Insular cell
PTC with dedifferentiation
Follicular thyroid carcinoma
Oncocytic (Hürthle cell)
Insular cell
a Other histological variants—such as PTC with lipomatous stroma or fascitis-like stroma or myxoid

and cribriform variant PTC—have been reported too rarely to be certain about prognosis.
Source: Adapted from Ref. 23.

Multiple tumor foci are found in up to 80% of PTC cases when the thyroid
gland is examined in detail but are found in only 20% to 45% of specimens
examined routinely (26). This is important because multifocal PTMCs are
associated with an increased rate of lymph node and distant metastases. One study
(27) found lymph node metastases in 20% of unifocal and 35% of multifocal PTC
tumors and that multifocality increased lymph node recurrences almost sixfold.
Another study of PTMC found that 89% of distant metastases from these small
tumors were associated with multifocal tumors (28). Thought to be intrathyroidal
metastases in the past (26), newer studies using X-chromosome (29) BRAF
(the gene encoding B-type Rat Kinase) (30,31) or PTC/RET (RET is rearranged
during transfection proto-oncogene) analyses (32) find that many of tumors
within the same thyroid have different oncogene patterns that support the concept
that multicentric tumors arise de novo as independent tumors. Still, one study (33)
found that multifocal PTC tumors often have similar X-chromosome inactivation
patterns suggesting intrathyroidal metastases are a cause of multicentric tumor.
Thus, both intrathyroid metastasis and de novo tumors may play an important
role in the intrathyroidal spread of PTC, findings that have important therapeutic,
diagnostic, and prognostic implications.
Focal areas of infiltration with lymphocytes or plasma cells or classic
Hashimoto’s disease are usually present in or around the tumor and are intense
in up to 20% of cases (34). The presence of neoplastic cell phagocytosis by
macrophages and lymphocytic reaction has been associated with a more favorable
prognosis and reduced likelihood of distant metastases (35). It has been hypothe-
sized that this lymphocytic infiltration represents an immune reaction which helps
Differentiated Thyroid Carcinoma 245

to control tumor growth and proliferation (36). Lymphocytic infiltration has been
associated with a more favorable clinical outcome (34).

Lymph Node Metastases


PTC metastasizes to lymph nodes in the lateral and central neck and mediastinum.
Lymph node metastases are found in almost half the cases at the time of diagnosis
(37), while even more—up to 85% in some studies (38,39)—have microscopic
nodal metastases found on more careful histological study (39). The number
and size of lymph node metastases increase as the primary tumor size enlarges
beyond 5 mm (14). When the isthmus or both lobes are involved with tumor,
nodal metastases are often bilateral or extend into the mediastinum, with the most
common site being the lower paratracheal area (level VI) (40,41). Level VI is the
most common site of lymph node recurrence (42). Lymph node metastases have
variably been found to be of no prognostic importance in some studies (43) and
of significant prognostic importance in others (37), but a study from the SEER
database of 19,919 patients found with multivariate analysis that lymph node
metastases predicted a poor outcome. In some cases tumor penetrates the lymph
node capsule and invades the soft tissues, which is a particularly poor prognostic
sign (44).

Distant Metastases
Less than 5% of patients have distant metastases at the time of diagnosis and
another 5% develop them over the next two or three decades (12). The lung is the
most common site of distant metastases and is the most common disease-specific
cause of death from PTC. In a review (12) of 1231 patients with distant metastases,
49% were in the lung alone and another 15% were in lung and bone, 25% were
in bone alone, and 12% were in the central nervous system (CNS) or in multiple
organs (12). Lung metastases may be large and discrete [Fig. 3(A)] or may have a
“snow-flake” appearance from diffuse small metastases [Fig. 3(B)].
The lungs may concentrate sufficient 131 I to be detected on whole-body
scan. Some lung metastases are not seen on radiographs but are visible only on
131
I whole-body scans, sometimes only after administration of therapeutic amounts
of 131 I [Fig. 3(C) and 3(D)] (45,46). In other cases, the lungs do not concentrate
131
I for a variety of reasons.

Papillary Microcarcinoma
PTMC is a tumor ≤1.0 cm in diameter that is usually impalpable. Histologically
malignant, PTMC is generally found by coincidence during surgery for benign
thyroid disease and generally displays a benign clinical behavior. Still, not all
PTMCs are innocuous. Depending upon the study, up to 43% are multifocal, as
many as 69% have lymph node metastases, and up to 2.8% have distant metastases
(24). Recurrence rates vary. One study found a recurrence rate of 25.8% (47), but
246 Sipos and Mazzaferri

Figure 3 X-ray with dense lung infiltrates from PTC (A); X-ray with multiple faint
pulmonary metastases and mediastinal metastasis from PTC (B); metastases from thyroid
carcinoma: normal chest X-ray in a patient with a serum Tg of 40 ng/mL after neck surgery
for PTC (C); scan of same patient 48 hours after 100 mCi 131 I (D); X-ray with multiple
pelvic metastases from FTC (E); scan of same patient after 200 mCi 131 I (F).

on average the rate is about 5% (24). Mortality from PTMC is rare, but rates as
high as 2% are found in some studies (48). Multifocal tumors tend to be more
aggressive than unifocal PTMCs (24,27,28,31). Lung metastases from PTMC are
rare but tend to occur more often in tumors with bulky cervical metastases (28,37).
Otherwise, the recurrence and cancer-specific mortality rates of PTMC are near
zero (24).
Differentiated Thyroid Carcinoma 247

Papillary Cancer within a Thyroglossal Duct


PTC within a thyroglossal duct is a rare occurrence but it is usually small
(⬍1 cm) and follows a benign course (49). Treatment is controversial, but a
Sistrunk procedure for removal of the thyroglossal duct is the mainstay (50,51).
Some also advocate thyroidectomy followed by radioiodine ablation as PTC may
be found within the thyroid gland upon careful inspection (52–55). PTC within a
thyroglossal duct is almost always small and usually has a benign course.
Encapsulated Papillary Carcinoma
This variant comprises approximately 10% of PTCs. Completely surrounded by a
fibrous capsule but otherwise a typical PTC, it is about half as likely as usual
to metastasize. It rarely recurs after initial therapy, and almost never causes
death (20).
Follicular Variant Papillary Carcinoma
Classic PTC comprises 55% to 65% (56) and FVPC approximately 20% to 40%
of all PTCs (57,58). Opinions differ concerning the exact diagnostic criteria for
FVPTC, resulting in the wide prevalence rates among studies. The diagnosis of
classic FTC requires the presence of capsular and/or vascular invasion (8), whereas
the diagnosis of FVPTC depends upon finding nuclear features of PTC. A major
diagnostic feature of FVPTC, according to the original description (59) and that
of LiVolsi (60), is that the microfollicular structure must involve the entire tumor;
however, some find this too restrictive and make the diagnosis of FVPTC when
80% of the tumor contain microfollicles (56). This distinction is important since
40% (61) to 80% (56) of classic PTCs show areas of follicle formation, thus
accounting for the wide range in incidence rates reported for FVPTC. Although
there are reports (62,63) that the rates of lung metastases are higher in FVPTC than
those in classic PTC, others have been unable to confirm this (56,58,64), and the
literature generally reflects that the clinical outcome and behavior of FVPTC and
classic PTC are the same (56,57,61). Nonetheless, a few case reports of FVPTC
describe metastases to lung (62,63), kidney (65), bone (66), and skin (67).
Macrofollicular Variant
Macrofollicular PTC is a rare variant of PTC in which ⬎50% of the follicles are
macrofollicles. The presence of abundant colloid, macrophages, macrofollicular
architectural arrangement and absence of typical cytologic features of PTC can
lead to an erroneous diagnosis of benign adenoma or hyperplastic nodule (68). As
a result, the tumor may be extremely difficult to diagnose by FNA because the
characteristic features of PTC are not present (69).
Diffuse Follicular Variant Papillary Carcinoma
This uncommon tumor may be confused with typical multinodular goiter or macro-
follicular adenoma on frozen section (70). It is seen mainly in younger women with
248 Sipos and Mazzaferri

goiter, about one-third of whom have hyperthyroidism. These aggressive tumors


are more likely to be multicentric, have extrathyroidal extension, nodal and distant
metastases, and vascular invasion compared to common PTC and FVPTC (71).
Mortality rates are very high with this tumor (70).

Tall Cell Variant


Ten percent of PTCs have papillae with cells twice as tall as they are wide that
constitute at least 30% of the tumor (70). Compared with typical PTC, tall cell
variants tend to be diagnosed about two decades later (in patients in their mid-
50s), are larger, and are more often associated with invasion into local soft tissues
and with distant metastases (70,72,73). This tumor can be identified by FNA
cytology. It often expresses the p53 oncogene, BRAF mutation, RET/PTC gene
rearrangement, or NTRK1 mutation (74,75). Tall cell variant often loses or lacks
131
I uptake and the mortality is two- to threefold higher than those of typical PTC
(70,73).

Columnar Cell Variant


This rare variant, which is possibly related to tall cell carcinoma, is found mainly
in males and is composed of rectangular cells with clear cytoplasm (37). Dis-
tant metastases develop in 90% and are usually unresponsive to 131 I therapy or
chemotherapy, resulting in death in most patients (37,38). When it is encapsulated,
it has a much better prognosis (28).

Diffuse Sclerosing Variant


Approximately 5% of spontaneously occurring PTC and 10% of those found
among the Chernobyl children are of this type (70,76). The tumor is usually
bilateral and presents as a goiter with extensive squamous metaplasia, sclerosis,
psammoma bodies, and abundant lymphatic invasion involving the whole thyroid
gland. Almost all develop lymph node metastases and approximately 25% have
lung metastases (70). FNA cytology reveals squamous metaplasia, inflammatory
cells, and psammoma bodies, but this tumor may be difficult to differentiate from
thyroiditis (77). Although local and pulmonary metastases are more frequent than
usual, there is disagreement about whether its long-term prognosis is worse than
that of typical PTC (78–80).

Solid or Trabecular Variant


This tumor has a predominantly (⬎75%) solid architectural pattern but maintains
the typical nuclear features of PTC. It has a propensity for extrathyroidal spread
and lung metastases that impart a poor prognosis (70), but some find it to be
more common in children, in whom its prognosis is the same as that of typical
PTC (81).
Differentiated Thyroid Carcinoma 249

Oxyphilic (Hürthle Cell) Variant


Approximately 2% of PTCs have cellular features resembling those of Hürthle cell
(oxyphilic) FTCs (82). Some cases have multiple oxyphilic thyroid tumors and a
familial occurrence (83). This tumor cannot be identified as PTC by FNA cytology
but is recognized by its papillary architecture on the final histological sections.
Compared with typical PTC, oxyphilic PTC has fewer neck nodal metastases
at diagnosis but has higher recurrence and mortality rates and in this respect
resembles oxyphilic FTC (70).

Insular Carcinoma
Approximately 5% of all thyroid carcinomas show solid clusters of cells with small
follicles that resemble pancreatic islet cells but contain Tg. Often categorized as a
FTC variant, some tumors show papillary differentiation. LiVolsi believes that this
tumor should be considered a separate entity derived from follicular epithelium
(84). The tumors are unusually large and invasive and tend to grow through the
tumor capsule and into tumor blood vessels. Compared with PTC, insular carci-
noma presents at an older age (54 vs. 36 years) with larger tumors (4.7 vs. 2.5 cm),
fewer neck metastases (36% vs. 50%) but more distant metastases (26% vs. 2%),
and has a worse 30-year cancer-specific mortality rate (25% vs. 8%) (73). Insular
carcinoma also displays aggressive behavior in children but is usually responsive
to thyroidectomy and 131 I therapy (85).

FOLLICULAR THYROID CARCINOMA


Pathology
Follicular thyroid carcinomas (FTCs) are solid invasive tumors which, unlike
PTCs, do not show necrotic degeneration but tend to be solitary and encapsulated,
even when they are biologically aggressive (5). Minimally invasive tumors show
just enough evidence—usually invasion into the tumor capsule without penetration
through it (and no vascular invasion)—to make a diagnosis of carcinoma. Others
have multiple foci of tumor capsule and vascular penetration but remain fairly
discrete masses, while a few are highly invasive tumors with satellite nodules.
Large, aggressive tumors may extend into the opposite lobe or adjacent cervical
tissue. Most FTCs cannot be differentiated from follicular adenomas on gross
tumor inspection, FNA, or frozen section and must be differentiated from follicular
variant PTC and follicular adenoma based on final histological sections (12).
FTCs typically are compact, highly cellular tumors composed of microfolli-
cles, trabeculae, and solid masses of cells. Less often, they contain medium-sized
or large follicles and such low invasive characteristics that are difficult to differ-
entiate from benign adenomas even on the final histological sections—a finding
associated with an excellent prognosis (12).
250 Sipos and Mazzaferri

FTC has compact cells with small, dark-staining, round nuclei that are more
uniform in shape, size, and location than the nuclei of PTCs and are difficult to
identify as carcinoma by FNA [Fig. 2(B)] (12,86).
Hürthle Cell Carcinoma
Oxyphilic cells, termed Hürthle or Askanazy cells, contain increased amounts
of acidophilic cytoplasm that contains numerous mitochondria on electron
microscopy. Greater than 75% of cells must be Hürthle cells to constitute HTC.
Some consider this to be a distinct clinicopathological entity while others consider
it to be a variant of FTC (87). HTCs demonstrate lymph node metastases more
frequently than typical FTC and metastases rarely respond to 131 I therapy. Regard-
less of its classification, HTC has a less favorable prognosis than nonoxyphilic
FTC, even though the former may not initially appear less differentiated or more
invasive (12).
Lymph Node Metastases
Compared to PTC, FTC metastasizes about half as often to regional lymph nodes,
occurring in approximately 20% of cases and is usually caused by the more
aggressive tumors that often have distant metastases (14).
Distant Metastases
FTC tends to metastasize to lung, bone, CNS, and other soft tissues with greater
frequency than does PTC, and the metastases often avidly concentrate 131 I. Distant
metastases at the time of diagnosis are about twice as common with FTC compared
to PTC (5). Unlike small PTCs that rarely metastasize to distant sites, small FTCs
can metastasize widely, and tumors ⬎3 cm are associated with a much higher
mortality rate (14,88)

DIAGNOSIS OF PAPILLARY AND FOLLICULAR


THYROID CARCINOMAS
Clinical Presentation
In the past, PTC was often diagnosed at a late stage, when the tumor was large
and invasive. Now, however, most are identified earlier by ultrasound-guided FNA
of small, asymptomatic thyroid nodules—frequently discovered incidentally by
imaging of the neck for other medical conditions (89). A few come to attention
as the result of pain, hoarseness, dysphagia, hemoptysis, or other signs of tis-
sue infiltration or rapid tumor growth. Such findings are associated with a high
probability of carcinoma in a nodule (90) and greater than usual mortality rates
(37). Sometimes palpably enlarged cervical lymph nodes are the only clue to the
diagnosis. A history of exposure to head or neck radiation is important, but only
about one-third of such nodules are malignant (17,91). A nodule in the setting
of a family history of PTC should be regarded with higher than usual suspicion.
Differentiated Thyroid Carcinoma 251

Nevertheless, the evaluation of familial or radiation-induced tumors is similar to


that of sporadic nodules. Most FTCs present as an asymptomatic neck mass with-
out palpable cervical lymphadenopathy. Less often, a distant metastasis is the first
manifestation, appearing as lung nodules, osteolytic bone lesion, or pathological
fracture or as a CNS tumor with neurological sequelae. Rarely, distant metastases
are seen in the absence of a palpable thyroid lesion (24). Bulky metastatic lesions
may be functional and cause thyrotoxicosis.
DTC usually becomes manifest by a palpably firm thyroid nodule that moves
upward when the patient swallows. PTC, however, may be cystic and soft or may
diffusely infiltrate one lobe or the entire thyroid gland; or its first manifestation
may be a palpable cervical lymph node metastasis. A midline mass above the
thyroid isthmus may be a metastatic (Delphian) lymph node or carcinoma within
a thyroglossal duct—the latter is suggested by upward movement with tongue
protrusion. Distant metastases are found less often at the time of diagnosis of
PTC than FTC, but when they are present, the primary tumor is usually large and
invasive. Only 10% of thyroid nodules have clear clinical evidence of malignancy
such as vocal cord paralysis or signs of invasion or metastases at the time of
diagnosis. Most appear benign and are associated with few or no symptoms. Thus
neither the history nor physical examination offers enough evidence of a nodule’s
benign nature that further testing can be deferred.

Fine-needle Aspiration
A serum thyrotropin (TSH) level should be measured in every patient found
to have a thyroid nodule. If the TSH level is below the lower reference range
(Fig. 4), then radionuclide imaging should be performed to determine if the nodule
is hyperfunctional (hot). Such a nodule is highly unlikely to be malignant and does
not routinely require FNA (92,93). Otherwise, FNA should be done to evaluate
a thyroid nodule in a clinically euthyroid patient (Fig. 4), whether the patient has
a single nodule or a multinodular goiter (92,93). Other tests—especially imaging
studies—are otherwise too nonspecific to be used early in the evaluation, except
perhaps in a patient with multinodular goiter and a suppressed TSH. All thyroid
nodules ⬎1 to 1.5 cm should undergo FNA (92), especially those with suspicious
ultrasound features. Fine-needle aspiration is best performed under ultrasound
guidance with on-site evaluation of cytology adequacy. This will decrease the
likelihood of obtaining a cytology specimen that is insufficient for diagnosis,
a common cause of missed cancer diagnoses, compared with performing the
procedure by palpation guidance alone. The ultrasonography-cytology assessment
approach is particularly important if the nodule is cystic (94). Thyroxine therapy
should not be used as a diagnostic test to identify thyroid carcinoma. Some
malignant nodules appear by palpation to shrink as the perinodular thyroid
parenchyma reduces in size in response to levothyroxine (86); in other cases,
ultrasound documented of reduction in size of malignant nodules has been
demonstrated (95).
252 Sipos and Mazzaferri

Figure 4 Diagnostic paradigm for evaluating a thyroid nodule. Source: Adapted from
Ref. 87.

FNA yields cytology that can be categorized as malignant, benign, indeter-


minate (suspicious), or inadequate for diagnosis (Fig. 4). FNA is highly effective in
obtaining sufficient cytology to identify the distinctive features of PTC, including
most of its variants as well as most other malignant lesions (86).
Benign Hürthle cell tumors and follicular adenomas may be difficult to dif-
ferentiate from their malignant counterparts by FNA and frozen tissue sections.
Large-needle aspiration biopsies and cutting-needle biopsies usually yield results
similar to those of FNA but cause more complications and have lower diagnostic
specificity, especially for PTC, than FNA (96). FNA cytology specimens showing
normal or atypical follicular or Hürthle cells are often simply designated as follic-
ular or Hürthle cell neoplasms because their benign or malignant character cannot
Differentiated Thyroid Carcinoma 253

be determined with certainty until the final histological sections are available.
Even then there may be difficulty separating malignant and benign tumors (86).
Cytological material sufficient for diagnosis can be obtained in most palpa-
ble nodules, but the accuracy of FNA is enhanced by ultrasonographically guided
FNA, especially for small solid, cystic, and hypoechoic nodules (94,97). Patients
with nodules that yield malignant cytology should undergo total or near-total
thyroidectomy (Fig. 4) (86). Before performing surgery on nodules with inde-
terminate cytology (highly cellular specimens with normal or atypical follicular
cells without colloid), an 123 I thyroid scan should be done even in patients without
suppressed TSH levels to identify those occasional hyperfunctioning nodules
that are not active enough to cause TSH suppression, and which are usually
benign (93).

Evaluating Patients with a History of Head-and-Neck Irradiation


Children and young adults with a history of exposure to ionizing radiation such as
X-ray or external beam radiation therapy, and patients irradiated during childhood
for tumors or Hodgkin’s disease often develop hypothyroidism (98,99) and thyroid
neoplasia (98–100). As a result, measurement of serum TSH is recommended,
along with an FNA for those with a palpable thyroid nodule. However, controversy
exists regarding the diagnostic approach to an asymptomatic previously irradiated
person with a palpably normal thyroid gland. Schneider et al. (101) found that
87% of 54 patients who had been exposed to head-and-neck irradiation (X-ray
treatment for acne as a child or young adult) had one or more discrete ultrasound-
detected thyroid nodules. Of this group, 25% had nodules ≥1 cm, and about half of
the nodules 1.5 cm or larger were impalpable. This study confirmed that irradiated
individuals continue to develop thyroid nodules more than 20 years’ follow-up
from the time of irradiation (101), prompting the authors to advise indefinite
routine ultrasonography every 3 to 5 years in such patients. Still, others argue
that routine ultrasonography is too sensitive; it identifies benign thyroid nodules
in about half of the healthy middle-aged population and in most people exposed
to thyroid irradiation (101–103). Still, most authorities agree that the majority of
thyroid nodules ⬍1 cm discovered by serendipity and without ultrasonographically
suspicious characteristics are benign and can be followed by ultrasonography for
one to two years without FNA, providing the nodule volume grows no more than
50% (104) and the patient understands the small risk imposed by this approach.

FACTORS INFLUENCING PROGNOSIS AND AFFECTING OUTCOME


The prognosis of DTC is determined by an interaction of three clinical variables—
tumor stage, patient age, and therapy—and ranges from excellent to dismal. Over-
all mortality rates for DTC are low (⬍10% over three decades), but recurrence rates
are high. Distant metastases or local recurrences are often detected years after ini-
tial therapy (Fig. 5) (37) unless modern follow-up paradigms are employed (105).
254 Sipos and Mazzaferri

Figure 5 Cancer-specific mortality rates and recurrence rates following initial therapy of
DTC (mean and standard error). Source: Adapted from Ref. 36.

Patient Variables Influencing Prognosis


Age over 40 years at the time of initial therapy is the most important adverse patient
prognostic factor, which becomes progressively worse thereafter, increasing at a
particularly steep rate after the age of 60 years (Fig. 6) (12,37).
The best responses to therapy are in younger patients whose tumors con-
centrate 131 I (45,46,106). Although survival rates are most favorable in children,
their tumors are typically more advanced at the time of diagnosis, with more
local and distant metastases than found in adults (45,46). Up to 80% of children
develop cervical lymph node metastases and 15% to 20% develop pulmonary
metastases (106,107), rates that are almost twofold than those in adults. Also,
tumor recurrence rates over several decades are approximately 40% in children
compared with 20% in adults (37). Yet the prognosis for survival in children is
excellent, with or without a history of irradiation, except for those younger than
10 years, who have higher mortality rates (106,108). Very young children may
have unusual tumors that lack typical papillary architecture and behave more
aggressively (109).
Gender is an independent prognostic factor for survival (37). Thyroid cancer
recurrence and mortality rates are higher in men than in women (72,110). Ten-year
cancer-specific mortality rates for PTC among men and women older than 40 years
Differentiated Thyroid Carcinoma 255

Figure 6 Cancer mortality and recurrence rates according to age at the time of diagnosis.
See Figure 5 for legend. Source: Adapted from Ref. 36.

are 13% and 7%, respectively (p ⬍ 0.01). Although estrogen and progesterone
receptors are expressed in up to 50% of PTCs, this does not explain the risk
imposed by male gender. Compared with women at the time of diagnosis, men
have higher rates of extrathyroidal tumor (51% vs. 39%), including more regional
metastases (40% vs. 32%) and twice the rate of distant metastases (9% vs. 4%).
This is most likely because of late diagnosis of thyroid cancer in men, which best
explains their higher cancer-mortality rates (7).
Serum from patients with Graves’ disease contains thyroid-stimulating
immunoglobulin that stimulates thyroid follicular cells in vitro that can produce
progression of thyroid carcinoma (111). One study of PTC associated with Graves’
disease found that the tumors were more often multifocal and the rate of distant
metastases was three times higher than usual (112). Other studies have failed to
show this effect (113,114).

Tumor Variables Influencing Prognosis


Tumor histology is a major determinant of outcome, being best with PTC, inter-
mediate with FTC, and worst with HTC. Distant metastases at the time of initial
diagnosis are found in 2.2%, 5.3%, and 35% of patients with PTC, FTC and
HTC, respectively (37). Tumor recurrence at distant sites occurs most often with
256 Sipos and Mazzaferri

HTC, is intermediate with FTC, and occurs least often with PTC. Recurrence is
particularly high with very invasive tumors, and becomes increasingly worse as
primary tumors grow ⬎5 mm (12,14,88,109,115,116). Mortality rates follow this
pattern. A study of 5925 patients treated in the USA between 1985 and 1995 found
10-year relative survival rates of 7%, 15%, and 25% for PTC, FTC, and HTC,
respectively (3). Marked cellular atypia or frank anaplastic transformation also
worsens prognosis.
Tumor size has an important influence on outcome. The primary tumors
with FTC tend to be larger and discovered at an older age than PTC (12,14,37).
Primary tumors ⬍1 cm in diameter rarely cause death (117), whereas larger tumors
are associated with higher mortality rates (12,37). One review (37) found that the
rate of distant metastases from both PTC and FTC was 4% with primary tumors
⬍1.5 cm, 10% with tumors 1.5 to 4.4 cm and 17% with tumors 4.5 cm or larger.
Thirty-year cancer-specific mortality rates in the three size groups were 0.5%, 8%,
and 22%, respectively.
Tumor multifocality may affect prognosis. PTC in one thyroid lobe predicts
a tumor rate of approximately 45% in the contralateral lobe in patients undergoing
completion thyroidectomy (118,119). This is one of the main reasons why the rates
of tumor recurrence or locally persistent disease are significantly higher following
hemithyroidectomy (120,121). For example, the recurrence rate in the contralateral
lobe was approximately 17% in a study (122) of 35 patients who had initially
refused completion thyroidectomy after hemithyroidectomy, a recurrence rate that
increases over prolonged periods of time (37). A study of 700 patients (123) found a
1.7-fold risk of recurrence in multifocal compared with unifocal tumors. Still, some
report almost no recurrences in the unresected thyroid lobe (124,125). However,
a number of studies find otherwise, even with PTMC. One study (28) of PTMC
found that only two parameters significantly influenced recurrence: the number
of histologic foci (p ⬍ 0.002) and the extent of initial thyroid surgery (p ⬍ 0.01).
Another study of PTMC (27) found that three parameters significantly influencing
lymph node recurrence: cervical lymph node metastases at presentation, multifocal
disease, and the absence of 131 I ablation.

Lymph Node Metastases


Cervical PTC lymph node metastases (12) reflect aggressive tumor behavior and
correlate with primary tumor size and multicentricity (72). The rate of lymph node
metastases depends upon the tumor features, the extent of surgery and the age of
the patient. A review (12) of 13 studies comprising 7845 PTC cases reported
before ultrasonography was widely employed, found a 36% rate of lymph node
metastases at the time of initial surgery. When neck ultrasonography is performed
preoperatively, a slightly larger number of patients are found to have lymph node
metastases (126–128). With immunohistochemical staining to identify micro-
scopic tumor, 53% have lymph node metastases (39) and with routine bilateral
cervical lymph node dissection, 60% of adults have them (40). The rates at the
Differentiated Thyroid Carcinoma 257

time of diagnosis are even higher in children, ranging as high as 65% (129) to
90% (130).
Although some report that metastatic lymph nodes have no effect on recur-
rence or survival (125,131), a number of studies find an increased risk for local
tumor recurrence when cervical lymph node metastases are present at the time of
initial surgery (37,132). Higher than usual cancer-specific mortality rates are seen
with bilateral cervical or mediastinal lymph node metastases or when tumor pen-
etrates the lymph node capsule and invades surrounding tissues (37,44,133,134).
FTC is less often metastatic to regional lymph nodes, but when it occurs the
prognosis is less favorable (12).

Thyroid Capsular Invasion and Extrathyroidal Extension


Up to one-third of PTCs penetrate the thyroid capsule, which may result in deep
tissue invasion, including tracheal or spinal cord invasion and penetration of the
major vessels (135). When this occurs, the mortality rate is approximately 20% at
five years, a rate 10-fold greater than that of noninvasive PTC (12,37). Aggressive
FTCs may also invade local tissues.

Distant Metastases
The main cause of death from DTC, distant metastases were associated with a
five-year mortality rate of 47% among 1231 patients with metastatic PTC or FTC
(12). A more recent study (136) of 49 patients with distant metastases found
that after a median follow-up of 3.5 years, 49% died of cancer; 3-year and 5-
year actuarial survivals were 69% and 50%, respectively. Children and young
adults with pulmonary metastases have a more favorable prognosis when their
distant metastases are discovered early, are small, and concentrate 131 I (137,138).
For example, 10-year survival rates in one study were approximately 20% in
adults with macronodular lung or bone metastases compared with approximately
80% in children and young adults with micronodular pulmonary metastases that
concentrate131 I (139).
The most recent long-term follow-up study (45) of 444 patients with distant
metastases from PTC and FTC treated from 1953 to 1994 found that 50% had
only lung metastases, 26% had only bone metastases, 18% had both lung and bone
metastases, and 5% had metastases at other sites. Overall 10-year survival after
131
I therapy was 92% in patients who achieved a negative posttreatment 131 I scan
and 19% in those who did not. Treatment was highly effective in younger patients
with 131 I uptake in the tumor and with small metastases. Thus the most important
elements in prolonging disease-free survival and improving the survival rate are
early diagnosis before the metastases are apparent on chest roentgenograms or
diagnostic 131 I whole-body scans (Fig. 3) and early 131 I treatment of tumors,
particularly lung metastases that concentrate the isotope.
258 Sipos and Mazzaferri

Irradiation-Induced Papillary Thyroid Carcinoma


The increase in thyroid carcinoma in post-Chernobyl children has been largely
confined to a specific subtype of papillary carcinoma (solid/follicular) that displays
more aggressive behavior with advanced primary tumor size and more multifocal-
ity and lymph node metastases than usual (140,141). From 44% to 70% express
RET/PTC3 mutations, depending upon how the specimens are studied (140–143).
Also, RET/PTC rearrangements are found in approximately 53% of benign tumors
in the Chernobyl children (144).
Cancer mortality rates are typically similar in other forms of radiation-
induced PTC compared with spontaneously occurring tumors, although tumors
associated with radiation are often large, multicentric and regionally metastatic
with high recurrence rates (91,145).

Other Tumor Factors


The histological variants of PTC and FTC affect prognosis (70). Also, coexistent
Hashimoto’s thyroiditis (usually with papillary thyroid carcinoma) is associated
with a low tumor stage and may be an independent predictor of a favorable prog-
nosis (34,146). Anaplastic tumor transformation that occurs in well-differentiated
thyroid carcinoma dramatically alters its course and results in aggressive local
tumor invasion and widespread, rapidly fatal metastases that do not concentrate
131
I (72).

Oncogenes
Protein tyrosine kinases (TKs) are enzymes that catalyze the transfer of phos-
phate from adenosine triphosphate (ATP) to tyrosine residues in polypeptides. The
human genome contains 90 TKs and 43 TK-like genes (147). The importance of
TKs became apparent when imatinib mesylate, an inhibitor of the BCR-ABL TK,
was shown to have a dramatic effect in the treatment of chronic myeloid leukemia.
It is now widely recognized that TKs play an important role in thyroid cancer and
are now the source of important targets for thyroid cancer chemotherapy.
The BRAF V600E mutation is the most common oncogene in sporadic PTC
in adults (148). The incidence of this mutation in PTC varies with the geographical
location of the study. In the Ukraine and Belarus it can be seen in 22.9% of tumors,
while the incidence is as high as 62% in France and Brazil. Pooled data analysis
of all patients studied reveals an overall incidence of 39.6% (149). The clinical
significance of this mutation has not yet been elucidated. One multicenter study
found an association between the presence of this mutation and lymph node
metastases, extrathyroidal invasion, and advanced tumor stage at initial surgery
(150). The BRAF mutation was also found to be an independent predictor of
tumor recurrence regardless of tumor stage, and was associated with absence of
131
I tumor avidity and treatment failure in patients with recurrent disease (150). A
recent study in Italy, however, found no correlation among the presence of BRAF
and tumor multicentricity, lymph node metastases, stage at diagnosis, or outcome
Differentiated Thyroid Carcinoma 259

(151). This discordance in findings between these two studies may be attributable
to geographic differences in the patient populations studied.

Treatment Variables Influencing Prognosis


Treatment of DTC has a major impact on long-term outcome, as discussed later
in sections dealing with surgical and medical therapy. Delay in therapy is an
important problem. We found that the median time from the first manifestation of
thyroid cancer—nearly always a neck mass—to initial therapy was 4 months in
patients who survived and 18 months in those who died of thyroid cancer (p ⬍
0.001) (37). The 30-year cancer mortality rate was nearly doubled when therapy
was delayed longer than a year (13% vs. 6%, p ⬍ 0.0001) (37).

TUMOR STAGING SYSTEMS AND PROGNOSTIC SCORING SYSTEMS


A number of tumor staging systems have been used to predict outcome with DTC.
Still, outcome cannot be accurately forecasted in individual patients. Two studies
(152,153) that critically compared the predictive accuracy of available prognos-
tic staging systems found that none accounted for more than a small portion of
the uncertainty in predicting outcome and that there was no statistically signif-
icant superiority of any system over that of the TNM (tumor, node metastasis)
classification of the American Joint Commission on Cancer (AJCC) and the Inter-
national Union Against Cancer (UICC) (Table 3) (154). The authors advised that,
because the TNM classification is universally available and widely accepted for
other disease sites, it should be used for all reports of the treatment and outcome
of patients with thyroid carcinoma, an opinion also expressed by the American
Thyroid Association (ATA) (93) and European Thyroid Associations (ETA) (155).
The greatest utility of staging systems is in epidemiological studies and
as tools to stratify patients for prospective therapy trials (72). Staging systems
may be less useful in determining treatment for individual patients unless a

Table 3 TNMa Classification of the AJCC and UICC Sixth


Edition Papillary or Follicular

Stage ⬍45 yr ≥45 yr Medullary

I M0 T1 T1
II M1 T2–3 T2–4
III T4 or N1 N1
IV M1 M1
a T is primary tumor: T1 , <2 cm; T2 , 2 cm to 4 cm; T3 , >4 cm; T4 ,
extension beyond thyroid capsule. N is regional lymph nodes: N1 , regional
lymph node metastases (cervical and upper mediastinal nodes). M is distant
metastases: M0 , no distant metastases; M1 , distant metastases present. All
undifferentiated (anaplastic) carcinomas are stage IV.
260 Sipos and Mazzaferri

reproducible group of patients can be identified with a very low risk of recur-
rence and cancer-specific mortality. Because the TNM classification of the AJCC
and UICC is universally available and widely accepted for other disease sites, it is
often recommended for thyroid carcinoma (152). Still, most patients are classified
as stage I with this classification (Table 3), which de facto categorizes most patients
as being at low risk (123). One study (123) found that patients with TNM stage-I
tumors had a 15% recurrence rate after an 11-year median follow-up, which would
argue against less aggressive therapy for this group. The numerous staging and
prognostic scoring systems that have been proposed underscore the fact that none
fully provides information to guide therapy.

TREATMENT OF PAPILLARY AND FOLLICULAR


THYROID CARCINOMAS
Surgery
Subtotal Lobectomy
Resection of less than a thyroid lobe, sometimes done as a nodulectomy, is inad-
equate therapy for thyroid carcinoma and is not the current standard of practice
(72,156). Even microscopic thyroid carcinoma requires more surgery than subtotal
lobectomy (157).
Ipsilateral Lobectomy and Isthmusectomy
A few surgeons prefer lobectomy, often referred to as subtotal thyroidectomy,
and regional lymph node dissection as the initial therapy for nearly all patients
with DTC (158). When the diagnosis of thyroid carcinoma is known preopera-
tively, however, most advise total or near-total thyroidectomy for all patients (159)
because it improves disease-free survival, even in children and adults with low-risk
tumors (160). Patients treated with lobectomy alone have a 5% to 10% recurrence
rate in the opposite thyroid lobe (5), an overall long-term recurrence rate ⬎30%
(12), and the highest frequency (11%) of subsequent pulmonary metastases (161),
compared with recurrence rates of only 1% after total thyroidectomy and 131 I ther-
apy (12). Higher recurrence rates are also observed with cervical node metastases.
Multicentric tumors—often found on study of the final histological sections—also
justify more complete initial thyroid resection (37). Lobectomy may be adequate
surgery for PTMC discovered serendipitously on the final pathology studies of
surgery done for benign disease provided that the patient has not been exposed to
radiation, has no other risk factors, and has a truly low-risk carcinoma—a tumor
⬍1.0 cm that is unifocal and confined to the thyroid without vascular invasion
(28,37,157,162). Complications with lobectomy are few, and survival in this latter
group is virtually assured (28,37,157,162). Nonetheless, the thyroid remnant tissue
hampers long-term follow-up with serum Tg determinations and whole-body 131 I
scans; therefore the decision to forgo complete thyroidectomy must be discussed
with the patient.
Differentiated Thyroid Carcinoma 261

Total or Near-total Thyroidectomy


Total or near-total thyroidectomy (ipsilateral total lobectomy, isthmusectomy, and
nearly total contralateral lobectomy) is the preferred surgical approach for the
majority of patients with a diagnosis of thyroid cancer. If any of the following
criteria are present, the patient should undergo total or near-total thyroidectomy:
tumors that are ≥l.0 cm in diameter, multicentric (any size), metastatic, penetrating
the thyroid capsule, histological variants with aggressive behavior, if the patient has
a first-degree relative with differentiated thyroid cancer, or patients with a history
of radiation therapy to the head and neck (163–165). Total thyroidectomy should
also be considered if there are contralateral nodules present, or if the patient is older
than 45 years. Complete or nearly complete thyroid resection removes multifocal
and bilateral carcinoma and provides the opportunity to ablate residual thyroid bed
uptake with small doses of 131 I, which substantially facilitates long-term follow-
up. Total thyroidectomy reduces the rates of recurrence and may also improve
long-term survival (166). If metastatic lymph nodes have been identified either
pre- or postoperatively, modified neck dissection (levels II to VI) that preserves
the sternocleidomastoid muscle is performed to remove involved cervical lymph
nodes. Radical neck dissection is done only for tumors that extensively invade the
strap muscles.

Completion Thyroidectomy
When subtotal thyroidectomy has been performed, it is best to consider completion
thyroidectomy for lesions that are anticipated to have the potential for recurrence
and because large thyroid remnants are difficult to ablate with 131 I (167,168). This
surgery has a low complication rate and is appropriate to perform routinely for
aggressive thyroid cancer variants, metastatic disease, PTC ⬎ 1 cm, FTC ⬎1 cm
with more than minimal capsular invasion, or multifocal carcinomas of any size,
because about half the patients have residual carcinoma in the contralateral thyroid
lobe (Table 4) (169). When there has been a local or distant tumor recurrence
following subtotal thyroidectomy, carcinoma is found in ⬎60% of the excised
contralateral lobes (170).

Table 4 Residual Carcinoma in Contralateral Thyroid Lobe Found with Completion


Thyroidectomy

Number of Patients (Percent with Residual Disease in Contralateral Lobe)

Total
Total number of patients residual
from seven studiesa Papillary Follicular Hürthle cell cancer

545 327 (58%) 206 (38%) 12 (57%) 244 (45%)


a Sources: Adapted from Refs. 43, 73, 76–78, 272, 273.
262 Sipos and Mazzaferri

Surgical Complications
The main complications of thyroidectomy are hypoparathyroidism and recurrent
laryngeal nerve damage, which are most common after total thyroidectomy. The
rates of hypoparathyroidism immediately after surgery are as high as 5% in adults
(128) and even higher in children (171,172) undergoing total thyroidectomy. How-
ever, the rates of persistent hypocalcemia are much lower. For example, one study
reported a 5.4% rate of hypocalcemia after total thyroidectomy that persisted in
only 0.5% of the patients one year after surgery (173). In a review of seven pub-
lished surgical series, the average rates of permanent recurrent laryngeal nerve
injury and hypoparathyroidism, respectively, were 3% and 2.6% after total thy-
roidectomy and 1.9% and 0.2% after subtotal thyroidectomy (174). When experi-
enced surgeons perform the surgery and the posterior thyroid capsule is left intact
on the contralateral side, hypoparathyroidism occurs at a lower rate. A study of
5860 patients treated in the state of Maryland found that surgeons who performed
more than 100 thyroidectomies a year had the lowest overall complication rates
(4.3%), which were fourfold lower than those of surgeons who performed ⬍10
cases annually (175).
Thyroidectomy during Pregnancy
Thyroid carcinoma may occasionally progress rapidly during pregnancy, per-
haps due to high maternal ␤-hCG levels, which have a TSH-like effect (133).
Nonetheless, most DTCs are slow growing and have an excellent prognosis during
pregnancy; therefore, surgery can usually be delayed until after delivery (176).

Radioiodine (131 I) Therapy


Sodium-Iodide Symporter
DTCs concentrate iodide much less avidly than normal thyroid tissue, perhaps
due to abnormalities in the sodium-iodide symporter. Increased sodium-iodide
symporter activity in PTC was reported in one study (177), but most find
reduced sodium-iodide symporter activity and heterogeneous immunohistochem-
ical sodium-iodide symporter staining in DTC (178,179).
Rationale for Thyroid Remnant Ablation
Because it is nearly impossible to remove all thyroid tissue with routine surgery,
131
I uptake is almost always seen postoperatively in the thyroid bed (including nor-
mal tissue of thyroglossal duct remnants). These foci must be ablated before 131 I
will optimally detect and be concentrated in metastatic deposits (180). Although
there continues to be debate concerning 131 I ablation of thyroid bed uptake after
near-total thyroidectomy, there are four compelling reasons to do this. First, a
thyroid remnant with high radioiodine uptake can obscure cervical or lung metas-
tases when their contribution to the total radioiodine uptake is low (180). Second,
high levels of circulating TSH, necessary to enhance tumor 131 I uptake, cannot
Differentiated Thyroid Carcinoma 263

be achieved in the presence of a large thyroid remnant (180). Patients with large
amounts of residual tissue in the neck (e.g., an entire lobe) prohibiting adequate
hypothyroidism should be strongly considered for completion thyroidectomy, and
only under unusual circumstances should 131 I be used to primarily ablate this
tissue. Third, serum Tg measurement (and its trend) is the most sensitive test
for detection of carcinoma when measured during hypothyroidism after thyroid
bed uptake ablation (181). Fourth, ablative doses may destroy undocumented
micrometastases, microscopic thyroid bed foci of malignancy, or eliminate “nor-
mal” residual tissue otherwise destined to become malignant. These arguments
would be consistent with decreased rates of thyroid bed recurrence and metastases
in patients treated with 131 I remnant ablation in the absence of residual malignancy.
Postoperative 131 I remnant ablation is done when the patient has a tumor
with the potential for recurrence (121). There are a large and growing number
of studies that demonstrate decreased recurrence and disease-specific mortality
rates from DTC attributable to 131 I therapy (Fig. 8) (37,121,132,182–186). The
lowest incidence of pulmonary metastases occurs after total thyroidectomy and
131
I. For example, in one study recurrences in the form of pulmonary metastases,
analyzed as a function of initial therapy for DTC, were reported to be as fol-
lows: thyroidectomy plus 131 I (ablation dose of 100 mCi), 1.3%; thyroidectomy
alone, 3%; partial thyroidectomy plus 131 I, 5%; and partial thyroidectomy alone,
11% (121).
Preparation for 131 I Therapy
Females with childbearing potential must have a negative pregnancy test doc-
umented shortly before receiving diagnostic imaging or therapeutic amounts of
131
I. For remnant ablation, 131 I therapy is ideally given approximately six weeks
after surgery. During the four to six weeks before 131 I therapy, iodine-containing
drugs must be carefully avoided. This is especially important for sources of
long-lasting iodine such as intravenous CT contrast, which routinely impairs the
response to 131 I for two to three months or longer, depending upon the num-
ber of scans that have been done. A low iodine diet should be ingested for
two weeks prior to therapy (187). The serum TSH levels must be high enough
(⬎30 mIU/mL) to stimulate sodium-iodide symporters in neoplastic and nor-
mal thyroid tissues to concentrate 131 I. For decades, the only way to do this
was to withdraw levothyroxine therapy and to administer liothyronine (Cytomel)
for four weeks and stop it for two more weeks, and then administer 131 I. This
causes profound hypothyroidism with TSH levels often well ⬎100 mIU/L (188).
However, a prospective randomized study found that administering liothyro-
nine for four weeks is unnecessary after thyroid hormone withdrawal (THW),
causing neither fewer symptoms of profound hypothyroidism nor improving the
rapidity of the rise in serum TSH levels (189). After THW, serum TSH lev-
els must be measured before diagnostic or therapeutic 131 I dosing because the
TSH response to THW is unpredictable or often fails to rise ⬎30 mIU/L. Inabil-
ity to adequately stimulate TSH elevation should raise suspicion of insufficient
264 Sipos and Mazzaferri

Figure 7 Preparation for 131 I imaging and/or therapy.

thyroidectomy, functioning metastases, continued thyroid hormone ingestion, or,


rarely, hypopituitarism, but is more often related to advanced age more than
60 years (190). On the other hand, elevating the TSH level ⬎30 mIU/L does not
improve the therapeutic response to 131 I (191), and the optimal magnitude of TSH
elevation is unknown and differs among patients.
The other way to increase serum TSH levels is by intramuscular adminis-
tration of recombinant human TSH (rhTSH), which was approved by the FDA in
2007 for remnant ablation, while the patient continues thyroid hormone therapy
and thus avoiding symptomatic hypothyroidism (192). The drug is administered
intramuscularly at a dosage of 0.9 mg for two consecutive days and 131 I is adminis-
tered 24 hours after the last injection (Fig. 7) (193). Mean peak TSH concentrations
are reached between 3 and 24 hours after injection (median of 10 hours) and the
mean half-life is 25 ± 10 hours (192,194). Serum TSH levels are almost always
⬎50 mU/L (192,195,196). However, measurement of serum TSH is not gener-
ally advised after rhTSH injection because by the time the patient undergoes 131 I
treatment, the serum TSH levels are below the usual limits recommended for
treatment. Still, when given to euthyroid patients taking levothyroxine, rhTSH
injection is as effective as thyroid hormone withdrawal in preparing patients for
remnant ablation and produces an equally favorable therapeutic response (193).
Moreover, the use of rhTSH reduces total body radiation from 131 I by 33% com-
pared with withdrawal-induced hypothyroidism, which delays the renal excretion
of 131 I thus increasing whole-body irradiation (193). Also, remnant ablation with
Differentiated Thyroid Carcinoma 265

Figure 8 Recurrence rates of DTC after various forms of medical therapy. The differences
are statistically significant between all for treatments shown. Source: Adapted from Ref. 4.

50 mCi of 131 I after rhTSH preparation is as effective as 100 mCi 131 I (197),
lowering whole-body irradiation to an even greater extent. Studies show that
short-term hypothyroidism after thyroid hormone withdrawal is associated with a
significant decline in quality of life that is abrogated by the use of rhTSH (198).

Diagnostic Whole-Body 131 I Scan and the Stunning Effect


Prior to remnant ablation, some physicians perform a diagnostic (“pretreatment”)
whole-body scan in order to establish the size and radioiodine avidity of the
thyroid remnant, and to search for the presence of cervical nodal disease. If a
diagnostic scan is ordered, it is usually obtained 24 to 72 hours after giving 2 to
4 mCi of 131 I (Fig. 7). Larger amounts of 131 I should not be given because focal
abnormalities not seen with 2 mCi are less likely to be ablated successfully (199),
and 131 I doses as small as 3 mCi diminish the subsequent uptake of therapeutic
131
I, which is termed the “stunning effect” (200,201). To avoid stunning, doses of
1 to 3 mCi of 131 I have been recommended; however, these doses are slightly less
sensitive than larger scanning doses of 131 I in identifying thyroid remnants, and
they require longer imaging times (202). Administration of the therapeutic dose
as soon as possible after the diagnostic dose of 131 I helps to minimize stunning
(202,203). Although 123 I in doses of ≥1.5 mCi has been reported to yield excellent
images without stunning, its use to date has been somewhat limited by issues of
266 Sipos and Mazzaferri

cost and availability. Furthermore, the logistics of doing pretreatment scanning


using rhTSH are complex, so that pretreatment scanning is usually not done in
this setting.
Amount of 131 I for Thyroid Remnant Ablation
Usually remnant ablation can be achieved with 30 to 50 mCi of 131 I, which is
as effective as larger doses in ablating the thyroid remnant and preventing tumor
recurrence (37,197,204). This has been a popular way to avoid hospitalization but
is no longer necessary in most states because of changes in federal regulations
permitting outpatient use of much larger 131 I doses in most states (205). Even
so, considering the differences in cost and radiation exposure and the fact that
doses to the thyroid remnant more than 30,000 rad do not substantially improve
the rate of successful ablation, it may be reasonable to use a relatively low dose
for remnant ablation (206). Successful thyroid remnant ablation with doses of
approximately 30 and 51 mCi of 131 I were reported as 63% and 78%, respectively,
in one study from New Delhi (206). However, this optimal dose of approximately
51 mCi was reported to deliver approximately 30,000 rad to the thyroid remnant.
In comparison, in one American study, 30,000 rad was achieved with a mean
I31
I dose of approximately 87 mCi and completely ablated the remnant in 86%
of cases (167). Another prospective study comparing 30 and 50 mCi doses of
131
I found no significant difference in ablation rates (207). Increasing the dose
to deliver more than 30,000 rad does not increase the success rate (167,206).
A recent meta-analysis of all the prospective trials showed no difference among
various doses of radioiodine (208). A subsequent prospective randomized study
of rhTSH to prepare patients for remnant ablation with 50 or 100 mCi also found
that the two were equally successful in achieving remnant ablation (197). The rate
of successful ablation is significantly lower when patients have less than a total or
near-total thyroidectomy or have a thyroid remnant calculated to be ⬎2g (167).
Posttreatment Scans and False-Positive Scans
Following therapy with 131 I, a posttreatment scan is performed with 5 to 10 days
later in order to visualize areas in the neck (and elsewhere) that contain iodine-
avid tissue. In most patients, an area in the midline corresponding to the thyroid
remnant is all that is seen. However, cervical nodal disease may be visualized in
papillary thyroid cancer, and more rarely diffuse lung uptake is seen, especially in
children and adolescents. Many things can infrequently cause false-positive 131 I
scans, including body secretions, transudates, inflammation, nonspecific mediasti-
nal uptake (e.g., blood pool), and neoplasms of nonthyroidal origin, which may
uncommonly concentrate 131 I (209). False-positive scans can also be seen with
physiological secretion of 131 I from the nasopharynx, salivary and sweat glands,
stomach, and genitourinary tract and from skin and hair contamination with spu-
tum or tears (210). Diffuse hepatic uptake of 131 I is rarely due to occult liver
metastases but more commonly due to hepatic clearance of Tg labeled with 131 I
by functioning thyroid remnants or extrahepatic thyroid cancer metastases. The
Differentiated Thyroid Carcinoma 267

more 131 I uptake by residual thyroid tissue, the more 131 I appears in the liver. In
one large study (211), 12% of all diagnostic scans showed uptake in the liver.
The frequency of hepatic uptake in posttherapy scans was related to the dose of
131
I, being 39% with 30 mCi, 61.5% with 75 to 100 mCi, and 71.3% with 150 to
200 mCi. On the other hand, hepatic uptake of 131 I without 131 I uptake elsewhere
suggests hidden metastases (211).

Treatment of Residual or Recurrent Carcinoma with 131 I


Thyroid carcinoma (especially macroscopic disease) should be treated surgically
whenever possible. Only 50% to 75% of DTCs and their metastases and about
one-third of Hürthle cell carcinomas concentrate 131 I (212–214). Moreover, the
larger the tumor mass, the less likely that 131 I therapy will successfully ablate
the tumor. One study found that two-thirds of 283 patients with lung or bone
metastases had tumors that concentrated 131 I, which is crucial to survival (215).
Another study found that 10-year survival rates were 83% or 0%, respectively,
depending on whether pulmonary metastases did or did not concentrate 131 I (139).
Lung metastases that concentrate 131 I most avidly are the smallest lesions found
in young patients (139).
There are three approaches to radioiodine therapy: empiric fixed doses,
upper-bound limits set by blood dosimetry, and quantitative dosimetry (205).
Dosimetric methods are often reserved for distant metastases or unusual cases,
as when renal failure is present or therapy with rhTSH stimulation is deemed
necessary.

Empirical Fixed Doses


With empirical doses, a fixed amount of 131 I is given based on tumor stage.
Generally, approximately 30 to 125 mCi are given to ablate thyroid remnants, 150
to 200 mCi for residual carcinoma in cervical nodes or neck tissues, and 200 mCi
for distant metastases. Tumor 131 I uptake in amounts adequate for imaging with
4-mCi diagnostic doses is usually sufficient for 131 I therapy, using empirical doses
from 30 to 200 mCi (216). However, two retrospective studies (217,218) found
that administering ⬍140 mCi of 131 I may rarely expose older patients to blood
doses of ⬎200 cGy, the upper allowable exposure limit. Administering 200 to
250 mCi of 131 I frequently exceeded the blood exposure limit in patients 70 years
of age or older. Consequently, dosimetry-guided 131 I therapy may be preferable
to fixed-dose 131 I treatment in older patients with thyroid cancer and in patients
with 131 I-avid diffuse bilateral pulmonary metastases, even when renal function is
normal (217,218).

Upper-Bound Limits Set by Blood Dosimetry


This approach establishes an upper limit on the amount of 131 I in a single dose that
can be given safely, which is generally considered to be 200 rad to the whole blood
(205). In patients with diffuse pulmonary metastases, the dose is also limited, so
that ⬍80 mCi of 131 I remains in the lungs after 48 hours to avoid pulmonary fibrosis.
268 Sipos and Mazzaferri

Without diffuse pulmonary metastases, most authors suggest that the whole-body
retention be ⬍120 mCi at 48 hours; however, this limit is usually greater than that
of the blood dosimetry and thus not a limiting factor. Based on these calculations,
therapy doses of 450 to 600 ± mCi are sometimes used. However, two studies
have shown that larger amounts of 131 I may be given to patients pretreated with
rhTSH injections compared with levothyroxine withdrawal with a lower risk of
bone marrow suppression (219,220).

Quantitative Tumor Dosimetry


This approach calculates the amount of 131 I that is required to deliver 30,000 rad
to ablate the thyroid remnant or 8000 to 12,000 rad to treat nodal or discrete soft
tissue metastases. For pulmonary metastases, an amount of 131 I is administered
that will deliver 200 rad to whole blood with no ⬎80 mCi of whole blood retention
at 48 hours (205). The mass of residual tissue and the effective half time of 131 I
in that tissue are the two most important factors in determining success (205). In
one study, an 80% response was found in tumor deposits that received at least
8000 rad (167). Lesions that receive ⬍3000 to 4000 rad from 150 to 200 mCi 131 I
should be considered for alternative therapy.

Repeat 131 I Treatments


Treatment with 131 I should be continued every 6 to 12 months, as long as metastatic
deposits are present and continue to concentrate 131 I and a favorable response
to previous therapy has been demonstrated, such as a decreasing Tg level or
decreasing tumor mass. Repeat 131 I doses should not be given until the bone
marrow has fully recovered from the previous dose. Few adverse effects occur
with this approach to 131 I therapy (219). Before large cumulative 131 I activities
(⬎500–600 mCi) are given over an extended period to treat neck metastases
or well-localized disease, especially in the CNS or spine, serious consideration
should be given to excising the tumor surgically.

Lithium
This drug inhibits iodine release from the thyroid without impairing iodine uptake,
thus enhancing 131 I retention in normal thyroid and tumor cells (221). One study
(222) showed that the mean increase in the biological or retention half-life was 50%
in tumors and 90% in thyroid remnants. The effect was greater in lesions with poor
131
I retention (222). Nevertheless, there are no outcome data showing improved
survival or efficacy of treatment in patients receiving lithium as an adjunct to
radioiodine therapy. In patients withdrawn from thyroid hormone, serum lithium
levels should be measured frequently if not daily and maintained between 0.8 and
1.2 nmol/L.

Immediate Complications of 131 I


There are few immediate serious risks of 131 I therapy except when metastases are
in critical locations that will not tolerate posttherapeutic swelling. For example,
Differentiated Thyroid Carcinoma 269

brain or spinal cord metastases can undergo potentially catastrophic edema and
hemorrhage 12 hours to 2 weeks after 131 I treatment (223). In patients with disease
in “critical areas” such as the CNS, pretreatment with high doses of glucocorticoids
has been recommended (224). Another example is severe radiation thyroiditis,
which can occur within a week of administering a large dose of 131 I to a patient
who has undergone only lobectomy, causing pain, swelling, and rarely airway
compromise that may require prednisone therapy (225). Thyroid storm may rarely
occur approximately 2 to 10 days after administering a therapeutic amount of 131 I,
especially when there is a large burden of functioning tumor (205). Acute bone pain
is sometimes experienced after 131 I treatment. Radiation sickness characterized
by headache, nausea, and occasional vomiting is experienced by about two-thirds
of patients approximately 4 to 12 hours after the oral administration of 200 mCi
or more of 131 I, which resolves in approximately 24 hours; this almost never
occurs with smaller amounts of 131 I (205). Patients with extensive neck tumor
may rarely develop transient vocal cord paralysis, and facial nerve paralysis has
been reported after very high doses of 131 I (205). Radiation cystitis does not occur
if the patient is well hydrated. Mild radiation sialadenitis, leukopenia, and a slight
drop in the number of platelets often occur approximately six weeks after therapy,
but ordinarily these effects are mild and transient unless very large doses of 131 I
are administered (226).

Parotid Dysfunction
Transient parotid swelling reminiscent of Stensen’s duct obstruction may occur
for nearly a year after 131 I therapy. Having the patient suck on hard lemon candy
starting 24 hours after therapy increases salivary flow, which may decrease but
does not prevent the adverse effect of 131 I radiation on the salivary gland. In
one study approximately 60% of patients reported side effects lasting longer than
three months, which included sialoadenitis (33%) and transient loss of taste or
smell (27%) (227). More than a year after the last 131 I treatment, 43% suffered
from reduced salivary gland function and ⬎4% had complete xerostomia, and
approximately 23% of the patients reported chronic or recurrent conjunctivitis,
complications that were related to the cumulative dose of 131 I (227).

Radiation Pneumonitis
Pulmonary fibrosis is a potential complication of 131 I therapy for diffuse pul-
monary metastases when the lung retention of 131 I is ⬎80 mCi 48 hours after treat-
ment. Some reports suggest that diffuse pulmonary metastases can be treated with
150 mCi of 131 I without risking pulmonary fibrosis (228) and smaller amounts of
131
I doses in the range of approximately 100 mCi are often given when there is dif-
fuse and intense uptake of the scanning dose in the lungs. However, a randomized
controlled study to investigate outcome of distant metastases treated with these
relatively low doses versus dosimetrically determined much higher doses (often
200–800 mCi) has not been conducted, so optimal therapy is not known.
270 Sipos and Mazzaferri

Leukemia, Second Tumors, and Other Bone Marrow Effects of 131 I


There is a small risk of developing acute myelogenous leukemia after 131 I therapy,
which in several studies has been estimated to range from 3 to 22 excess cases
per 1000 patients treated with 131 I, depending upon the cumulative amount of 131 I
administered (205,229). The late health effects associated with 131 I therapy for
thyroid cancer have been difficult to fully assess since the number of patients who
develop leukemia after being treated with 131 I is limited. For example, a Swedish
study found only two cases of leukemia among 834 thyroid carcinoma patients
treated with 131 I, which was not a statistically significant increase rate over that
found in the general population (230).
However, much larger studies have found a significantly greater risk of
second primary malignancies (SPM) in patients treated with 131 I. A European study
of 6841 patients with thyroid cancer (mean age 44 years) that was diagnosed from
1934 to 1995 found that 17% had been treated with external beam radiotherapy
and 62% had received 131 I therapy. An SPM was found in 576 (8%) of the thyroid
cancer patients, representing a 27% increase over that found in the population (95%
CI: 15–40). There was a significantly increased risk of cancer of the digestive tract,
bone and soft tissue, skin melanoma, kidney, CNS, and endocrine glands other
than thyroid, and leukemia. The risk of solid tumors and leukemias increased
as the cumulative amounts of administered 131 I increased, resulting in an excess
absolute risk of 14.4 solid cancers and of 0.8 leukemias per GBq (27 mCi) of 131 I
per 100,000 person-years of follow-up. A group of Utah investigators determined
the risk of a nonthyroidal SPM in 30,278 American patients with thyroid cancer
diagnosed between 1973 and 2002 in centers participating in the National Cancer
Institute’s Surveillance, Epidemiology, and End Results (SEER) program (231).
Median follow-up was 103 months (range, 2–359). A total of 2158 patients (7%)
developed nonthyroidal SPMs, which was significantly more than expected in the
general population [observed/expected (O/E) = 1.09; 95% CI: 1.05–1.14; p ⬍
0.05]. The absolute excess risk per 10,000 person-years was 6.39. Compared with
the general population, the risk of SPM was significantly greater than expected
(O/E = 1.20; 95% CI, 1.07–1.33; AER = 11.8) as was the increased risk observed
in nonirradiated patients (O/E = 1.05; 95% CI, 1.00–1.10; AER = 3.53). Still,
the risk of SPM was greater for irradiated patients than the nonirradiated cohort
(relative risk = 1.16; 95% CI, 1.05–1.27; p ⬍ 0.05). The greatest risk of second
primary cancers occurred within five years of the diagnosis of thyroid cancer and
was also elevated for younger patients.
When 131 I treatments are given at 12-month intervals and total cumulative
doses are limited to 500 mCi in children and 600 mCi in adults, long-term effects
on the bone marrow are minimal and few cases of leukemia occur (229,232,233).
Limiting the cumulative dose is usually not a problem unless large single doses of
131
I are given, because tumor tissue that concentrates 131 I is likely to be ablated by
a few 131 I treatments, leaving either no residual tumor or metastases that do not
concentrate 131 I. For example, all patients cured of pulmonary metastases by 131 I
Differentiated Thyroid Carcinoma 271

in one study (234) did so with a cumulative dose of ≤1500 mCi. It is reasonable
to give high cumulative doses of 131 I to patients with extensive metastatic disease
responsive to therapy, as the risk posed by the known thyroid cancer outweighs
the risk of a potential second cancer from radiation. These observations of SPM,
including colon cancer following 131 I underscores the need for laxatives and
hydration after 131 I treatment, especially for hypothyroid patients.
Lacrimal Duct Obstruction Induced by 131 I
In 2002, Kloos et al. (235) described a patient who developed complete bilateral
nasolacrimal duct obstruction (epiphora) after 131 I therapy for thyroid cancer,
which first prompted awareness of this potential complication. After studying 390
patients who had received 131 I for thyroid remnant ablation or tumor therapy, 10
were found to have epiphora. All had evidence of nasolacrimal duct obstruction
that occurred after being treated with an individual dose of 180 ± 15 mCi of 131 I
(mean ± SE) and a cumulative dose of 467 ± 79 mCi of 131 I. Symptoms appeared
6.5 ± 1.4 months (range, 3–16) after the last 131 I treatment; however, the time from
symptom onset to correct diagnosis was 18 ± 5 months. This complication did
not develop in patients who did not receive 131 I therapy or were treated with ⬍150
mCi. In all, 3% of the cohort had evidence of the problem, which was manifest
by epiphora, discharge on the eyelids, recurring conjunctivitis, dacryocystitis and
a mass below the median canthal tendon.
Patients reporting epiphora should be promptly evaluated by an oculoplastic
surgeon. Management of patients with complete obstructions requires more exten-
sive surgical procedures than does management of patients with an incomplete
obstruction. Early intervention with balloon dilation of the nasolacrimal duct
and/or stents may prevent complete obstruction until radiation-induced inflamma-
tion subsides. Although the natural history of partial obstruction is unknown, data
from several patients suggests that spontaneous improvement may occur without
intervention.
Infertility and Gonadal Failure
Gonadal damage may be caused by large doses of 131 I, but it is infrequently
observed (205,236). A large European study of 2113 pregnancies in women who
had been treated with surgery and 131 I for thyroid cancer found that the miscarriage
rate increased from 11% before surgery to 20% after surgery, remaining at this
level after 131 I therapy (237). In this study, miscarriages were more frequent when
women were treated during the year preceding conception; however, whether this
is related to gonadal irradiation or to insufficient control of hormonal thyroid
status is uncertain (237). The incidences of stillbirth, preterm birth, low birth
weight, congenital malformation, and death during the first year of life were not
significantly different before or after 131 I therapy; the incidence of thyroid disease
and nonthyroidal malignancy was similar in children born either before or after
their mothers were exposed to 131 I. Testicular germinal cell function may be
transiently impaired when men are given 131 I therapy, although the damage may
272 Sipos and Mazzaferri

become permanent when large doses of 131 I are delivered year after year (238).
Since this might pose a significant risk of infertility, it seems prudent to advise
young men to bank their sperm before repeated high dose 131 I therapy.

Thyroid Hormone Therapy


Levothyroxine (T4) Suppression of TSH
DTCs contain TSH receptors that stimulate the cell growth and iodine uptake of
well-differentiated follicular cancer cells (111). Thyroid hormone therapy signifi-
cantly reduces recurrence rates and cancer-specific mortality rates (Fig. 8) (121).
A meta-analysis of thyroid hormone suppression therapy in thyroid cancer patients
showed an association with reduced risk of major adverse clinical events, defined
as disease progression and/or recurrence and death (239). The National Thyroid
Cancer Treatment Cooperative Study (NCTCS) that has prospectively performed
follow-up of 2936 patients with DTC found that thyroid hormone therapy signifi-
cantly reduced recurrence rates and cancer-specific mortality rates in patients with
stage II to IV disease (Fig. 8) (240). However, the study was unable to show any
impact, positive or negative, of any form of surgical or medical therapy in stage I
patients, which may be due to the relatively short duration of follow-up (median
three years, range 0–14) in this group of patients.
The levothyroxine dosage needed to attain serum TSH levels in the euthy-
roid range is greater among patients with thyroid cancer (2.11 ␮g/kg/day) than
among those with primary hypothyroidism caused by nonmalignant disease
(1.62 ␮g/kg/day) (241). One study found that patients who had undergone total
thyroid ablation for thyroid carcinoma required 2.7 ± 0.4 (SD) ␮g/kg/d of levothy-
roxine to achieve an undetectable basal serum TSH level that does not increase
after TRH administration (242). A French study found that a constantly sup-
pressed TSH (⬍0.05 ␮U/mL) was associated with a longer relapse-free survival
than when serum TSH levels were always 1 ␮U/mL or greater, and that the degree
of TSH suppression was an independent predictor of recurrence (243). A more
recent study showed a beneficial effect of thyroid hormone therapy, but only in
those patients whose serum TSH level was consistently ⬎2 mU/L (244). Hence,
these data do not support the concept that a great degree of TSH suppression
(into the undetectable, thyrotoxic range) is required in patients with stage I or II
disease but may be beneficial in patients with more advanced disease stage. The
most appropriate dose of thyroid hormone for patients with DTC who are free
of disease is that which reduces the serum concentration to just below the lower
limit of the normal range for the assay being used (93). The ATA management
guidelines for patients with thyroid cancer recommend that the TSH in patients
with persistent disease should be maintained ⬍0.1 mU/L indefinitely as long as
there are no contraindications. For patients at high risk of recurrence, the TSH
should be maintained at 0.1 to 0.5 mIU/L for 5 to 10 years. Patients with low-risk
tumors who are free of disease are advised to maintain the TSH within the low
normal range (0.3–2 mU/L).
Differentiated Thyroid Carcinoma 273

Complications of Levothyroxine Therapy


Potential problems associated with subclinical thyrotoxicosis are an increased
risk of atrial fibrillation (245) in older patients (older than 60 or 65 years), a
higher 24-hour heart rate, more atrial premature contractions per day, ventricular
hypertrophy, diastolic dysfunction, and impaired cardiac reserve (246). Patients
with thyroid carcinoma treated with suppressive doses of levothyroxine have a high
rate of bone turnover that decreases acutely after withdrawing treatment (247),
which is of most concern in postmenopausal women not receiving estrogen-like
or bisphosphonate therapy (248). Studies of fracture risk in women treated with
thyroid hormone suggest that there may be an increased risk when suppressive
doses are used (249). However, TSH suppression has no significant effects on
bone mass in men, according to one study (250).

Other Therapy
External Beam Radiation Therapy
This therapy is generally considered third-line therapy for localized DTC after
surgical resection and 131 I therapy. Typical candidate sites of disease for external
beam radiation therapy include the neck, upper mediastinum, or bone lesions
that are symptomatic or in critical locations to prevent fracture. External beam
radiation therapy is frequently reserved for inoperable, non-iodine-avid disease;
however, in some iodine-avid situations, it is considered for adjuvant therapy
following 131 I therapy such as in patients with T4 N1 disease and over the age
of 45 years, aerodigestive invasion, nonresectable local bulk disease, or selected
osseous metastases (132).
Gamma KnifeTM
Brain metastases from thyroid carcinoma are an extremely poor prognostic sign.
Surgical resection of brain metastases may be associated with a limited prolonga-
tion of survival and rarely with apparent cure (251). While controlled studies are
lacking, inoperable CNS metastases should probably be treated with gamma knife
rather than external beam radiation therapy if possible (251).
Chemotherapy
Patients with papillary thyroid cancer who no longer respond to the usual modes
of therapy—surgery, radioiodine, and external beam radiotherapy—and still show
signs of progressive disease may be candidates for investigational drugs. Histori-
cally, traditional chemotherapy (doxorubicin) has proven to be of limited benefit in
thyroid cancer and at best provides palliation. However, recent discoveries of the
molecular pathogenesis of papillary thyroid cancer have led to the use of promis-
ing molecular-targeted therapies. Numerous clinical trials are currently underway
to investigate the effects of various targeted therapies. These new agents can
be divided into several categories: oncogene inhibitors, angiogenesis inhibitors,
274 Sipos and Mazzaferri

redifferentiation agents, and gene therapy (252). Two recently published stud-
ies suggest some modest benefit from multi-kinase inhibitors such as sorafenib
(253) and motesanib (254). The RET/PTC oncogene is targeted by the tyrosine
kinase (TK) inhibitors; various steps in the molecular pathway may be inhibited
by this family of agents. Potential targets include monoclonal antibodies directed
against the tyrosine kinase ligand or its receptor, inhibition of ATP binding to
the TK receptor, inhibition of receptor phosphorylation or activation, or blocking
downstream signals (252). Several compounds are also being studied to determine
their ability to induce redifferentiation of tumors, particularly the ability to restore
function of the sodium-iodine symporter and thereby increase uptake of radioac-
tive iodine by the tumor cells. It may be necessary to use combination therapy
with these agents to achieve maximal tumor response in light of the complex and
overlapping relationships between the various mitogen activated protein kinase
pathways (252).
The patient should have evidence of disease progression before initiating
such therapies as many of these agents offer disease stabilization and a treatment
effect cannot be demonstrated unless the tumor is actively growing. In this fast
moving area, the best way to get current information on clinical trials for thyroid
cancer is to use the National Institutes of Health website or that of the ATA
(http://www.thyroid.org) or the National Cancer Institute (http://www.cancer.gov/
clinicaltrials).

Follow-up
It is convenient to stratify follow-up into three stages (Fig. 9): Phase 1 is four to
six weeks after surgery when the completeness of tumor resection is evaluated,
Phase 2 occurs approximately 6 to 12 months after remnant ablation to assess
the status of the initial surgical and medical therapy together, and Phase 3 is the

Figure 9 The three phases of follow-up in differentiated thyroid cancer.


Differentiated Thyroid Carcinoma 275

long-term follow-up that occurs after patients are deemed free of disease. At each
stage of follow-up different studies may be necessary, depending upon the patient’s
response to therapy.

Changing Follow-up Paradigms


The follow-up of patients with PTC has changed considerably over the past decade.
In the past, follow-up depended heavily upon the use of diagnostic whole-body
131
I scanning (DxWBS) and the assessment of serum Tg levels performed during
levothyroxine suppression of TSH (Tg-on). This approach, which was less accurate
than current follow-up strategies, resulted in long delays in identifying patients
with persistent metastases (37), which has the potential to reduce survival rates
(37,255). Risk stratification is now established immediately after completion of the
initial therapy, when follow-up strategy is adapted to fit the patient’s clinical status.

Identifying Patients Who Are Free of Disease


Identifying patients who are free of disease following initial therapy is a key
follow-up issue because this identifies 80% of those who have undergone adequate
therapy, which usually consists of total thyroidectomy and 131 I remnant ablation.
The ATA (93) and ETA (155) guidelines define disease-free status as follows: (1)
no clinical evidence of residual tumor; (2) No imaging evidence of tumor, which
generally means that the posttreatment whole-body 131 I scan (RxWBS) shows 131 I
uptake only in the thyroid bed and that neck ultrasonography is negative; and (3)
an undetectable Tg (⬍1 ng/mL) during both TSH suppression and stimulation, in
the absence of interfering serum antibodies.

Risk Stratification Based upon the Patient’s Clinical Status After


Initial Therapy
The AJCC TNM patient risk stratification and most tumor staging classifications
depend heavily upon patient age and tumor stage at the time of diagnosis. While this
provides a uniform means of comparing patient outcome in studies, it fails to take
into account other important tumor features that predict outcome. For example, the
TNM system does not account for tumor variables (e.g., tall cell papillary thyroid
cancer), tumor molecular features such as BRAF mutation, or patient variables
such as familial non-medullary thyroid cancer, all of which have an impact on
outcome. The ATA (93) and ETA (155) thyroid cancer guidelines stratify risk
on the basis of patient age and tumor stage at the time of diagnosis, and patient
response to initial therapy, which is usually defined by TSH-stimulated serum Tg
levels and neck ultrasonography. This is considerably more accurate than DxWBS
and Tg-on that was used in the past to define patient status after initial therapy.
Patients with low-risk tumors undergo a follow-up strategy that is substantially
different from that for high-risk patients (Fig. 10). This is done by stratifying risk
after initial surgery with or without postoperative radioiodine ablation into three
groups as follows:
276 Sipos and Mazzaferri

Figure 10 Follow-up algorithm for papillary and follicular thyroid carcinoma.

Low-risk patients are those whose tumors have no local or distant metastases;
all macroscopic tumors have been resected; no invasion of locoregional
tissues or structures by the tumor; no aggressive histology of the tumor
(e.g., tall cell, insular, or columnar cell tumors); no vascular invasion; and
if radioiodine ablation has been given, there is no evidence of 131 I uptake
outside the thyroid bed on posttreatment whole-body scanning (93).
Intermediate-risk patients have tumors which show microscopic invasion into
perithyroidal soft tissues, have tumors with aggressive histology, or tumors
which have vascular invasion (93).
High-risk patients have macroscopic tumor invasion, incomplete tumor resection,
distant metastases, or 131 I uptake outside the thyroid bed on posttreatment
whole-body scanning (93).

rhTSH in Follow-up
During follow-up, serum TSH concentration must be periodically increased to
levels sufficiently elevated to stimulate thyroid tissue sodium-iodide symporters
so that serum Tg measurement and radioiodine scanning can be performed. For
follow-up rhTSH is given as an intramuscular dose of 0.9 mg for two consecutive
days followed by 4 mCi of 131 I orally on the third day and a whole-body scan and Tg
Differentiated Thyroid Carcinoma 277

measurement on the fifth day (Fig. 7). Whole-body 131 I images are acquired after
30 minutes of scanning or after obtaining 140,000 counts, because a 4-mCi dose
of 13I I may have the same body retention as a 2-mCi dose given to a hypothyroid
patient. When a large- or small-field-of-view camera is used, a minimum of 60,000
and 35,000 counts per view, respectively, are required.
Serum Thyroglobulin and Cervical Ultrasound
Serum Tg determinations and cervical ultrasonography can almost always detect
residual thyroid tissue, whether benign or malignant, in patients who have under-
gone thyroidectomy. This has been shown in a number of clinical studies that
have provided new information regarding optimal surveillance protocols for low-
and high-risk patients with differentiated thyroid cancer. A 2003 review (256) of
eight follow-up studies comprising 1028 patients found a growing consensus on
the clinical value of TSH-stimulated Tg measurements as part of routine surveil-
lance. One of the early findings was that an undetectable serum Tg measured
during thyroid hormone suppression of TSH (THST) was often misleading (256).
Analysis of eight follow-up studies found that 21% of 784 patients who had no
clinical evidence of tumor with baseline serum Tg levels ⬍1 ng/dL had a rise in
rhTSH-stimulated serum Tg that was ⬎2 ng/mL. When this occurred, 36% of the
patients were found to have metastases, about one-third of which were at distant
sites, which in almost all the cases (91%) was associated with an rhTSH-stimulated
Tg ⬎2 ng/mL (256). On the other hand, a DxWBS performed after either rhTSH
stimulation or THW, identified only 19% of the cases of metastases (over an 80%
false-negative rate). Ten studies comprising 1599 patients demonstrate that a TSH-
stimulated Tg using a Tg cutoff of 2 ␮g/L (either after THW or 72 hours after
rhTSH) is sufficiently sensitive to be used as the principal test in the follow-up
management of low-risk patients with DTC and that the routine use of DxWBS
in follow-up should be discouraged. On this basis, a surveillance paradigm was
proposed using TSH-stimulated Tg for patients who have had a total or near-total
thyroidectomy and 131 I ablation and have no clinical evidence of residual tumor
with a baseline serum Tg ⬍1 ng/mL during THST (256).
However, careful analysis found that although the negative predictive value
(NPV) of TSH-stimulated serum Tg measurements was approximately 100%,
the positive predictive value (PPV) was only approximately 50%, which was
the case for both THW (257) and rhTSH (258) stimulation; moreover, when the
patient’s TSH-Tg was studied over time, the PPV increased to approximately 85%
(257,259). A serum rhTSH-stimulated Tg followed over a span of three to five
years showed that the PPV of a serum rhTSH-stimulated Tg was 80% while half
the patients with rhTSH-stimulated Tg values ⬍2 ng/mL experienced a gradual
decrease in Tg values to undetectable levels. A meta-analysis (260) found that the
highest accuracy of Tg-guided follow-up is obtained if treatment includes thyroid
remnant ablation, and Tg testing is performed while the patient is off thyroxine
(sensitivity 96% and specificity 94%). The sensitivity of rhTSH-stimulated Tg
was 93% but specificity was only 76%. For this reason, following serum Tg trends
278 Sipos and Mazzaferri

after initial therapy has become a widely accepted means of anticipating the result
of therapy.

Tg and Whole-Body 131 I Scans


Serum Tg determinations and whole-body 131 I imaging can almost always detect
residual thyroid tissue, whether benign or malignant, in patients who have under-
gone thyroidectomy. The serum Tg concentration correlates with the mass of
normal or malignant thyroid tissue, the amount of thyroid physical damage or
inflammation, and the level of TSH receptor stimulation. Tg measurement is more
sensitive when thyroid hormone has been stopped or rhTSH is given to elevate
the serum TSH and, under these conditions, has a lower false-negative rate than
whole-body 131 I scanning (256). The highest NPV (99.5%) and sensitivity (96%)
was achieved with rhTSH-stimulated serum Tg and neck ultrasonography com-
pared with a 93% sensitivity and 99% NPV for DxWBS and rhTSH-stimulated
Tg measurements (261).

Serum Anti-Tg Antibodies


Serum antithyroglobulin antibodies must be measured in the serum sample
obtained for Tg assay because they are present in up to 25% patients with thyroid
carcinoma and almost always invalidate serum Tg measurement (181,262). These
antibodies must be quantitated because they can serve as a surrogate marker for
Tg, rising when there is an exacerbation of tumor and falling when the tumor
burden declines (263).

Imaging Studies
Although imaging with 131 I is the “gold standard” in detecting thyroid tissue,
several other scanning techniques are available. Some are particularly useful in
identifying the location of tumor in patients with high serum Tg levels and negative
diagnostic 131 I scans and negative neck ultrasonography.

Whole-Body Positron Emission Tomography


Scanning with F-18-fluorodeoxyglucose (FDG) provides two important pieces of
information. First, it may identify DTC metastasis that cannot be identified by
scintigraphy with 131 I or 99m Tc. Second, Fluorine-18-FDG uptake is an indicator
of poor functional differentiation and poor prognosis in thyroid cancer (264). A
retrospective study of 400 patients with thyroid cancer studied at one institution
(265) found an inverse relationship between patient survival and the glycolytic
rate of the most active lesion. Likewise, the number of FDG-avid lesions was
inversely correlated with survival (265). The likelihood of observing an FDG-
avid lesion increases with the serum Tg level, especially when the serum Tg
is ⬎10 ng/mL (266). False-positive 18 F-FDG uptake may occur with benign
lung disease, inflammatory conditions, and other malignancies (267). Positron
Differentiated Thyroid Carcinoma 279

Emission Tomography (PET) scanning may be of most value in the setting of high
serum Tg levels and negative neck ultrasonography and other imaging studies.

Treatment of Patients with High Serum Tg Levels and Negative


Imaging Studies
When the serum Tg level is elevated and a tumor cannot be found by localizing
techniques—including 131 I diagnostic scans, neck ultrasonography, and CT or
MRI scans—metastases are sometimes found only after administrating therapeutic
doses of 131 I.

Tg Cutoffs for Treatment with 131 I


A serum Tg above the lower detection limit (usually 0.2–1.0 ng/mL in newer
assays) during levothyroxine therapy in a patient who has undergone total or near-
total thyroidectomy and 131 I ablation is a sign of persistent normal tissue (thyroid
remnant) or DTC. This is an indication for repeat 131 I scanning in high-risk patients
or neck ultrasonography alone in low-risk patients when there is no other evidence
of disease (Fig. 9). If serum Tg rises ⬎10 ng/mL after levothyroxine is discontinued
or rises ⬎5 ng/mL after rhTSH is administered, normal or malignant thyroid tissue
is usually present, even if a 2 to 4 mCi (74–148 MBq) 131 I diagnostic scan is
negative (⬍1% 131 I uptake) (159,192). A serum Tg that is rising over time and
increases to ⬎10 ng/mL after rhTSH stimulation or THW may be an indication
of persistent tumor. If tumor is not identified by ultrasound and other imaging
modalities, it is reasonable to give one therapeutic dose of 131 I, usually 100 mCi,
and to perform a posttreatment scan. Depending upon the age of the patient, as
many as 20% of children and young adults with greatly elevated serum Tg levels
and negative diagnostic 131 I scans have lung metastases. When the RxWBS is
negative, an 18 FDG-PET/CT scan may detect occult tumor that is amenable to
surgical excision. Even if it does not identify tumor, this is a good prognostic sign
(265) that can reassure the patient.

Rationale for 131 I Therapy


Although some skepticism has been voiced about empirically treating patients with
100 mCi of 131 I without imaging evidence of disease, there is increasing evidence
that this approach is beneficial to some patients. A multivariate analysis has shown
the independent prognostic significance of the size of pulmonary metastases at the
time of therapy (268). Another multivariate analysis of prognostic factors in 134
patients with pulmonary metastases showed that an early diagnosis (a normal chest
roentgenogram with pulmonary metastases found only on 131 I scintigraphy) and
treatment of the metastases with 131 I were the most important elements giving
rise to a significant improvement in survival rate and a prolonged disease-free
time interval (269). Two studies (270,271) found clearly beneficial effects of 131 I
therapy for such patients: 80% achieved a negative whole-body 131 I posttherapy
scan, 60% had a serum Tg ⬍5 ng/mL off thyroid hormone and six of eight patients
280 Sipos and Mazzaferri

had normalization of the CT scan, and two patients had negative lung biopsies.
Improvement sometimes occurs with one or two 131 I treatments, but complete
resolution of pulmonary metastases after 131 I therapy is often difficult to achieve
(137). When a partial reduction of metastatic disease is achieved, patients usually
have a good quality of life with no further disease progression and a low mortality
rate (137). It seems intuitively wrong to withhold therapy in this group of patients
who are usually young and have a small tumor burden. Withholding therapy seems
especially harsh given the fact that 131 I treatment directly targets the metastatic
deposits and is effective in reducing tumor burden. Early diagnosis and treatment of
cancer is a desirable and effective therapeutic goal, especially when the treatment
has relatively few serious side effects. On the other hand treating patients on the
basis of a rising Tg when distant metastases are not found on a previous RxWBS
on the basis of an elevated serum Tg level is not advisable. Some patients are
found in retrospect to have regional lymph node metastases or a thyroid remnant
not seen on the diagnostic 131 I scan to account for the high serum Tg level. In
other cases, the Tg begins to gradually fall spontaneously without treatment or
following an apparently unsuccessful treatment. Giving repeated 131 I treatments
without clear evidence of efficacy is strongly discouraged.

REFERENCES
1. American Cancer Society. Cancer Facts and Figures. American Cancer Society.
9–4–2007.
2. Edwards BK, Brown ML, Wingo PA, et al. Annual report to the nation on the status
of cancer,1975–2002, featuring population-based trends in cancer treatment. J Natl
Cancer Inst 2005; 97(19):1407–1427.
3. Hundahl SA, Fleming ID, Fremgen AM, et al. A National Cancer Data Base report
on 53,856 cases of thyroid carcinoma treated in the US,1985–1995. Cancer 1998;
83:2638–2648.
4. Davies L, Welch HG. Increasing incidence of thyroid cancer in the United
States,1973–2002. JAMA 2006; 295(18):2164–2167.
5. Hayat MJ, Howlader N, Reichman ME, et al. Cancer statistics, trends, and multiple
primary cancer analyses from the Surveillance, Epidemiology, and End Results
(SEER) Program. Oncologist 2007; 12(1):20–37.
6. Mazzaferri EL. Managing small thyroid cancers. JAMA 2006; 295(18):2179–2182.
7. Ries LAG, Harkins D, Krapcho D, et al. SEER Cancer Statistics Review,1997–2003
Based on November 2005 SEER data submission, posted to the SEER web site,
2006. SEER Surveillance Epidemiology and End Results Cancer Stat Fact Sheets,
2006:6–13.
8. World Health Organization. Pathology & Genetics Tumours of Endocrine Organs.
Lyon, Farnce: IARC Press, 2004.
9. Begum S, Rosenbaum E, Henrique R, et al. BRAF mutations in anaplastic thyroid
carcinoma: Implications for tumor origin, diagnosis and treatment. Mod Pathol 2004;
17(11):1359–1363.
10. Bevan S, Pal T, Greenberg CR, et al. A comprehensive analysis of MNG1, TCO1,
fPTC, PTEN, TSHR, and TRKA in familial nonmedullary thyroid cancer: Confir-
mation of linkage to TCO1. J Clin Endocrinol Metab 2001; 86(8):3701–3704.
Differentiated Thyroid Carcinoma 281

11. Uchino S, Noguchi S, Kawamoto H, et al. Familial nonmedullary thyroid carcinoma


characterized by multifocality and a high recurrence rate in a large study population.
World J Surg 2002; 26(8):897–902.
12. Mazzaferri EL. Thyroid carcinoma: Papillary and follicular. In: Mazzaferri EL,
Samaan N, eds. Endocrine Tumors. Cambridge, MA: Blackwell Scientific Publica-
tions Inc, 1993:278–333.
13. Trimboli P, Ulisse S, Graziano FM, et al. Trend in thyroid carcinoma size, age at
diagnosis, and histology in a retrospective study of 500 cases diagnosed over 20
years. Thyroid 2006; 16(11):1151–1155.
14. Machens A, Holzhausen HJ, Dralle H. The prognostic value of primary tumor
size in papillary and follicular thyroid carcinoma. Cancer 2005; 103(11):2269–
2273.
15. Brennan MD, Bergstralh EJ, van Heerden JA, et al. Follicular thyroid cancer treated
at the Mayo Clinic,1946 through 1970: Initial manifestations, pathologic findings,
therapy, and outcome. Mayo Clin Proc 1991; 66:11–22.
16. Samaan NA, Schultz PN, Ordonez NG, et al. A comparison of thyroid carcinoma
in those who have and have not had head and neck irradiation in childhood. J Clin
Endocrinol Metab 1987; 64:219–223.
17. Schneider AB, Ron E, Lubin J, et al. Dose-response relationships for radiation-
induced thyroid cancer and thyroid nodules: Evidence for the prolonged effects of
radiation on the thyroid. J Clin Endocrinol Metab 1993; 77:362–369.
18. Somerville HM, Steinbeck KS, Stevens G, et al. Thyroid neoplasia following irradi-
ation in adolescent and young adult survivors of childhood cancer. Med J Aust 2002;
176(12):584–587.
19. Ron E, Lubin JH, Shore RE, et al. Thyroid cancer after exposure to exter-
nal radiation: A pooled analysis of seven studies. Radiat Res 1995; 141:259–
277.
20. Moreno A, Rodriguez JM, Sola J, et al. Encapsulated papillary neoplasm of the
thyroid: Retrospective clinicopathological study with long term follow up. Eur J
Surg 1996; 162:177–180.
21. Schroder S, Bocker W, Dralle H, et al. The encapsulated papillary carcinoma of the
thyroid. A morphologic subtype of the papillary thyroid carcinoma. Cancer 1984;
54:90–93.
22. Massoll N, Nizam MS, Mazzaferri EL. Cystic thyroid nodules: Diagnostic and
therapeutic dilemmas. Endocrinologist 2002; 12(3):185–198.
23. Bui A, Mazzaferri E, Massoll N. Cystic papillary thyroid cancer. Thyroid 2006;
16(12):1319–1320.
24. Mazzaferri EL. Management of low risk differentiated thyroid cancer. Endocr Pract
2007; 13(6):498–512.
25. LiVolsi VA. Papillary lesions of the thyroid. In: LiVolsi VA, ed. Surgical Pathology
of the Thyroid. Philadelphia, PA: W.B.Saunders Company, 1990:136–172.
26. Katoh R, Sasaki J, Kurihara H, et al. Multiple thyroid involvement (intraglandu-
lar metastasis) in papillary thyroid carcinoma. A clinicopathologic study of 105
consecutive patients. Cancer 1992; 70(6):1585–1590.
27. Chow SM, Law SC, Chan JK, et al. Papillary microcarcinoma of the thyroid-
Prognostic significance of lymph node metastasis and multifocality. Cancer 2003;
98(1):31–40.
28. Baudin E, Travagli JP, Ropers J, et al. Microcarcinoma of the thyroid gland: The
Gustave-Roussy Institute experience. Cancer 1998; 83(3):553–559.
282 Sipos and Mazzaferri

29. Shattuck TM, Westra WH, Ladenson PW, et al. Independent clonal origins of dis-
tinct tumor foci in multifocal papillary thyroid carcinoma. N Engl J Med 2005;
352(23):2406–2412.
30. Giannini R, Ugolini C, Lupi C, et al. The heterogeneous distribution of BRAF
mutation supports the independent clonal origin of distinct tumor foci in multifocal
papillary thyroid carcinoma. J Clin Endocrinol Metab 2007; 92(9):3511–3516.
31. Park SY, Park YJ, Lee YJ, et al. Analysis of differential BRAF(V600E) mutational
status in multifocal papillary thyroid carcinoma: Evidence of independent clonal
origin in distinct tumor foci. Cancer 2006; 107(8):1831–1838.
32. Sugg SL, Ezzat S, Rosen IB, et al. Distinct multiple RET/PTC gene rearrangements
in multifocal papillary thyroid neoplasia. J Clin Endocrinol Metab 1998; 83:4116–
4122.
33. McCarthy RP, Wang M, Jones TD, et al. Molecular evidence for the same clonal ori-
gin of multifocal papillary thyroid carcinomas. Clin Cancer Res 2006; 12(8):2414–
2418.
34. Schäffler A, Palitzsch KD, Seiffarth C, et al. Coexistent thyroiditis is associated
with lower tumour stage in thyroid carcinoma. Eur J Clin Invest 1998; 28:838–
844.
35. Fiumara A, Belfiore A, Russo G, et al. In situ evidence of neoplastic cell phagocytosis
by macrophages in papillary thyroid cancer. J Clin Endocrinol Metab 1997; 82:1615–
1620.
36. Loh KC, Greenspan FS, Dong F, et al. Influence of lymphocytic thyroiditis on the
prognostic outcome of patients with papillary thyroid carcinoma. J Clin Endocrinol
Metab 1999; 84:458–463.
37. Mazzaferri EL, Jhiang SM. Long-term impact of initial surgical and medical therapy
on papillary and follicular thyroid cancer. Am J Med 1994; 97:418–428.
38. Noguchi M, Yamada H, Ohta N, et al. Regional lymph node metastases in well-
differentiated thyroid carcinoma. Int Surg 1987; 72:100–103.
39. Qubain SW, Nakano S, Baba M, et al. Distribution of lymph node micrometastasis
in pN0 well-differentiated thyroid carcinoma. Surgery 2002; 131(3):249–256.
40. Mirallie E, Visset J, Sagan C, et al. Localization of cervical node metastasis of
papillary thyroid carcinoma. World J Surg 1999; 23(9):970–973.
41. Arch-Ferrer J, Velazquez D, Fajardo R, et al. Accuracy of sentinel lymph node in
papillary thyroid carcinoma. Surgery 2001; 130(6):907–913.
42. Wada N, Duh QY, Sugino K, et al. Lymph node metastasis from 259 papillary
thyroid microcarcinomas: Frequency, pattern of occurrence and recurrence, and
optimal strategy for neck dissection. Ann Surg 2003; 237(3):399–407.
43. Hay ID. Management of patients with low-risk papillary thyroid carcinoma. Endocr
Pract 2007; 13(5):521–533.
44. Yamashita H, Noguchi S, Murakami N, et al. Extracapsular invasion of lymph node
metastasis is an indicator of distant metastasis and poor prognosis in patients with
thyroid papillary carcinoma. Cancer 1997; 80(12):2268–2272.
45. Durante C, Haddy N, Baudin E, et al. Long term outcome of 444 patients with distant
metastases from papillary and follicular thyroid carcinoma: Benefits and limits of
radioiodine therapy. J Clin Endocrinol Metab 2006; 92(2):450–455.
46. Bal CS, Kumar A, Chandra P, et al. Is chest x-ray or high-resolution computed
tomography scan of the chest sufficient investigation to detect pulmonary metastasis
in pediatric differentiated thyroid cancer? Thyroid 2004; 14(3):217–225.
Differentiated Thyroid Carcinoma 283

47. Pellegriti G, Scollo C, Lumera G, et al. Clinical behavior and outcome of papillary
thyroid cancers smaller than 1.5 cm in diameter: Study of 299 cases. J Clin Endocrinol
Metab 2004; 89(8):3713–3720.
48. Sugitani I, Fujimoto Y. Symptomatic versus asymptomatic papillary thyroid micro-
carcinoma: A retrospective analysis of surgical outcome and prognostic factors.
Endocr J 1999; 46(1):209–216.
49. Mazzaferri EL. Thyroid cancer in thyroglossal duct remnants: A diagnostic and
therapeutic dilemma. Thyroid 2004; 14(5):335–336.
50. Doshi SV, Cruz RM, Hilsinger RL Jr. Thyroglossal duct carcinoma: A large case
series. Ann Otol Rhinol Laryngol 2001; 110(8):734–738.
51. Patel SG, Escrig M, Shaha AR, et al. Management of well-differentiated thyroid
carcinoma presenting within a thyroglossal duct cyst. J Surg Oncol 2002; 79(3):134–
139.
52. Kurzen F, Flugel W, Brauer CF, et al. Papillary carcinoma of the thyroid in a
thyroglossal duct cyst with metastasis to an ipsilateral cervical lymph node. HNO
1999; 47(8):741–744.
53. Luna-Ortiz K, Hurtado-Lopez LM, Valderrama-Landaeta JL, et al. Thyroglossal duct
cyst with papillary carcinoma: What must be done? Thyroid 2004; 14(5):363–366.
54. Miccoli P, Pacini F, Basolo S, et al. Thyroid carcinoma in a thyroglossal duct cyst:
Tumor resection alone or a total thyroidectomy?. Ann Chir 1998; 52(5):452–454.
55. Moncet D, Manavela M, Cross GE, et al. Papillary carcinoma in thyroglossal duct
cyst. Endocr Pract 2001; 7(6):463–466.
56. Burningham AR, Krishnan J, Davidson BJ, et al. Papillary and follicular variant
of papillary carcinoma of the thyroid: Initial presentation and response to therapy.
Otolaryngol Head Neck Surg 2005; 132(6):840–844.
57. Passler C, Prager G, Scheuba C, et al. Follicular variant of papillary thyroid carci-
noma: A long-term follow-up. Arch Surg 2003; 138(12):1362–1366.
58. Sebastian SO, Gonzalez JM, Paricio PP, et al. Papillary thyroid carcinoma: Prognostic
index for survival including the histological variety. Arch Surg 2000; 135(3):272–
277.
59. Chen KT, Rosai J. Follicular variant of thyroid papillary carcinoma: A clinicopatho-
logic study of six cases. Am J Surg Pathol 1977; 1(2):123–130.
60. LiVolsi VA. Pure versus follicular variant of papillary thyroid carcinoma: Clinical
features, prognostic factors, treatment, and survival. Cancer 2003; 98(9):1997–1998.
61. Zidan J, Karen D, Stein M, et al. Pure versus follicular variant of papillary thyroid
carcinoma. Cancer 2003; 97(5):1181–1185.
62. Carcangiu ML, Zampi G, Pupi A, et al. Papillary carcinoma of the thyroid. A
clinicopathologic study of 241 cases treated at the University of Florence, Italy.
Cancer 1985; 55:805–828.
63. Chang HY, Lin JD, Chou SC, et al. Clinical presentations and outcomes of surgical
treatment of follicular variant of the papillary thyroid carcinomas. Jpn J Clin Oncol
2006; 36(11):688–693.
64. Evans HL. Follicular neoplasms of the thyroid. A study of 44 cases followed for
a minimum of 10 years, with emphasis on differential diagnosis. Cancer 1984;
54(3):535–540.
65. Gamboa-Dominguez A, Tenorio-Villalvazo A. Metastatic follicular variant of papil-
lary thyroid carcinoma manifested as a primary renal neoplasm. Endocr Pathol 1999;
10(3):256–268.
284 Sipos and Mazzaferri

66. Baloch ZW, LiVolsi VA. Encapsulated follicular variant of papillary thyroid carci-
noma with bone metastases. Mod Pathol 2000; 13(8):861–865.
67. Smit JW, Zelderen-Bhola S, Merx R, et al. A novel chromosomal translocation
t(3;5)(q12;p15.3) and loss of heterozygosity on chromosome 22 in a multifocal
follicular variant of papillary thyroid carcinoma presenting with skin metastases.
Clin Endocrinol (Oxf) 2001; 55(4):543–548.
68. Fadda G, Fiorino MC, Mule A, et al. Macrofollicular encapsulated variant of papil-
lary thyroid carcinoma as a potential pitfall in histologic and cytologic diagnosis. A
report of three cases. Acta Cytol 2002; 46(3):555–559.
69. Chung D, Ghossein RA, Lin O. Macrofollicular variant of papillary carcinoma: A
potential thyroid FNA pitfall. Diagn Cytopathol 2007; 35(9):560–564.
70. LiVolsi VA. Unusual variants of papillary thyroid carcinoma. In: Mazzaferri EL,
Kreisberg RA, Bar RS, eds. Advances in Endocrinology and Metabolism. St. Louis,
MO: Mosby-Year Book Inc, 1995:39–54.
71. Ivanova R, Soares P, Castro P, et al. Diffuse (or multinodular) follicular variant of
papillary thyroid carcinoma: A clinicopathologic and immunohistochemical analysis
of ten cases of an aggressive form of differentiated thyroid carcinoma. Virchows Arch
2002; 440(4):418–424.
72. Ain KB. Papillary thyroid carcinoma etiology, assessment, and therapy. Endocrinol
Metab Clin North Am 1995; 24:711–760.
73. Burman KD, Ringel MD, Wartofsky L. Unusual types of thyroid neoplasms.
Endocrinol Metab Clin North Am 1996; 25(1):49–68.
74. Nikiforova MN, Kimura ET, Gandhi M, et al. BRAF mutations in thyroid tumors are
restricted to papillary carcinomas and anaplastic or poorly differentiated carcinomas
arising from papillary carcinomas. J Clin Endocrinol Metab 2003; 88(11):5399–
5404.
75. Frattini M, Ferrario C, Bressan P, et al. Alternative mutations of BRAF, RET and
NTRK1 are associated with similar but distinct gene expression patterns in papillary
thyroid cancer. Oncogene 2004; 23(44):7436–7440.
76. Nikiforov Y, Gnepp DR. Pediatric thyroid cancer after the Chernobyl disaster: Path-
omorphologic study of 84 cases (1991–1992) from the Republic of Belarus. Cancer
1994; 74:748–766.
77. Kumarasinghe MP. Cytomorphologic features of diffuse sclerosing variant of papil-
lary carcinoma of the thyroid—A report of two cases in children. Acta Cytol 1998;
42:983–986.
78. Chow SM, Chan JK, Law SC, et al. Diffuse sclerosing variant of papillary thyroid
carcinoma—clinical features and outcome. Eur J Surg Oncol 2003; 29(5):446–449.
79. Falvo L, Giacomelli L, D’Andrea V, et al. Prognostic importance of sclerosing variant
in papillary thyroid carcinoma. Am Surg 2006; 72(5):438–444.
80. Lam AK, Lo CY. Diffuse sclerosing variant of papillary carcinoma of the thyroid: A
35-year comparative study at a single institution. Ann Surg Oncol 2006; 13(2):176–
181.
81. Rosai J. Papillary carcinoma. Monogr Pathol 1993; 138–165.
82. Herrera MF, Hay ID, Wu PS, et al. Hurthle cell (oxyphyilic) papillary thyroid
carcinoma: A variant with more aggressive biologic behavior. World J Surg 1994;
16:669–674.
83. Katoh R, Harach HR, Williams ED. Solitary, multiple, and familial oxyphil tumours
of the thyroid gland. J Pathol 1998; 186:292–299.
Differentiated Thyroid Carcinoma 285

84. LiVolsi VA. Follicular lesions of the thyroid. In: LiVolsi VA, ed. Surgical Pathology
of the Thyroid. Philadelphia, PA: W.B.Saunders Company, 1990:173–212.
85. Hassoun AAK, Hay ID, Goellner JR, et al. Insular thyroid carcinoma in
adolescents—A potentially lethal endocrine malignancy. Cancer 1997; 79:1044–
1048.
86. Mazzaferri EL. Management of a solitary thyroid nodule. N Engl J Med 1993;
328:553–559.
87. LiVolsi V, Asa SL. The demise of follicular carcinoma of the thyroid gland. Thyroid
1994; 4:233–236.
88. Lin JD, Chao TC, Ho J, et al. Poor prognosis of 56 follicular thyroid carcinomas
with distant metastases at the time of diagnosis. Cancer J 1998; 11:190–195.
89. Leenhardt L, Bernier MO, Boin-Pineau MH, et al. Advances in diagnostic practices
affect thyroid cancer incidence in France. Eur J Endocrinol 2004; 150(2):133–139.
90. Hamming JF, Goslings BM, vanSteenis GJ, et al. The value of fine-needle aspiration
biopsy in patients with nodular thyroid disease divided into groups of suspicion of
malignant neoplasms on clinical grounds. Arch Intern Med 1990; 150:113–116.
91. Schneider AB, Shore Freedman E, Ryo UY, et al. Radiation-induced tumors of
the head and neck following childhood irradiation. Prospective studies. Medicine
(Baltimore) 1985; 64:1–15.
92. Hegedus L. Clinical practice. The thyroid nodule. N Engl J Med 2004; 351(17):1764–
1771.
93. Cooper DS, Doherty GM, Haugen BR, et al. Management guidelines for patients with
thyroid nodules and differentiated thyroid cancer. Thyroid 2006; 16(2):109–141.
94. Redman R, Zalaznick H, Mazzaferri EL, et al. The impact of assessing specimen
adequacy and number of needle passes for fine-needle aspiration biopsy of thyroid
nodules. Thyroid 2006; 16(1):55–60.
95. Asanuma K, Kobayashi S, Shingu K, et al. The rate of tumour growth does not distin-
guish between malignant and benign thyroid nodules. Eur J Surg 2001; 167(2):102–
105.
96. Renshaw AA, Pinnar N. Comparison of thyroid fine-needle aspiration and core
needle biopsy. Am J Clin Pathol 2007; 128(3):370–374.
97. Leenhardt L, Hejblum G, Franc B, et al. Indications and limits of ultrasound-guided
cytology in the management of nonpalpable thyroid nodules. J Clin Endocrinol
Metab 1999; 84:24–28.
98. Illes A, Biro E, Miltenyi Z, et al. Hypothyroidism and thyroiditis after therapy for
Hodgkin’s disease. Acta Haematol 2003; 109(1):11–17.
99. Atahan IL, Yildiz F, Ozyar E, et al. Thyroid dysfunction in children receiving neck
irradiation for Hodgkin’s disease. Radiat Med 1998; 16(5):359–361.
100. Fleming ID, Black TL, Thompson EI, et al. Thyroid dysfunction and neoplasia in
children receiving neck irradiation for cancer. Cancer 1985; 55(6):1190–1194.
101. Schneider AB, Bekerman C, Leland J, et al. Thyroid nodules in the follow-up of irra-
diated individuals: Comparison of thyroid ultrasound with scanning and palpation.
J Clin Endocrinol Metab 1997; 82:4020–4027.
102. Ezzat S, Sarti DA, Cain DR, et al. Thyroid incidentalomas: Prevalence by palpation
and ultrasonography. Arch Intern Med 1994; 154:1838–1840.
103. Tan GH, Gharib H. Thyroid incidentalomas: Management approaches to nonpal-
pable nodules discovered incidentally on thyroid imaging. Ann Intern Med 1997;
126(3):226–231.
286 Sipos and Mazzaferri

104. Brauer VF, Eder P, Miehle K, et al. Interobserver variation for ultrasound determi-
nation of thyroid nodule volumes. Thyroid 2005; 15(10):1169–1175.
105. Mazzaferri EL, Kloos RT. Current approaches to primary therapy for papillary and
follicular thyroid cancer. J Clin Endocrinol Metab 2001; 86(4):1447–1463.
106. Dottorini ME, Vignati A, Mazzucchelli L, et al. Differentiated thyroid carcinoma in
children and adolescents: A 37-year experience in 85 patients. J Nucl Med 1997;
38:669–675.
107. Hung W. Well-differentiated thyroid carcinomas in children and adolescents: A
review. Endocrinologist 1994; 4:117–126.
108. Harach HR, Williams ED. Childhood thyroid cancer in England and Wales. Br J
Cancer 1995; 72:777–783.
109. Young RL, Mazzaferri EL, Rahe AJ, et al. Pure follicular thyroid carcinoma: Impact
of therapy in 214 patients. J Nucl Med 1980; 21:733–737.
110. Kosary CL, Ries LAG, Miller BA, et al. SEER Cancer Statistic Review,1973–1992.
In: Kosary CL, Ries LAG, Miller BA, Hankey BF, Harras A, Edwards BK, eds.
Tables and Graphs. Bethesda, MD: National Cancer Institute. NIH Pub. No., 1995:
96–2789.
111. Filetti S, Belfiore A, Amir SM, et al. The role of thyroid-stimulating antibodies of
Graves’ disease in differentiated thyroid cancer. N Engl J Med 1988; 318:753–779.
112. Belfiore A, Garofalo MR, Giuffrida D, et al. Increased aggressiveness of thyroid
cancer in patients with Graves’ disease. J Clin Endocrinol Metab 1990; 70:830–835.
113. Hay ID. Papillary thyroid carcinoma. Endocrinol Metab Clin North Am 1990;
19(3):545–576.
114. Yano Y, Shibuya H, Kitagawa W, et al. Recent outcome of Graves’ disease patients
with papillary thyroid cancer. Eur J Endocrinol 2007; 157(3):325–329.
115. Samaan NA, Schultz PN, Hickey RC, et al. Well-differentiated thyroid carcinoma
and the results of various modalities of treatment. A retrospective review of 1599
patients. J Clin Endocrinol Metab 1992; 75:714–720.
116. Machens A, Holzhausen HJ, Lautenschlager C, et al. Enhancement of lymph node
metastasis and distant metastasis of thyroid carcinoma. Cancer 2003; 98(4):712–719.
117. Pazaitou-Panayiotou K, Capezzone M, Pacini F. Clinical features and therapeutic
implication of papillary thyroid microcarcinoma. Thyroid 2007; 17(11):1085–1092.
118. Pacini F, Elisei R, Capezzone M, et al. Contralateral papillary thyroid cancer is
frequent at completion thyroidectomy with no difference in low- and high-risk
patients. Thyroid 2001; 11(9):877–881.
119. Alzahrani AS, Al Mandil M, Chaudhary MA, et al. Frequency and predictive factors
of malignancy in residual thyroid tissue and cervical lymph nodes after partial
thyroidectomy for differentiated thyroid cancer. Surgery 2002; 131(4):443–449.
120. DeGroot LJ, Kaplan EL, McCormick M, et al. Natural history, treatment, and course
of papillary thyroid carcinoma. J Clin Endocrinol Metab 1990; 71:414–424.
121. Mazzaferri EL. Thyroid remnant 131 I ablation for papillary and follicular thyroid
carcinoma. Thyroid 1997; 7:265–271.
122. Sarda AK, Bal S, Kapur MM. Near-total thyroidectomy for carcinoma of the thyroid.
Br J Surg 1989; 76:90–92.
123. Loh KC, Greenspan FS, Gee L, et al. Pathological tumor-node-metastasis (pTNM)
staging for papillary and follicular thyroid carcinomas: A retrospective analysis of
700 patients. J Clin Endocrinol Metab 1997; 82(11):3553–3562.
Differentiated Thyroid Carcinoma 287

124. Sanders LE, Cady B. Differentiated thyroid cancer—Reexamination of risk groups


and outcome of treatment. Arch Surg 1998; 133:419–424.
125. Shaha AR, Loree TR, Shah JP. Prognostic factors and risk group analysis in follicular
carcinoma of the thyroid. Surgery 1995; 118:1131–1138.
126. Solorzano CC, Carneiro DM, Ramirez M, et al. Surgeon-performed ultrasound in
the management of thyroid malignancy. Am Surg 2004; 70(7):576–580.
127. Kouvaraki MA, Shapiro SE, Fornage BD, et al. Role of preoperative ultrasonog-
raphy in the surgical management of patients with thyroid cancer. Surgery 2003;
134(6):946–954.
128. Karwowski JK, Jeffrey RB, McDougall IR, et al. Intraoperative ultrasonography
improves identification of recurrent thyroid cancer. Surgery 2002; 132(6):924–
929.
129. Kumar A, Bal CS. Differentiated thyroid cancer. Indian J Pediatr 2003; 70(9):707–
713.
130. La Quaglia MP, Black T, Holcomb GW III, et al. Differentiated thyroid cancer:
Clinical characteristics, treatment, and outcome in patients under 21 years of age who
present with distant metastases. A report from the Surgical Discipline Committee of
the Children’s Cancer Group. J Pediatr Surg 2000; 35(6):955–959.
131. Cady B. Staging in thyroid carcinoma. Cancer 1998; 83:844–847.
132. Tsang RW, Brierley JD, Simpson WJ, et al. The effects of surgery, radioiodine, and
external radiation therapy on the clinical outcome of patients with differentiated
thyroid carcinoma. Cancer 1998; 82(2):375–388.
133. Kurozumi K, Nakao I, Nishida T, et al. Significance of biologic aggressiveness and
proliferating activity in papillary thyroid carcinoma. World J Surg 1998; 22:1237–
1242.
134. Sugitani I, Yanagisawa A, Shimizu A, et al. Clinicopathologic and immunohis-
tochemical studies of papillary thyroid microcarcinoma presenting with cervical
lymphadenopathy. World J Surg 1998; 22:731–737.
135. Chiang FY, Lin JC, Lee KW, et al. Thyroid tumors with preoperative recurrent
laryngeal nerve palsy: Clinicopathologic features and treatment outcome. Surgery
2006; 140(3):413–417.
136. Sampson E, Brierley JD, Le LW, et al. Clinical management and outcome of papillary
and follicular (differentiated) thyroid cancer presenting with distant metastasis at
diagnosis. Cancer 2007; 110(7):1451–1456.
137. Samuel AM, Rajashekharrao B, Shah DH. Pulmonary metastases in children and
adolescents with well-differentiated thyroid cancer. J Nucl Med 1998; 39:1531–
1536.
138. Sisson JC, Giordano TJ, Jamadar DA, et al. 131-I treatment of micronodular pul-
monary metastases from papillary thyroid carcinoma. Cancer 1996; 78:2184–2192.
139. Nêmec J, Zamrazil V, Pohunková D, et al. Radioiodide treatment of pulmonary
metastases of differentiated thyroid cancer. Results and prognostic factors. Nuk-
learmedizin 1979; 18:86–90.
140. Unger K, Zitzelsberger H, Salvatore G, et al. Heterogeneity in the Distribution
of RET/PTC Rearrangements within Individual Post-Chernobyl Papillary Thyroid
Carcinomas. J Clin Endocrinol Metab 2004; 89(9):4272–4279.
141. Rabes HM, Demidchik EP, Sidorow JD, et al. Pattern of radiation-induced RET
and NTRK1 rearrangements in 191 post-chernobyl papillary thyroid carcinomas:
288 Sipos and Mazzaferri

Biological, phenotypic, and clinical implications. Clin Cancer Res 2000; 6(3):1093–
1103.
142. Thomas GA, Bunnell H, Cook HA, et al. High prevalence of RET/PTC rearrange-
ments in Ukrainian and Belarussian post-Chernobyl thyroid papillary carcinomas:
A strong correlation between RET/PTC3 and the solid-follicular variant. J Clin
Endocrinol Metab 1999; 84:4232–4238.
143. Santoro M, Thomas GA, Vecchio G, et al. Gene rearrangement and Chernobyl related
thyroid cancers. Br J Cancer 2000; 82(2):315–322.
144. Elisei R, Romei C, Vorontsova T, et al. Ret/ptc rearrangements in thyroid nodules:
Studies in irradiated and not irradiated, malignant and benign thyroid lesions in
children and adults. J Clin Endocrinol Metab 2001; 86(7):3211–3216.
145. Viswanathan K, Gierlowski TC, Schneider AB. Childhood thyroid cancer: Charac-
teristics and long-term outcome in children irradiated for benign conditions of the
head and neck. Am J Dis Child 1994; 148:260–265.
146. Kashima K, Yokoyama S, Noguchi S, et al. Chronic thyroiditis as a favorable prog-
nostic factor in papillary thyroid carcinoma. Thyroid 1998; 8:197–202.
147. Krause DS, Van Etten RA. Tyrosine kinases as targets for cancer therapy. N Engl J
Med 2005; 353(2):172–187.
148. Puxeddu E, Moretti S, Elisei R, et al. BRAF (V599E) mutation is the leading genetic
event in adult sporadic papillary thyroid carcinomas. J Clin Endocrinol Metab 2004;
89(5):2414–2420.
149. Fugazzola L, Puxeddu E, Avenia N, et al. Correlation between B-RAFV600E muta-
tion and clinico-pathologic parameters in papillary thyroid carcinoma: Data from a
multicentric Italian study and review of the literature. Endocr Relat Cancer 2006;
13(2):455–464.
150. Xing M, Westra WH, Tufano RP, et al. BRAF mutation predicts a poorer clinical
prognosis for papillary thyroid cancer. J Clin Endocrinol Metab 2006; 90(12):6373–
6379.
151. Fugazzola L, Mannavola D, Cirello V, et al. BRAF mutations in an Italian cohort of
thyroid cancers. Clin Endocrinol (Oxf) 2004; 61(2):239–243.
152. Brierley JD, Panzarella T, Tsang RW, et al. A comparison of different staging systems
predictability of patient outcome—Thyroid carcinoma as an example. Cancer 1997;
79:2414–2423.
153. Sherman SI, Brierley JD, Sperling M, et al. Prospective multicenter study of thy-
roid carcinoma treatment—Initial analysis of staging and outcome. Cancer 1998;
83:1012–1021.
154. American Joint Committee on Cancer. AJCC Comparison Guide: Cancer Staging
Manual. American Joint Committee on Cancer. 2005:8–15.
155. Pacini F, Schlumberger M, Dralle H, et al. European consensus for the management
of patients with differentiated thyroid carcinoma of the follicular epithelium. Eur J
Endocrinol 2006; 154(6):787–803.
156. Pasieka JL, Rotstein LE. Consensus conference on well-differentiated thyroid cancer:
A summary. Can J Surg 1999; 36:298–301.
157. Moosa M, Mazzaferri EL. Occult thyroid carcinoma. Cancer J 1997; 10:180–188.
158. Cady B. Our AMES is true: How an old concept still hits the mark: Or, risk group
assignment points the arrow to rational therapy selection in differentiated thyroid
cancer. Am J Surg 1997; 174:462–468.
Differentiated Thyroid Carcinoma 289

159. Schlumberger MJ. Medical progress—Papillary and follicular thyroid carcinoma. N


Engl J Med 1998; 338:297–306.
160. Handkiewicz-Junak D, Wloch J, Roskosz J, et al. Total thyroidectomy and adjuvant
radioiodine treatment independently decrease locoregional recurrence risk in child-
hood and adolescent differentiated thyroid cancer. J Nucl Med 2007; 48(6):879–888.
161. Massin JP, Savoie JC, Garnier H, et al. Pulmonary metastases in differentiated thyroid
carcinoma. Study of 58 cases with implications for the primary tumor treatment.
Cancer 1984; 53:982–992.
162. Tourniaire J, Bernard MH, Bizollon-Roblin MH, et al. Papillary microcarcinoma of
the thyroid. 179 cases reported since 1973. Presse Med 1998; 27(29):1467–1469.
163. Hay ID, Bergstralh EJ, Goellner JR et al. Predicting outcome in papillary thyroid
carcinoma: Development of a reliable prognostic scoring system in a cohort of 1779
patients surgically treated at one institution during 1940 through 1989. Surgery 1993;
114(6):1050–1057.
164. Hay ID, Thompson GB, Grant CS, et al. Papillary thyroid carcinoma managed at
the Mayo Clinic during six decades (1940–1999): Temporal trends in initial therapy
and long-term outcome in 2444 consecutively treated patients. World J Surg 2002;
26(8):879–885.
165. Rouxel A, Hejblum G, Bernier MO, et al. Prognostic factors associated with the
survival of patients developing loco-regional recurrences of differentiated thyroid
carcinomas. J Clin Endocrinol Metab 2004; 89(11):5362–5368.
166. Hay ID, Bergstralh EJ, Grant CS, et al. Impact of primary surgery on outcome in 300
patients with pathologic tumor-node-metastasis stage III papillary thyroid carcinoma
treated at one institution from 1940 through 1989. Surgery 1999; 126(6):1173–
1181.
167. Maxon HR, Englaro EE, Thomas SR, et al. Radioiodine-131 therapy for well-
differentiated thyroid cancer—a quantitative radiation dosimetric approach: Out-
come and validation in 85 patients. J Nucl Med 1992; 33:1132–1136.
168. De Jong SA, Demeter JG, Lawrence AM, et al. Necessity and safety of completion
thyroidectomy for differentiated thyroid carcinoma. Surgery 1992; 112:734–737.
169. Chao TC, Jeng LB, Lin JD, et al. Completion thyroidectomy for differentiated thyroid
carcinoma. Otolaryngol Head Neck Surg 1998; 118:896–899.
170. Pasieka JL, Thompson NW, McLeod MK, et al. The incidence of bilateral well-
differentiated thyroid cancer found at completion thyroidectomy. World J Surg 1992;
16(4):711–716.
171. Miccoli P, Antonelli A, Spinelli C, et al. Completion total thyroidectomy in children
with thyroid cancer secondary to the Chernobyl accident. Arch Surg 1998; 133:89–
93.
172. Dralle H, Gimm O, Simon D, et al. Prophylactic thyroidectomy in 75 children and
adolescents with hereditary medullary thyroid carcinoma: German and Austrian
experience. World J Surg 1998; 22:744–751.
173. Pattou F, Combemale F, Fabre S, et al. Hypocalcemia following thyroid surgery:
Incidence and prediction of outcome. World J Surg 1998; 22:718–724.
174. Udelsman R, Lakatos E, Ladenson P. Optimal surgery for papillary thyroid carci-
noma. World J Surg 1996; 20:88–93.
175. Sosa JA, Bowman HM, Tielsch JM, et al. The importance of surgeon experience for
clinical and economic outcomes from thyroidectomy. Ann Surg 1998; 228:320–328.
290 Sipos and Mazzaferri

176. Moosa M, Mazzaferri EL. Outcome of differentiated thyroid cancer diagnosed in


pregnant women. J Clin Endocrinol Metab 1997; 82:2862–2866.
177. Saito T, Endo T, Kawaguchi A, et al. Increased expression of the Na+ /I − symporter
in cultured human thyroid cells exposed to thyrotropin and in Graves’ thyroid tissue.
J Clin Endocrinol Metab 1997; 82:3331–3336.
178. Jhiang SM, Cho JY, Ryu K-Y, et al. An immunohistochemical study of Na+/I-
symporter in human thyroid tissues and salivary gland tissues. Endocrinology 1998;
139:4416–4419.
179. Arturi F, Russo D, Schlumberger M, et al. Iodide symporter gene expression in
human thyroid tumors. J Clin Endocrinol Metab 1998; 83:2493–2496.
180. Vassilopoulou-Sellin R, Klein MJ, Smith TH, et al. Pulmonary metastases in children
and young adults with differentiated thyroid cancer. Cancer 1993; 71:1348–1352.
181. Spencer CA, Takeuchi M, Kazarosyan M, et al. Serum thyroglobulin autoantibodies:
Prevalence, influence on serum thyroglobulin measurement, and prognostic signif-
icance in patients with differentiated thyroid carcinoma. J Clin Endocrinol Metab
1998; 83:1121–1127.
182. Hodgson DC, Brierley JD, Tsang RW, et al. Prescribing 131 iodine based on neck
uptake produces effective thyroid ablation and reduced hospital stay. Radiother
Oncol 1998; 47:325–330.
183. Taylor T, Specker B, Robbins J, et al. Outcome after treatment of high-risk papillary
and non-Hurthle-cell follicular thyroid carcinoma. Ann Intern Med 1998; 129:622–
627.
184. Cunningham MP, Duda RB, Recant W, et al. Survival discriminants for differentiated
thyroid cancer. Am J Surg 1990; 160:344–347.
185. DeGroot LJ, Kaplan EL, Straus FH, et al. Does the method of management of
papillary thyroid carcinoma make a difference in outcome? World J Surg 1994;
18:123–130.
186. Samaan NA, Maheshwari YK, Nader S, et al. Impact of therapy for differentiated
carcinoma of the thyroid: An analysis of 706 cases. J Clin Endocrinol Metab 1983;
56:1131–1138.
187. Park JT, Hennessey JV. Two-week low iodine diet is necessary for adequate outpatient
preparation for radioiodine rhTSH scanning in patients taking levothyroxine. Thyroid
2004; 14(1):57–63.
188. Maini CL, Sciuto R, Tofani A. Delayed thyroid-stimulating hormone suppression by
L-thyroxine in the management of differentiated thyroid carcinoma. Eur J Cancer
[A] 1993; 29 A:2071–2072.
189. Leboeuf R, Perron P, Carpentier AC, et al. L-T(3) preparation for whole-body scintig-
raphy: A randomized-controlled trial. Clin Endocrinol (Oxf) 2007; 67(6):839–844.
190. Menzel C, Kranert WT, Dobert N, et al. rhTSH Stimulation before radioiodine
therapy in thyroid cancer reduces the effective half-life of (131)I. J Nucl Med 2003;
44(7):1065–1068.
191. Goldman JM, Line BR, Aamodt RL, et al. Influence of triiodothyronine withdrawal
time on 131I uptake postthyroidectomy for thyroid cancer. J Clin Endocrinol Metab
1980; 50(4):734–739.
192. Haugen BR, Pacini F, Reiners C, et al. A comparison of recombinant human thy-
rotropin and thyroid hormone withdrawal for the detection of thyroid remnant or
cancer. J Clin Endocrinol Metab 1999; 84:3877–3885.
Differentiated Thyroid Carcinoma 291

193. Pacini F, Ladenson PW, Schlumberger M, et al. Radioiodine ablation of thyroid


remnants after preparation with recombinant human thyrotropin in differentiated
thyroid carcinoma: Results of an international, randomized, controlled study. J Clin
Endocrinol Metab 2006; 91(3):926–932.
194. Meier CA, Braverman LE, Ebner SA, et al. Diagnostic use of recombinant human
thyrotropin in patients with thyroid carcinoma (phase I/II study). J Clin Endocrinol
Metab 1994; 78:188–196.
195. Ladenson PW, Braverman LE, Mazzaferri EL, et al. Comparison of administra-
tion of recombinant human thyrotropin with withdrawal of thyroid hormone for
radioactive iodine scanning in patients with thyroid carcinoma. N Engl J Med 1997;
337(13):888–896.
196. Torres MS, Ramirez L, Simkin PH, et al. Effect of various doses of recombinant
human thyrotropin on the thyroid radioactive iodine uptake and serum levels of
thyroid hormones and thyroglobulin in normal subjects. J Clin Endocrinol Metab
2001; 86(4):1660–1664.
197. Pilli T, Brianzoni E, Capoccetti F, et al. A comparison of 1850 MBq (50 mCi)
and 3700 MBq (100 mCi) 131-iodine administered doses for recombinant TSH-
stimulated postoperative thyroid remnant ablation in differentiated thyroid cancer. J
Clin Endocrinol Metab 2007; 92(9):3542–3546.
198. Schroeder PR, Haugen BR, Pacini F, et al. A comparison of short-term changes in
health-related quality of life in thyroid carcinoma patients undergoing diagnostic
evaluation with recombinant human thyrotropin compared with thyroid hormone
withdrawal. J Clin Endocrinol Metab 2006; 91(3):878–884.
199. Maxon HR, Thomas SR, Hertzberg VS, et al. Relation between effective radiation
dose and outcome of radioiodine therapy for thyroid cancer. N Engl J Med 1983;
309:937–941.
200. Medvedec M. Thyroid stunning in vivo and in vitro. Nucl Med Commun 2005;
26(8):731–735.
201. Morris LF, Waxman AD, Braunstein GD. Thyroid stunning. Thyroid 2003;
13(4):333–340.
202. Muratet JP, Giraud P, Daver A, et al. Predicting the efficacy of first iodine-131
treatment in differentiated thyroid carcinoma. J Nucl Med 1997; 38:1362–1368.
203. Cholewinski SP, Yoo KS, Klieger PS, et al. Absence of thyroid stunning after
diagnostic whole-body scanning with 185 MBq 131I. J Nucl Med 2000; 41(7):1198–
1202.
204. Bal CS, Kumar A, Pant GS. Radioiodine dose for remnant ablation in differentiated
thyroid carcinoma: A randomized clinical trial in 509 patients. J Clin Endocrinol
Metab 2004; 89(4):1666–1673.
205. Brierley J, Maxon HR. Radioiodine and external radiation therapy. In: Fagin JA,
ed. Thyroid Cancer. Boston/Dordrecht, London: Kluwer Academic Publishers,
1998:285–317.
206. Bal C, Padhy AK, Jana S, et al. Prospective randomized clinical trial to evaluate the
optimal dose of 131 I for remnant ablation in patients with differentiated thyroid
carcinoma. Cancer 1996; 77(12):2574–2580.
207. Maenpaa HO, Heikkonen J, Vaalavirta L, et al. Low vs. high radioiodine activity to
ablate the thyroid after thyroidectomy for cancer: A randomized study. PLoS ONE
2008; 3(4):e1885.
292 Sipos and Mazzaferri

208. Hackshaw A, Harmer C, Mallick U, et al. 131I activity for remnant ablation in
patients with differentiated thyroid cancer: A systematic review. J Clin Endocrinol
Metab 2006.
209. Greenler DP, Klein HA. The scope of false-positive iodine-131 images for thyroid
carcinoma. Clin Nucl Med 1989; 14(2):111–117.
210. Carlisle MR, Lu C, McDougall IR. The interpretation of 131I scans in the evaluation
of thyroid cancer, with an emphasis on false positive findings. Nucl Med Commun
2003; 24(6):715–735.
211. Chung JK, Lee YJ, Jeong JM, et al. Clinical significance of hepatic visualization on
iodine-131 whole-body scan in patients with thyroid carcinoma. J Nucl Med 1997;
38(8):1191–1195.
212. Simpson WJ, Panzarella T, Carruthers JS, et al. Papillary and follicular thyroid
cancer: Impact of treatment in 1578 patients. Int J Radiat Oncol Biol Phys 1988;
14:1063–1075.
213. Samaan NA, Schultz PN, Haynie TP, et al. Pulmonary metastasis of differentiated
thyroid carcinoma: Treatment results in 101 patients. J Clin Endocrinol Metab 1985;
60:376–380.
214. Ruegemer JJ, Hay ID, Bergstralh EJ, et al. Distant metastases in differentiated thyroid
carcinoma: A multivariate analysis of prognostic variables. J Clin Endocrinol Metab
1988; 67:501–558.
215. Schlumberger M, Tubiana M, De Vathaire F, et al. Long-term results of treatment of
283 patients with lung and bone metastases from differentiated thyroid carcinoma. J
Clin Endocrinol Metab 1986; 63:960–967.
216. Mazzaferri EL. Carcinoma of follicular epithelium: Radioiodine and other treatment
outcomes. In: Braverman LE, Utiger RD, eds. The Thyroid: A Fundamental and
Clinical Text. Philadelphia, PA: Lippencott-Raven Pub, 1996:922–945.
217. Tuttle RM, Leboeuf R, Robbins RJ, et al. Empiric radioactive iodine dosing regimens
frequently exceed maximum tolerated activity levels in elderly patients with thyroid
cancer. J Nucl Med 2006; 47(10):1587–1591. Erratum in: J Nucl Med. 2007; 48(1):7.
218. Kulkarni K, Nostrand DV, Atkins F, et al. The relative frequency in which empiric
dosages of radioiodine would potentially overtreat or undertreat patients who have
metastatic well-differentiated thyroid cancer. Thyroid 2006; 16(10):1019–1023.
219. de Keizer B, Hoekstra A, Konijnenberg MW, et al. Bone marrow dosimetry and safety
of high 131I activities given after recombinant human thyroid-stimulating hormone
to treat metastatic differentiated thyroid cancer. J Nucl Med 2004; 45(9):1549–1554.
220. Hanscheid H, Lassmann M, Luster M, et al. Iodine biokinetics and dosimetry in
radioiodine therapy of thyroid cancer: Procedures and results of a prospective inter-
national controlled study of ablation after rhTSH or hormone withdrawal. J Nucl
Med 2006; 47(4):648–654.
221. Pons F, Carrio I, Estorch M, et al. Lithium as an adjuvant of iodine-131 uptake when
treating patients with well-differentiated thyroid carcinoma. Clin Nucl Med 1987;
8:644–647.
222. Koong SS, Reynolds JC, Movius EG, et al. Lithium as a potential adjuvant to 131 I
therapy of metastatic, well differentiated thyroid carcinoma. J Clin Endocrinol Metab
1999; 84:912–916.
223. Datz FL. Cerebral edema following iodine-131 therapy for thyroid carcinoma
metastatic to the brain. J Nucl Med 1986; 27:637–640.
Differentiated Thyroid Carcinoma 293

224. Luster M, Lippi F, Jarzab B, et al. rhTSH-aided radioiodine ablation and treatment
of differentiated thyroid carcinoma: A comprehensive review. Endocr Relat Cancer
2005; 12(1):49–64.
225. DiRusso G, Kern KA. Comparative analysis of complications from I-131 radioabla-
tion for well-differentiated thyroid cancer. Surgery 1994; 116:1024–1030.
226. Allweiss P, Braunstein GD, Katz A, et al. Sialadenitis following I-131 therapy for
thyroid carcinoma: Concise communication. J Nucl Med 1984; 25:755–758.
227. Alexander C, Bader JB, Schaefer A, et al. Intermediate and long-term side effects of
high-dose radioiodine therapy for thyroid carcinoma. J Nucl Med 1998; 39:1551–
1554.
228. Brown AP, Greening WP, McCready VR, et al. Radioiodine treatment of metastatic
thyroid carcinoma: The Royal Marsden Hospital experience. Br J Radiol 1984;
57:323–327.
229. Maxon H III, Smith HS. Radioiodine-131 in the diagnosis and treatment of metastatic
well differentiated thyroid cancer. Endocrinol Metabol Clin North Am 1990; 19:685–
718.
230. Hall P, Holm L-E. Cancer in iodine-131 exposed patients. J Endocrinol Invest 1995;
18:147–149.
231. Brown AP, Chen J, Hitchcock YJ, et al. The risk of second primary malignancies up
to three decades after the treatment of differentiated thyroid cancer. J Clin Endocrinol
Metab 2008; 93(2):504–515.
232. Van Nostrand D, Neutze J, Atkins F. Side effects of “rational dose” iodine-131
therapy for metastatic well-differentiated thyroid carcinoma. J Nucl Med 1986;
27:1519–1527.
233. De Vathaire F, Schlumberger M, Delisle MJ, et al. Leukaemias and cancers
following iodine-131 administration for thyroid cancer. Br J Cancer 1997; 75:734–
739.
234. Menzel C, Grünwald F, Schomburg A, et al. “High-dose” radioiodine therapy in
advanced differentiated thyroid carcinoma. J Nucl Med 1996; 37:1496–1503.
235. Kloos RT, Duvuuri V, Jhiang SM, et al. Nasolacrimal drainage system obstruction
from radioactive iodine therapy for thyroid carcinoma. J Clin Endocrinol Metab
2002; 87(12):5817–5820.
236. Schlumberger M, De Vathaire F, Ceccarelli C, et al. Outcome of pregnancy in women
with thyroid carcinoma. J Endocrinol Invest 1995; 18:150–151.
237. Schlumberger M, De Vathaire F, Ceccarelli C, et al. Exposure to radioactive iodine-
131 for scintigraphy or therapy does not preclude pregnancy in thyroid cancer
patients. J Nucl Med 1996; 37:606–612.
238. Pacini F, Gasperi M, Fugazzola L, et al. Testicular function in patients with differen-
tiated thyroid carcinoma treated with radioiodine. J Nucl Med 1994; 35:1418–1422.
239. McGriff NJ, Csako G, Gourgiotis L, et al. Effects of thyroid hormone suppression
therapy on adverse clinical outcomes in thyroid cancer. Ann Med 2002; 34(7–8):554–
564.
240. Jonklaas J, Sarlis NJ, Litofsky D, et al. Outcomes of patients with differentiated
thyroid carcinoma following initial therapy. Thyroid 2006; 16(12):1229–1242.
241. Burmeister LA, Goumaz MO, Mariash CN, et al. Levothyroxine dose requirements
for thyrotropin suppression in the treatment of differentiated thyroid cancer. J Clin
Endocrinol Metab 1992; 75(2):344–350.
294 Sipos and Mazzaferri

242. Bartalena L, Martino E, Pacchiarotti A, et al. Factors affecting suppression of


endogenous thyrotropin secretion by thyroxine treatment: Retrospective analysis
in athyreotic and goitrous patients. J Clin Endocrinol Metab 1987; 64:849–855.
243. Pujol P, Daures JP, Nsakala N, et al. Degree of thyrotropin suppression as a prognostic
determinant in differentiated thyroid cancer. J Clin Endocrinol Metab 1996; 81:4318–
4323.
244. Hovens GC, Stokkel MP, Kievit J, et al. Associations of serum thyrotropin concen-
trations with recurrence and death in differentiated thyroid cancer. J Clin Endocrinol
Metab 2007; 92(7):2610–2615.
245. Sawin CT, Geller A, Wolf PA, et al. Low serum thyrotropin concentrations as a risk
factor for atrial fibrillation in older persons. N Engl J Med 1994; 331(19):1249–
1252.
246. Biondi B, Fazio S, Carella C, et al. Cardiac effects of long term thyrotropin-
suppressive therapy with levothyroxine. J Clin Endocrinol Metab 1993; 77:
334–338.
247. Toivonen J, Tahtela R, Laitinen K, et al. Markers of bone turnover in patients with
differentiated thyroid cancer with and following withdrawal of thyroxine suppressive
therapy. Eur J Endocrinol 1998; 138(6):667–673.
248. Uzzan B, Campos J, Cucherat M, et al. Effects on bone mass of long term treatment
with thyroid hormones: A meta-analysis. J Clin Endocrinol Metab 1996; 81:4278–
4289.
249. Bauer DC, Ettinger B, Nevitt MC, et al.; for the Study of Osteoporotic Fractures
Research Group. Risk for Fracture in Women with Low Serum Levels of Thyroid-
Stimulating Hormone. Ann Intern Med 2001; 134:561–568.
250. Marcocci C, Golia F, Vignali E, et al. Skeletal integrity in men chronically treated
with suppressive doses of L-thyroxine. J Bone Miner Res 1997; 12:72–77.
251. Chiu AC, Delpassand ES, Sherman SI. Prognosis and treatment of brain metastases
in thyroid carcinoma. J Clin Endocrinol Metab 1997; 82:3637–3642.
252. Tuttle RM, Leboeuf R. Investigational therapies for metastatic thyroid carcinoma. J
Natl Compr Canc Netw 2007; 5(6):641–646.
253. Gupta-Abramson V, Troxel AB, Nellore A, et al. Phase II trial of sorafenib in
advanced thyroid cancer. J Clin Oncol 2008. [Epub ahead of print]
254. Sherman SI, Wirth LJ, Droz JP, et al. Motesanib diphosphate in progressive differ-
entiated thyroid cancer. N Engl J Med 2008; 359(1):31–42.
255. Links TP, Van Tol KM, Jager PL, et al. Life expectancy in differentiated thyroid
cancer: A novel approach to survival analysis. Endocr Relat Cancer 2005; 12(2):273–
280.
256. Mazzaferri EL, Robbins RJ, Spencer CA, et al. A consensus report of the role of
serum thyroglobulin as a monitoring method for low-risk patients with papillary
thyroid carcinoma. J Clin Endocrinol Metab 2003; 88(4):1433–1441.
257. Baudin E, Cao CD, Cailleux AF, et al. Positive predictive value of serum thy-
roglobulin levels, measured during the first year of follow-up after thyroid hormone
withdrawal, in thyroid cancer patients. J Clin Endocrinol Metab 2003; 88(3):1107–
1111.
258. Mazzaferri EL, Kloos RT. Is diagnostic iodine-131 scanning with recombinant
human TSH (rhTSH) useful in the follow-up of differentiated thyroid cancer after
thyroid ablation? J Clin Endocrinol Metab 2002; 87:1490–1498.
Differentiated Thyroid Carcinoma 295

259. Kloos RT, Mazzaferri EL. A single recombinant human thyrotrophin-stimulated


serum thyroglobulin measurement predicts differentiated thyroid carcinoma metas-
tases three to five years later. J Clin Endocrinol Metab 2005; 90(9):5047–5057.
260. Eustatia-Rutten CF, Smit JW, Romijn JA, et al. Diagnostic value of serum thyroglob-
ulin measurements in the follow-up of differentiated thyroid carcinoma, a structured
meta-analysis. Clin Endocrinol (Oxf) 2004; 61(1):61–74.
261. Pacini F, Molinaro E, Castagna MG, et al. Recombinant human thyrotropin-
stimulated serum thyroglobulin combined with neck ultrasonography has the highest
sensitivity in monitoring differentiated thyroid carcinoma. J Clin Endocrinol Metab
2003; 88(8):3668–3673.
262. Spencer CA. Recoveries cannot be used to authenticate thyroglobulin (Tg) measure-
ments when sera contain Tg autoantibodies. Clin Chem 1996; 42:661–663.
263. Chiovato L, Latrofa F, Braverman LE, et al. Disappearance of humoral thyroid
autoimmunity after complete removal of thyroid antigens. Ann Intern Med 2003;
139(5 Pt 1):346–351.
264. Feine U, Lietzenmayer R, Hanke JP, et al. Fluorine-18-FDG and iodine-131-iodide
uptake in thyroid cancer. J Nucl Med 1996; 37:1468–1472.
265. Robbins RJ, Wan Q, Grewal RK, et al. Real-time prognosis for metastatic thyroid car-
cinoma based on 2-[18 F]fluoro-2-deoxy-D-glucose-positron emission tomography
scanning. J Clin Endocrinol Metab 2006; 91(2):498–505.
266. Kloos RT. Approach to the patient with a positive serum thyroglobulin and a neg-
ative radioiodine scan after initial therapy for differentiated thyroid cancer. J Clin
Endocrinol Metab 2008; 93(5):1519–1525.
267. Bakheet SMB, Powe J. Fluorine-18-fluorodeoxyglucose uptake in rheumatoid
arthritis-associated lung disease in a patient with thyroid cancer. J Nucl Med 1998;
39:234–236.
268. Schlumberger M, Challeton C, De Vathaire F, et al. Radioactive iodine treatment
and external radiotherapy for lung and bone metastases from thyroid carcinoma. J
Nucl Med 1996; 37:598–605.
269. Casara D, Rubello D, Saladini G, et al. Different features of pulmonary metastases
in differentiated thyroid cancer: Natural history and multivariate statistical analysis
of prognostic variables. J Nucl Med 1993; 34:1626–1631.
270. Schlumberger M, Arcangioli O, Piekarski JD, et al. Detection and treatment of lung
metastases of differentiated thyroid carcinoma in patients with normal chest X-rays.
J Nucl Med 1988; 29:1790–1794.
271. Pineda JD, Lee T, Ain K, et al. Iodine-131 therapy for thyroid cancer patients with
elevated thyroglobulin and negative diagnostic scan. J Clin Endocrinol Metab 1995;
80:1488–1492.
272. Filie AC, Asa SL, Geisinger KR, et al. Utilization of ancillary studies in thyroid fine
needle aspirates: A synopsis of the National Cancer Institute Thyroid Fine Needle
Aspiration State of the Science Conference. Diagn Cytopathol 2008; 36(6):438–441.
273. Basolo F, Giannini R, Monaco C, et al. Potent Mitogenicity of the RET/PTC3
Oncogene Correlates with Its Prevalence in Tall-Cell Variant of Papillary Thyroid
Carcinoma. Am J Pathol 2002; 160(1):247–254.
7
Medullary Thyroid Carcinoma,
Anaplastic Thyroid Carcinoma, and
Thyroid Lymphoma

Jennifer A. Sipos
The Ohio State University, Columbus, Ohio, U.S.A.

Ernest L. Mazzaferri
University of Florida, Gainesville, Florida, U.S.A.

MEDULLARY THYROID CANCER


Prevalence and Demographics
Medullary thyroid carcinoma (MTC) accounts for approximately 4% of all thyroid
malignancies (1). Familial MTC occurs with equal frequency in both sexes, while
sporadic MTC has a female/male ratio of 1.5:1. Only about 10% to 20% of MTC
cases occur as familial tumors. The other 80% to 90% are sporadic and may occur
at any age, but usually are detected later in life than familial MTC. For instance, the
median age of patients with sporadic MTC seen at the Mayo Clinic was 51 years,
compared with 21 years for those with familial tumors (2). Familial Non-MEN
Medullary Thyroid Carcinoma (FMTC) is usually detected later, around 40 to 50
years, as compared with an average age of 20 to 30 years for Multiple Endocrine
Neoplasia type 2A (MEN-2A) when detected by screening affected kindreds with
calcitonin tests. The diagnosis is made even earlier now that genetic testing is
available.

297
298 Sipos and Mazzaferri

Classification
MTC was first recognized in 1959 as a pleomorphic neoplasm with amyloid
stroma. A few years later, it became apparent that the tumor arises from the
calcitonin-secreting C cells of the thyroid. Approximately 70% to 80% are sporadic
tumors and 20% to 30% are familial tumors that are transmitted as an autosomal
dominant trait, which is often associated with other endocrine neoplasms. The
genes that are responsible for the familial forms of MTC map to the pericentromeric
region of chromosome 10 and typically cause three familial syndromes, described
below (3).

Multiple Endocrine Neoplasia Type 2 Syndromes


MEN Type 2A
This is the most common Multiple endocrine neoplasia type 2 (MEN-2) syndrome,
which comprises bilateral MTC that occurs in more than 90% of the gene carriers,
and bilateral pheochromocytoma and hyperparathyroidism approximately 40%
and 25% of the carriers, respectively (4). The syndrome is also associated with
cutaneous lichen amyloidosis, a pigmented pruritic lesion that typically is found
on the back (5). Also, MEN-2A may occur with congenital absence of enteric
innervation resulting in Hirschsprung’s disease (6).

MEN Type 2B
This is the most distinctive and aggressive but least common familial MEN-
2 syndrome (7). Patients have bilateral MTC and pheochromocytomas but not
hyperparathyroidism. They have a distinct phenotype with a marfanoid habitus
with a decreased upper/lower segment body ratio, long limbs, hyperextensible
joints, scoliosis, and anterior chest deformities; however, this is not associated
with the ectopic lens or cardiovascular abnormalities characteristic of Marfan’s
syndrome. The other unique phenotypic characteristic is ganglioneuromatosis
involving the salivary glands, pancreas, intestine, gallbladder, upper respiratory
tract, and urinary bladder. Patients also have visible ganglioneuromas on the
tongue, lips, and eyelids, producing a lumpy patulous appearance of the lips.
Alimentary ganglioneuromas may be associated with constipation, diarrhea, and
megacolon.

Familial Non-MEN Medullary Thyroid Carcinoma


This syndrome consists of bilateral MTC with no other endocrine tumors or
somatic abnormalities. It is transmitted as an autosomal dominant trait and is
the least common form of MTC, manifest at a later age than the other familial
syndromes (4). There is, however, considerable overlap in the presentation of this
syndrome and MEN-2A (8).
Medullary Thyroid Carcinoma 299

Genetic Alterations in MEN-2 and FMTC Syndromes


Point mutations of the Rearranged during transfection (RET) proto-oncogene
occur in germ-line and tumor DNA of unrelated patients from kindreds with MEN-
2A, MEN type 2B (MEN-2B), and FMTC (9). Several different and independent
point mutations in the genomic sequence of the RET proto-oncogene have been
identified, all involving codons for cysteine residues, which provides an important
direct means of identifying affected MEN-2 and FMTC kindreds. Nonetheless, the
normal function of RET is not yet known and its role in the development of these
inherited syndromes remains unclear. There is, however, a relationship between
specific RET proto-oncogene mutations and the MEN-2 phenotypes (9).
MTC Tumor Location
MTC in the familial syndromes is generally bilateral and multicentric, as opposed
to sporadic MTC, which is usually unilateral. The C cells, which are normally
located in the upper and middle thirds of the lateral thyroid lobes, initially undergo
hyperplasia in the familial syndromes before developing into MTC. A European
study (10) of 207 patients with familial MTC from 145 families found significant
age-related progression from C-cell hyperplasia (CCH) to MTC and ultimately to
nodal metastasis in patients whose RET mutations were grouped according to the
affected extracellular and intracellular-domain codons and in those with the codon
634 genotype. No lymph-node metastases were found in patients younger than
14 years. The authors of this study concluded that the codon-specific differences
in the age at presentation of MTC and the familial rates of concomitant adrenal
and parathyroid involvement suggest that the risk of progression was based on
the transforming potential of each individual RET mutation (10). The earliest
reported ages for the onset of MEN-2 was the first year of life for families with a
918 mutation (MEN-2B) to 20 years of age or older with RET codon 791 and 768
mutations (10).
Parathyroid Disease
Hyperparathyroidism develops in one-third to half of the patients with the
MEN-2A syndrome, most of whom (85%) develop parathyroid hyperplasia
(11), which almost never occurs in MEN-2B (7). MEN-2A patients seldom
present with symptoms of hypercalcemia but often form renal stones. When
MTC develops in MEN-2A patients, parathyroid hyperplasia is often discovered
during thyroidectomy even when there is no clinical or biochemical evidence of
hyperparathyroidism (7).
Pheochromocytoma
Adrenal medullary disease occurs in both MEN-2A and MEN-2B syndromes. Its
manifestations range from diffuse or nodular adrenal hyperplasia to large, bilateral,
multilobular pheochromocytomas. These abnormalities are typically bilateral and
occur in approximately 40% of affected family members, although this ranges
300 Sipos and Mazzaferri

widely occurring in 6% to 100% in different kindreds (11). Recent studies suggest


a codon-specific, age-related development of pheochromocytoma in RET carriers
(12). A study of (11) of 206 RET carriers with a mean observation period of
27 years found that pheochromocytomas developed in 28% of the carriers with
mutations in codon 918, 29% of carriers with mutations in 634, 14% of carriers
with mutation in codon 618, 13% of carriers with mutations in codon 620, and
13% of carriers with mutations in codon 791. The earliest age of manifestation for
each genotype was 22, 18, 29, 22, and 39 years. Contralateral pheochromocytomas
developed after four years in carriers of codon 618 and after 5.2 years in carriers
with mutations in codon 634. No pheochromocytomas were identified in carriers
of mutations in codons 609, 611, 630, 768, 790, 804, and 891. The authors of
this study recommended annual screening for pheochromocytoma from age 10 in
carriers of RET mutations in codons 918, 634, and 630, and from age 20 in the
remainder (11).
Pheochromocytoma symptoms are typically subtler than those encountered
with sporadic pheochromocytoma (2). The diagnosis is established by demonstrat-
ing high urinary or plasma catecholamine, or metanephrine levels, although total
urinary catecholamines may be normal, and only the epinephrine/norepinephrine
ratio may be increased, particularly with medullary hyperplasia (2).
Genetic Alterations in Sporadic MTC
Approximately 50% of sporadic MTC tumors harbor a somatic mutation at codon
918 of the RET proto-oncogene without showing a similar mutation in other tissues
(13). One study (14) detected a RET somatic mutation at codon 918 in fine-needle
aspiration (FNA) cytology specimens obtained from both a thyroid nodule and
two enlarged neck lymph nodes but not in the peripheral blood, thus establishing
a diagnosis of sporadic MTC before surgery. This is important because exclud-
ing MEN-2 preoperatively permits immediate thyroidectomy without search for
pheochromocytoma (14).
The Relationship Between Somatic RET Mutation and Prognosis
Although an association has been described between somatic mutations and a poor
prognosis, this has until recently been controversial, mainly because the studies
were small and had inadequate follow-up. However, a recent study of 100 patients
with sporadic MTC who had a mean follow-up of 10.2 years found a strong
relationship between somatic RET mutations and outcome (15). After analyzing
RET exons, 10 to 11 and 13 to 16, a somatic RET mutation was found in 43% of
the patients. The most frequent mutation (79%) was M918 T. Patients with a RET
mutation were more likely to have larger tumors (p = 0.03) with more lymph node
and distant metastases (p ⬍ 0.0001 and p = 0.02, respectively). Thus, there was
a significant correlation between the RET somatic mutation and more advanced
disease stage at the time of diagnosis (p = 0.004). Multivariate analysis found
that only advanced disease stage at the time of diagnosis and the presence of a
RET mutation were independent factors predicting poor outcome (p ⬍ 0.0001 and
Medullary Thyroid Carcinoma 301

Figure 1 Anaplastic thyroid carcinoma (ATC) and medullary thyroid carcinoma (MTC):
ATC with tumor infiltrating the entire gland and invading the thyroid capsule at 4 o’clock
(top left); histology of ATC showing large, bizarre-appearing cells (top right); MTC with
spindle cells and dense stroma that stains positive for amyloid (lower right); spindle cell
tumor that was a MTC, which can mimic ATC (lower right).

p = 0.01, respectively). Survival curves showed a significantly lower rate of


survival in patients with RET mutations (p = 0.006).

Pathology
Primary Tumor
Ranging from large bulky tumors to microscopic lesions, sporadic and familial
tumors are nonetheless histologically similar. However, the widest spectrum is
encountered in familial tumors, which range from isolated microscopic C cells
that are hypertrophied to large, bilateral, multicentric tumors that are usually in the
superior portions of the thyroid lobes. Sporadic MTC is usually unilateral. MTC
is typically composed of fusiform or polygonal cells surrounded by irregular
masses of amyloid and abundant collagen (Fig. 1). About half the tumors have
calcifications, which occasionally appear as trabecular bone formation. Calcitonin
can usually be demonstrated in the tumors by immunohistochemical studies.
Metastases
Cervical lymph node metastases occur early in the disease and can be seen with
primary lesions as small as several millimeters. Tumors larger than 1.5 cm in
diameter are more likely to metastasize to distant sites, especially to lung, bone,
and liver but also to the CNS. Metastatic deposits usually contain calcitonin and
stain for amyloid.
302 Sipos and Mazzaferri

Tumor Calcitonin
Immunohistochemical staining for calcitonin serves not only to identify MTC but
also, by the staining pattern, to differentiate virulent from less aggressive tumors.
In one study, patients with primary tumors that showed intense homogeneous cal-
citonin staining were all clinically well on follow-up examination, while those with
tumors that showed patchy localization of calcitonin either developed metastatic
disease or died of cancer within six months to five years of initial surgery (16).
Still, the presence of RET oncogene in sporadic tumors appears to be the most
helpful prognostic marker (15).

Diagnosis
Clinical Features
Patients with sporadic disease or unrecognized familial MTC usually present with
one or more painless thyroid nodules in an otherwise normal gland, although
some experience pain, dysphagia, and hoarseness. The presenting feature may be
enlarged cervical lymph nodes or occasionally distant metastases, most commonly
to the lung, followed in frequency by metastases to the liver, bone (osteolytic or
osteoblastic lesions), and brain. Cervical lymph node metastases are present at the
time of diagnosis in about half the patients with sporadic MTC. The thyroid nodule
is usually solid on ultrasonography and malignant on FNA, and may be cold or
warm on 123 I imaging. Radiographs may show dense, irregular calcifications of the
primary tumor and cervical nodes, and mediastinal widening due to metastases.
The abnormal phenotype of MEN-2B may lead to the clinical diagnosis. Rarely,
a paraneoplastic syndrome such as diarrhea or flushing may be the presenting
manifestation.

Hormonal Features
In addition to calcitonin, MTC may synthesize calcitonin gene-related peptide, L-
dopa decarboxylase, serotonin, prostaglandins, adrenocorticotropin (ACTH), his-
taminase, carcinoembryonic antigen (CEA), nerve growth factor, and substance P
(2). Elevated serum levels of calcitonin, histaminase, L-dopa decarboxylase, and
CEA are frequently found in MTC patients. Approximately 10% of patients have
episodes of flushing, often induced by alcohol ingestion, calcium infusion, and pen-
tagastrin injection; these episodes may be due to tumor release of prostaglandins
and serotonin (2). Approximately 30% of patients also experience a secretory
diarrhea that may be related to the high circulating levels of calcitonin, though this
is typically seen in those with advanced tumors. Because of the indolent course of
MTC in many patients, ectopic ACTH produced by the tumor may cause typically
Cushing’s syndrome, which is not seen in most other tumors that produce ACTH
such as lung cancer (2).
Medullary Thyroid Carcinoma 303

Calcitonin and Tumor Mass


Elevated basal plasma calcitonin levels are almost always found once the MTC
becomes palpable, and correlate directly with tumor mass (2). As a result, basal
calcitonin often is not elevated with small tumors and is almost invariably normal
in those with CCH; however, after stimulation with pentagastrin and/or calcium,
serum calcitonin levels increase to abnormally high levels.

Calcitonin in the Evaluation of Thyroid Nodules


Some clinicians advise routine measurement of basal serum calcitonin in those
with nodular thyroid disease undergoing FNA biopsy, finding that it allows early
preoperative diagnosis of subclinical MTC and improves the outcome of surgery.
Most of the studies that advocate this approach are from Europe where the inci-
dence of multinodular goiter is high and pentagastrin is available for clinical use
(17). This has not been the practice among U.S. thyroidologists, mainly because it
is viewed as not being cost-effective because baseline serum calcitonin levels are
often falsely elevated in patients without MTC.
Still, there is incontrovertible evidence that the survival rate of patients with
MTC is substantially better than usual when the tumor is limited to the thyroid.
Elisei et al. (17) found that calcitonin screening in a cohort of 10,864 patients
with thyroid nodular disease had a higher diagnostic sensitivity and specificity
compared with FNA biopsy. Calcitonin screening resulted in a diagnosis of MTC
at an earlier stage compared with those in whom the test was not performed
(p = 0.004). Moreover, normalization of serum calcitonin levels (undetectable)
after surgery was more frequently observed in patients who had calcitonin screen-
ing before surgery. Complete remission was observed in 59% of the group that had
calcitonin testing compared with 2.7% of the group without calcitonin testing 9
(p = 0.0001). This study confirms that MTC is a relatively frequent finding among
patients with thyroid nodules—nearly 1 in 250 patient—that benefits patients with
MTC. The American Thyroid Association (ATA) guidelines for the management
of thyroid nodules and differentiated thyroid carcinoma suggest that if the test is
done, patients should undergo surgery if in the serum calcitonin level is higher
than 100 pg/mL (18).

Stimulation Tests
Some patients with small MTC tumors have normal basal calcitonin levels, which
rise to high levels with stimulation. The combination of intravenous pentagastrin
and calcium is a more potent stimulus for calcitonin release than either agent
alone. Unfortunately, pentagastrin is no longer available in the United States. In
the calcium stimulation test, elemental calcium (2 mg/kg) in 50 mL of 0.9% saline
is given intravenously over one minute and plasma is collected for calcitonin
determination every 10 minutes for 30 minutes. Patients with CCH and MTC
generally have plasma calcitonin levels that rise fivefold above baseline. Basal
and stimulated levels are higher in men than in women and decline with age. Of
304 Sipos and Mazzaferri

course, the diagnosis of inherited MTC is now possible with genetic screening long
before the thyroid tumor is clinically manifest or calcitonin levels are elevated,
rendering stimulation testing obsolete in most situations.
Omeprazole Stimulation Test
Omeprazole stimulates intrinsic gastrin release, which may stimulate calcitonin
and be useful in the diagnosis and follow-up of MTC, but it is less potent than
pentagastrin for this indication (17).
Differential Diagnosis of Hypercalcitoninemia
An elevated serum calcitonin level is not absolutely diagnostic of MTC because it
occurs in other conditions (2) such as patients on chronic hemodialysis and other
neuroendocrine tumors. When the differential diagnosis of a high plasma calcitonin
value is between MTC and another malignancy, a higher calcitonin value and a
palpable thyroid tumor usually identify patients with MTC. In addition, calcitonin
secretion by other cancers is poorly stimulated by pentagastrin.

Factors Influencing Prognosis


Age and Tumor Stage
MTC is more aggressive than papillary and follicular carcinoma, having a cancer-
specific mortality rate of approximately 25% at 10 years (1). However, pheochro-
mocytoma causes a number of deaths among those with familial MTC (19). The
mortality rate is substantially worse with sporadic tumors or when metastases
are found at the time of diagnosis, or with the MEN-2B phenotype, and among
patients older than 50 years at the time of diagnosis. Children with familial disease
operated on during the first decade of life generally have no evidence of resid-
ual disease postoperatively (2). However, persistent or recurrent disease occurs
in about one-third of patients operated on in the second decade and it gradually
increases in frequency until the seventh decade when about two-thirds of patients
have persistent disease after surgery (2). This is largely due to the patient’s age,
tumor stage at the time of surgery, and the genetic makeup (RET mutation) of the
tumor, which are the three most powerful prognostic factors predicting outcome
(20,21).
Familial Medullary Thyroid Carcinoma
Prognosis is best with FMTC and MEN-2 A. Early detection and treatment have
a profound impact on the clinical course of MTC. The 10-year survival rates
are nearly similar to those in unaffected subjects when nodal metastases are not
present, but the rates fall to approximately 45% when nodal metastases are present
(2). Before 1970, MTC was usually diagnosed in the fifth or sixth decade of life.
For the next 20 years, biologic markers (pentagastrin- or calcium-stimulated calci-
tonin) assumed a preeminent role in diagnosis. With periodic calcitonin screening,
patients from MEN kindreds were diagnosed at a much earlier age, usually in
Medullary Thyroid Carcinoma 305

the second decade or earlier, when they had CCH or microscopic carcinoma con-
fined to the thyroid (22). Nonetheless, the results of long-term studies using this
approach are somewhat disappointing. One large study comparing calcitonin with
genetic testing found 14 young gene carriers (18% of the 80 known gene car-
riers in the study) who had normal plasma calcitonin tests, eight of whom had
undergone thyroidectomy at the time of the report and were found to have foci
of MTC (23). Another study (24) reported long-term follow-up of 22 children
in whom pentagastrin-stimulated calcitonin screening studies were routinely per-
formed once a year and thyroidectomy was recommended for any elevation in
serum calcitonin levels. MTC had already developed in 17 children (77%), and
only five of them had CCH. Of the 17 with MTC, 13 had macroscopic tumors,
and recurrent disease developed in four children (24%) (24).

Therapy
Initial Surgery
Surgery offers the only chance for cure and should be performed as soon as
the disease is detected (20,21). However, pheochromocytoma must be rigorously
searched for and excised before thyroidectomy is performed. The treatment of
MTC confined to the neck is total thyroidectomy because the disease is often
bilateral, even with a negative family history, because patients are often unsus-
pected relatives of affected MEN-2 kindred (20,21). All patients with palpable
MTC or clinically occult disease that is visible on cut section of the thyroid
should undergo routine dissection of lymph nodes in the central neck compart-
ment because nodal metastases occur early and adversely influence survival. The
lateral cervical lymph nodes should be dissected when they contain tumor, but
radical neck dissection is not recommended unless the jugular vein, accessory
nerve, or sternocleidomastoid muscle in invaded by tumor (25).
Residual or Recurrent MTC
Persistent or recurrent disease, which is ordinarily manifest by elevated calcitonin
levels postoperatively, occurs commonly following primary treatment of the tumor.
Nonetheless, survival is often good. For example, a French study reported that 57%
of 899 patients were not cured (57%), but that survival was 80% at 5 years and
70% at 10 years. Thus, reoperation in appropriately selected patients is advisable.
It is the only treatment that consistently and reliably reduces calcitonin levels and
often results in excellent local disease control. There is no curative therapy for
widely metastatic disease, but preoperative neck microdissections may normal-
ize calcitonin levels when metastatic MTC is confined to regional lymph nodes.
Improved results have been reported in recent years with the surgical manage-
ment of recurrent MTC, mainly through better preoperative selection of patients
and the use of routine laparoscopic liver examination preoperatively, which iden-
tifies hepatic metastases in MTC patients with normal computed tomography
(CT) scan and magnetic resonance imaging (MRI) (26). Although patients with
306 Sipos and Mazzaferri

widely metastatic MTC often live for years, many develop symptoms from dis-
ease progression. Despite the presence of widespread incurable tumor, patients
with metastatic MTC causing significant symptoms or physical compromise may
respond to palliative, preoperative resection (27) or tumor embolization (28). Judi-
cious, palliative, preoperative resection of discrete symptomatic lesions provides
substantial long-term relief of symptoms with minimal operative mortality and
morbidity (27).

Inoperable Disease
Patients with inoperable disease are often given palliative treatment with external
radiation for localized disease or with doxorubicin or other chemotherapy combi-
nations for widespread, life-threatening disease, which is of limited benefit. Radi-
olabeled metaiodobenzyguanidine (MIBG) or somatostatin analogues and 131 I or
yttrium-90 (99 Y)-labeled humanized anti-CEA monoclonal antibodies are poten-
tially useful toward tumor debulking. Nonradioactive somatostatin analogues may
importantly control the paraneoplastic symptoms of flushing or diarrhea, but they
do not improve the natural course of advanced stages of MTC (29). Embolization
is especially effective for liver metastases, often relieving symptoms of flushing
and diarrhea that are characteristic of hepatic metastases (28,30). Recent trials
(31,32) of multikinase inhibitors have not proven effective at inducing a remission
in these tumors, however, several studies involving different agents are currently
underway and may reveal treatments that offer stabilization of disease.

Follow-up
The success of surgery is assessed postoperatively by measuring plasma calcitonin
levels. Although it may require up to six months for calcitonin to normalize,
undetectable basal and stimulated calcitonin levels usually indicate a cure. Persis-
tent basal calcitonin elevations are often seen after surgery in patients who remain
well for many years, particularly those from MEN-2A kindred who should be
followed without further aggressive therapy. If postoperative plasma calcitonins
are extremely high or the patient has flushing and/or diarrhea, imaging studies
must be performed. There are conflicting data concerning which imaging studies
have the highest sensitivity in this situation. However, a prospective study of 55
consecutive patients with MTC who had elevated serum calcitonin levels found
that the most efficient imaging work-up for detecting MTC tumor was as follows:
ultrasonography for the neck, CT scan for the chest, MRI for the liver, bone
scintigraphy, and axial skeleton MRI for bone metastases (33). Unlike papillary
and follicular thyroid carcinoma, FDG-PET scanning was less sensitive and of
low prognostic value for MTC. Another prospective study of the same cohort
of 55 patients with MTC found that calcitonin and CEA levels were correlated
with tumor burden and their doubling times were strongly related to disease
progression (34). When the CEA and calcitonin doubling times were shorter than
Medullary Thyroid Carcinoma 307

25 months, 94% of the patients had progressive disease while most (86%) of the
patients with doubling times longer than 25 months had stable disease.

Family Screening
Genetic Screening
Affected patients can be identified at birth with proper genetic RET testing of
codons 10, 11, 13 to 16, which hopefully will result in prevention of disease
when the patient is operated on early in life. To do this, all patients with presumed
sporadic MTC should undergo genetic screening, although only approximately 3%
of such cases turn out to have familial disease. Nonetheless, this is an extremely
important concept given the possibility of overlooking potentially lethal occult
pheochromocytoma and the implications for failing to identify affected kindred.
All first-degree relatives of any patient who tests positive for a MEN-2 or FMTC
mutation should be screened. One study found that without genetic testing, even
when the mean age at the time of thyroidectomy was about 10 years, a significant
number of patients (21%) with MEN-2A or -2B developed persistent or recurrent
MTC over a follow-up of about a decade (35). However, the sensitivity of genetic
screening for MEN-2A offered by diagnostic laboratories that limit RET analysis
to exons 11, 13–16 results in a more complete and accurate analysis with a
sensitivity of nearly 95% (35). It is recommended that clinicians confirm the
comprehensiveness of a laboratory’s genetic screening approach for MEN-2A to
ensure thoroughness of sample analysis.

Prophylactic Total Thyroidectomy


In 1993, when mutations of the RET proto-oncogene were identified in hereditary
MTC, surgeons began to operate on patients prophylactically before the disease
was clinically manifest.
Nonetheless, microscopic or grossly evident MTC was often present in the
excised thyroid glands but almost none were metastatic to regional lymph nodes
at the time of surgery (36). The debate now is focused on the optimal age for
operating on children and the extent of lymph node resection. There are differing
views on this issue from three authoritative sources on this issue.

The Gubbio Conference Recommendation


This consensus opinion comes from the Seventh International Workshop on MEN
held in Gubbio, Italy in June 1999 (4). The report from this meeting represents
a consensus opinion from an international group of experts (4). Patients with
familial MTC are stratified into three groups.
Level 3 (highest risk group) is children with MEN-2B and/or RET codon
883, 918, or 922 mutation. The participants recommended that all of these children
should have thyroidectomy performed within the first six months and preferably
within the first month of life. This is based upon the fact that microscopic MTC
308 Sipos and Mazzaferri

within the first year of life in this group is common and metastases may occur
during the first year. They recommended a central neck lymph node dissection,
and that more extensive surgery may be appropriate if metastases are identified at
surgery.
Level 2 (high-risk group) is children with any RET codon 611, 618, 620,
or 634 mutation (MEN-2A). For these children the participants recommended
total thyroidectomy before the age of five years, including removal of the thyroid
posterior capsule. However, there was no consensus regarding the need for pro-
phylactic central lymph node dissection, which recommended by surgeons but not
by internists. This differing opinion is related to the high risk of hypothyroid and
laryngeal nerve damage by such aggressive surgery.
Level 1 (least high-risk group) is children with RET codon 609, 768, 790,
791, 804, and 891 mutations. Total thyroidectomy, including the posterior thyroid
capsule, was recommended for this group of children. However, there was no
consensus as what age the thyroidectomy should be done. This is based upon the
fact that the biologic behavior of MTC with the RET codon mutations indicated
is variable, but generally grows more slowly than tumors in the other two groups.
No recommendation was made for several other RET codon mutations.
The EUROMEN Study
This is a European multicenter study (10) of 207 patients from 145 families that
was conducted from July 1993 to February 2001. The patients who were 20 years
of age or younger, were asymptomatic and had undergone total thyroidectomy for
MTC after confirmation of the RET mutation. As noted above, the main findings
of the study were that there is a significant age-related progression from CCH
to MTC, and ultimately, to lymph-node metastases among patients with a RET
codon 634 genotype (MEN-2A and FMTC). As a result, the EUROMEN authors
provide initial age guidelines for prophylactic total thyroidectomy based upon the
earliest age at which MTC was identified for each of the known RET mutations for
each of the MEN-2 syndromes. It is important to note that the authors provide the
following caveat “. . . but the benefit of cure offered by prophylactic thyroidectomy
is offered by the potential overtreatment of some carriers of RET mutations” (10).

ANAPLASTIC THYROID CARCINOMA


Its extremely aggressive behavior and poor prognosis distinguish this exceptionally
virulent and invasive neoplasm. Occurring in an older population, Anaplastic
thyroid carcinoma (ATC) demonstrates a biologic behavior that is among the
worst encountered in humans.

Incidence and Demographics


The frequency of ATC relative to other thyroid cancers was approximately 5% to
10% in the past (37), but more recently it has been approximately 2% in the United
States (1). Part of this apparent reduction in incidence may be to improvements in
Medullary Thyroid Carcinoma 309

discrimination of this tumor from other poorly differentiated tumors of the thyroid
as a result of consensus conferences and improved histopathologic classification.
In a review of 475 patients with ATC from six large studies, the mean age was 65
years, with only a slight female predominance (37). It is almost never seen before
age 20.
The incidence of ATC is influenced by dietary iodine. In one study (38) the
frequency of ATC was threefold higher in an iodine-deficient area compared with
an iodine-rich area. This finding may account for differences in disease frequency
reported from around the world, which were especially evident several decades
ago.

Origin
Although some ATCs appear to arise de novo (38,39), there are many examples
in laboratory animals and humans of well differentiated tumors transforming
into ATCs, events that may evolve over many years (40). One large study (41)
demonstrated elements of differentiated thyroid carcinoma in almost 90% of ATC
specimens, which, in more than 20% of the cases, came from persons previously
diagnosed and treated for well-differentiated thyroid carcinoma.

Pathology
These tumors are composed wholly or in part of undifferentiated cells and tend to
behave according to their most aggressive tumor element.
Small-Cell Carcinoma
In the past, ATCs or undifferentiated thyroid carcinomas were divided into two
broad categories: spindle- or giant-cell carcinomas and small-cell carcinoma.
Now, however, almost all small-cell carcinomas are recognized both by elec-
tron microscopy and immunohistochemical study as primary thyroid lymphomas
(PTLs), MTCs, insular variants of thyroid carcinoma, or occasionally a small-cell
metastasis from lung cancer. The WHO classification of thyroid tumors recom-
mends against using the term.
Large-Cell Carcinoma
The three major cell types of ATC are spindle cells, giant cells, and squamoid
patterns. Subdivision based upon these cellular features is not clinically helpful
and is not recommended in the current WHO classification. For example, one large
study (42) found spindle cells in 53%, giant cells in 50%, and a squamoid element
in 19% of the tumors, but outcome was similar regardless of these features.
Gross Features
ATCs often involve both thyroid lobes and are typically invasive and poorly demar-
cated from surrounding neck tissues (Fig. 1). Extensive local invasion into the soft
tissues and other structures of the neck is common at the time of presentation. On
310 Sipos and Mazzaferri

gross examination, the tumors are gray-white, fibrous, calcified, or even ossified,
and they frequently show areas of necrosis.
Histologic Features
ATCs exhibit the three distinct morphologic patterns (Fig. 1)—spindle cell, giant
cell, and squamoid features—often with frequent mitotic figures. The tumors are
very pleomorphic and may resemble a fibrosarcoma or rhabdomyosarcoma, or they
may contain multinucleated giant cells that resemble osteoclasts and, very rarely,
malignant osteoid or cartilage may be seen simulating an osteogenic sarcoma or
chondrosarcoma.
Immunohistochemical Studies
The immunohistochemical proof of the follicular origin of ATC is its staining
with epithelial markers, the most useful of which is low-molecular-weight keratin
(cytokeratin), expressed in up to 80% of the cases (37). Other markers may be
detected, but some (43) show that 30% of ATCs express none of the tumor markers
examined. Less than one-third react with epithelial membrane antigen or CEA and
some (0–70%) stain for Tg. An occasional case of anaplastic MTC occurs (37).
p53 Suppressor Gene
Mutations in this gene occur in thyroid carcinoma almost exclusively in the poorly
differentiated tumors and thyroid cancer cell lines, suggesting that inactivation of
p53 may confer aggressive properties and further loss of differentiated function.
For example, one study found the frequency of p53 mutations to be zero in normal
thyroid, follicular adenomas, PTCs, and MTCs; however, it was approximately
9% in FTCs and 83% in ATCs (44).

Diagnosis
Clinical Presentation
History
A study (45) from MD Anderson found that almost two-thirds of patients presented
with a rapidly enlarging neck mass and about one-third had symptoms of tracheal
compression with invasion. Another study (46) from the Mayo Clinic also found
that about two-thirds had a rapidly enlarging neck mass, either with (37%) or
without (32%) a preexisting goiter, and about half had dyspnea at the time of
diagnosis. In this series, symptoms had been present for less than three months in
almost half the patients, while 20% had them for a year before diagnosis. Rapid
tumor enlargement often causes neck pain that may mimic subacute thyroiditis,
probably due to tumor necrosis and invasion of neck tissues (so-called “malignant
pseudothyroiditis”). Hoarseness and cough, with systemic symptoms such as fever
or weight loss, are often present. Most patients have normal thyroid function,
although thyrotoxicosis rarely occurs, probably from rapid thyroid tissue necrosis
releasing thyroid hormone.
Medullary Thyroid Carcinoma 311

Physical Examination
The tumor is typically hard, poorly circumscribed, and fixed to surrounding struc-
tures. In the Mayo Clinic series (46), 60% presented as a multinodular goiter and
38% presented as an apparently isolated thyroid nodule, while only 2% caused
diffuse thyroid enlargement. The tumors are characteristically quite large and may
be associated with palpable cervical lymph nodes. In the Mayo Clinic series (46),
approximately 80% were ⬎5 cm, half the patients had palpably enlarged cervical
lymph nodes, and one-third of the patients had vocal cord paralysis. Stridor that
occurs when the patient’s neck is extended portends serious airway obstruction
and should raise the question of elective tracheostomy to protect the airway. The
tumor may cause a superior mediastinal syndrome with venous distention and
edema of the face, arms, and neck.

Distant Metastases
Only about one-third of patients have distant metastases, mostly to the lung,
recognized at the time of diagnosis, but about half the patients eventually developed
them (45,46). The lung is the most common site of distant metastases, comprising
75% to 90% of all distant metastases reported in large series (37). They are
typically manifest as large isolated metastases that do not concentrate 131 I. Skeletal
metastases are the second most frequent but only account for approximately 5%
of the distant metastases at the time of diagnosis and approximately 15% of all
distant metastases that eventually develop (37). Uncommonly, patients with ATC
have distant metastases to other sites, such as the CNS (37).
The diagnosis of malignancy is usually quite evident with ATC because of
the ominous symptoms and physical findings suggesting an aggressive tumor. The
main diagnostic problems are differentiating ATC from less aggressive undiffer-
entiated thyroid tumors and delineating the extent of the neck disease.

Fine-Needle Aspiration
Most cases of ATC diagnosed by FNA should at least have open biopsy with
immunohistochemical staining to confirm the diagnosis. Although a diagnosis of
thyroid malignancy is almost always possible by FNA cytology, ATC may be diffi-
cult to distinguish from thyroid lymphoma, MTC, and other forms of poorly differ-
entiated thyroid carcinoma or cancers metastatic to the thyroid gland. Nonetheless,
the giant- and spindle-cell patterns of ATC usually predominate, sometimes with
multinucleated giant cells, suggesting the correct diagnosis. Although MTC often
shows a spindle-cell population, it can usually be correctly identified by FNA
cytology that may stain for amyloid or calcitonin (37). Non-Hodgkin’s thyroid
lymphoma can be identified by FNA, usually by its small cells, but open biopsy
and histochemical staining are often necessary to differentiate it from ATC, espe-
cially when the lymphoma is composed of large cells.
312 Sipos and Mazzaferri

Radionuclide Studies
Thyroid scanning with radioiodine usually discloses one or more cold nodules in
ATC, which is a nonspecific finding. However, gallium-67 (67 Ga) uptake is low
in well-differentiated thyroid carcinomas but high in anaplastic carcinomas and
malignant lymphomas and may detect distant metastases (47).

Thyroid Ultrasonography
The majority of differentiated carcinomas and all anaplastic tumors present as
hypoechoic masses by ultrasonography. Tumor clearly extending beyond the thy-
roid capsule or infiltrating adjacent tissues strongly suggests a malignant diagnosis.

Computed Tomography
CT scan is indispensable preoperatively since it may alter surgical planning, espe-
cially when the tumor extends into the thorax or invades the larynx, esophagus,
or other neck structures. ATC appears as a large mass of low attenuation accom-
panied by dense calcification in over half the patients in most of whom there is
also evidence of tumor necrosis (48). CT scan usually identifies tumor infiltrating
neck structures, including the carotid artery, internal jugular vein, larynx, trachea,
esophagus, mediastinum, and regional lymph nodes. CT scan with contrast iden-
tifies lymph nodes, which are often necrotic, more consistently than palpation
(48).

Natural History and Mortality Rates


ATC has a dismal prognosis. In 1961, Woolner et al. (49) reported from the Mayo
Clinic that 61% of ATC patients were dead within six months and 77% died
within a year of diagnosis. Almost 50 years later, a study from MD Anderson (45)
reported a mean survival of seven months. Only 8% (20 of 240 patients) from both
series survived longer than one year. Median survival in a later Mayo Clinic series
was only four months (37). Overall cause-specific mortality of 516 patients from
the SEER database was 69.3% at 6 months and 80.7% at 12 months (50). This is
why lymphomas and MTCs, which histologically resemble ATC but have a sub-
stantially better prognosis, must be carefully identified by immunohistochemical
studies.

Cause of Death
Death occurs most commonly from the effects of local tumor invasion, particularly
asphyxiation. In one large study (51) over half of the patients died of suffocation,
while the others died of a combination of effects from the primary tumor and
metastases.
Medullary Thyroid Carcinoma 313

Prognostic Factors
Although prognosis with ATC is very poor, certain features predict a more favor-
able response to therapy. Like differentiated thyroid carcinoma, patient age (45)
and tumor stage at the time of diagnosis are the most important variables influ-
encing prognosis. In one large study (45), mean survival was 8.1 months if the
disease was confined to the neck, compared with 3.3 months when it was distantly
metastatic. The Mayo Clinic (46) reported that relatively favorable prognostic
features were unilateral tumors, a tumor diameter ⬍5 cm, no invasion of adja-
cent tissue, and absence of cervical node involvement. Cox multivariate analysis
showed that extent of disease, tumor size, histologic cell type, and surgical treat-
ment were the most significant variables. In a recent report spanning 1985–1995
years, with 893 cases (1), the 10-year relative survival in the United States was
14%, with 77% of deaths within 1 year and 100% of all deaths by year 4. Age ⬍45
years at diagnosis demonstrated dramatically better 5-year survival at 55%. Small
tumor size and absent extrathyroidal extension were more common in patients sur-
viving more than five years. In a study of 516 patients from the SEER database, on
multivariate analysis only age ⬍60, extent of disease, and combined therapy with
surgical resection and external beam radiotherapy were independent prognostic
factors (50).

Treatment
ATCs are very resistant to any form of therapy and are rarely cured. Surgery,
chemotherapy, or radiotherapy used separately generally have not been effective
(52–54). The best survival rates occur with combined surgery, accelerated and
hyperfractionated external irradiation, and chemotherapy (55).

Surgical Therapy
Thyroid Surgery
Less than total thyroidectomy, with resection of the involved adjacent neck tissues
and cervical lymph nodes, should be done if possible, without resorting to radical
surgery in surgically incurable patients. In the Mayo Clinic series (46), 41% of the
patients had surgical resection—mostly total lobectomy with subtotal resection of
the contralateral lobe—in an attempt to cure the patient. In the MD Anderson series
(45), approximately 45% of patients underwent total thyroidectomy. However, in
neither series did the surgical extent influence the survival. Total thyroidectomy and
radical neck dissection result in an increased complication rate without conferring
a clear advantage over a more conservative approach. When the tumor is resectable,
an appropriately aggressive and safe surgical approach is to resect the tumor with
wide margins of adjacent soft tissue on the involved side. When the tumor is not
resectable, attention is directed toward airway management.
314 Sipos and Mazzaferri

Airway Management
Tumor may compromise the airway by compression, displacement, infiltration, and
less often as the result of neurogenic dysfunction. Management involves thyroid
gland resection with decompression of the airway and tracheostomy when the
airway is infiltrated with tumor. Many patients require tracheostomy during their
course, mainly to relieve airway obstruction but also as a precautionary measure
before initiating external radiation therapy. External irradiation may cause tumor
edema that acutely exacerbates airway obstruction. More than half the patients
in the Mayo Clinic series required tracheostomy at some time during the course
of their disease (46). In patients who are not surgical candidates or who do not
wish to have invasive procedures, tracheal stenting is an alternative therapy which
may provide symptomatic relief from the dyspnea while maintaining quality of
life without requiring a tracheostomy (56).

Medical Therapy
External Radiotherapy
After treatment with surgery and conventional radiotherapy, ⬍5% of patients with
ATC survive five years, although survivors who are disease-free at two years
may live for longer periods (37). Because conventional external irradiation fails
to eradicate local disease, larger doses of radiation have been given at closer
intervals with limited success in controlling local disease, but the toxicity is very
high, resulting in severe esophagitis, dysphagia, spinal cord necrosis, and death
(54,57).

Chemotherapy
There is no consensus regarding the selection of chemotherapeutic agents for ATC,
mainly because this tumor is so resistant to chemotherapy. None provides clearly
superior therapeutic efficacy, and the selection is often based as much upon the
side-effect profiles of the drugs as its potential efficacy. Doxorubicin is perhaps
the most commonly used, administered either alone or in combination with other
drugs. There is a dose–response relationship such that patients receiving lower
doxorubicin doses, in the range of 45 mg/m2 , show little or no response, while those
treated with larger doses demonstrate therapeutic responses more consistently
(37). The most frequently applied doxorubicin dosage has been between 60 and
90 mg/m2 body surface (37). With this dosage, one study (58) found the response
rate (complete or partial remission) to be almost 40% in patients with advanced
thyroid carcinoma of all types, although the response rate was only 22% for
ATC treated with doxorubicin monotherapy. Even when remission does occur,
its duration is relatively short, with a median period of 90 (33–560) days in the
20% to 30% who responded to doxorubicin (37). Others believe that doxorubicin
alone does not improve the extremely poor prognosis of ATC. When there is a
response to doxorubicin, pulmonary metastases most frequently respond, followed
Medullary Thyroid Carcinoma 315

by bone metastases and local tumor growth; however, the drug may not control
local disease (37). Because of the poor response to standard chemotherapy, it is
reasonable to recommend a clinical trial of one of the new targeted molecular
therapies.
Doxorubicin and Hyperfractionated Radiation Therapy
In 1983, a new treatment regimen consisting of combination doxorubicin and
hyperfractionated radiation therapy was reported for the treatment of ATC (59).
The protocol consisted of once-weekly administration of low-dose doxorubicin
(10 mg/m2 ) and hyperfractionated radiation therapy carried out with a fractional
dose of 160 cGy per treatment twice a day for 3 day/wk. The total tumor dose
was 5760 cGy delivered in 40 days. The protocol was well tolerated with little
morbidity. Complete tumor response was observed in 84% of the patients and rate
of local tumor control was 68% at two year. Median survival was one year, but
most patients developed distant metastases and died from the disease.
In 1990, a Swedish group (55) reported prospectively treating 16 patients
with a protocol of hyperfractionated radiotherapy, doxorubicin, and debulking
surgery. The radiotherapy was administered preoperatively to a target dose of
30 Gy in three weeks and postoperatively to an additional dose of 16 Gy in 1.5
weeks. Radiotherapy was administered twice daily, five days a week, with a target
dose of 1 Gy per fraction and with a minimum interval of six hour. Doxorubicin
(20 mg) was administered intravenously one to two before the first radiotherapy
session every week. Debulking surgery was feasible in nine patients. Complete
remission of local disease was achieved in 5 of 16, of whom three were alive and
disease-free at 10, 30, and 30 months after diagnosis. Only six patients died of
local disease. This combination regimen was well tolerated despite the patients’
advanced age and disease stage.
In 1994, the same group (60) reported prospectively treating 33 consecutive
patients with ATC according to the above protocol except that the daily fraction of
radiotherapy was increased to 1.3 Gy per fraction after 1988. No patient failed to
complete the protocol because of toxicity. There were no local recurrences in 16
patients (48%). Death was attributed to local failure in only eight patients (24%).
In four patients, survival with no evidence of disease exceeded two years. Local
tumor control was marginally improved in patients treated in the latter part of
the study. The authors concluded that combination-modality treatment of ATC
is feasible and effective despite the patients’ advanced age and locally advanced
disease.
In 1991, a French group (61) reported their results in 20 ATC patients treated
prospectively according to a combination protocol of chemotherapy and external
radiation therapy. Depending on the patient’s age, two types of chemotherapy
were used every four weeks. For those younger than 65 years, a combination
of doxorubicin (60 mg/m2 ) and cisplatin (90 mg/m2 ) was employed; for older
patients, mitoxantrone (14 mg/m2 ) alone was used. Between days 10 and 20 of the
first four chemotherapy cycles, radiotherapy (17.5 Gy) was given in seven fractions
316 Sipos and Mazzaferri

to the neck and superior mediastinum. Of the total, three patients (15%) survived
longer than 20 months. Complete tumor response was seen in distant metastases,
which were the cause of death in 70% of the patients in this series. Toxicity was the
main limiting factor—all patients developed pharyngoesophagitis and tracheitis
after the first or second cycle of radiotherapy. It was severe in 60% of the patients,
requiring one to two weeks’ rest before resuming treatment. Hematologic toxicity
occurred in 40% of the patients, while cardiotoxicity was seen in 25% of those
treated with doxorubicin.

PRIMARY THYROID LYMPHOMA


PTL is an uncommon but potentially life-threatening disorder that often poses
a major diagnostic and therapeutic dilemma because most clinicians rarely see
these tumors. PTL typically arises in the setting of chronic thyroiditis, sometimes
in patients with long-standing hypothyroidism. The true character of the neoplasm
may go unrecognized for several months, until the development of airway obstruc-
tion symptoms; then, unfortunately, an erroneous diagnosis of ATC is sometimes
made. It is important to identify the tumor correctly as a PTL, since treatments and
prognoses differ, and diagnosis at an early stage is associated with an excellent
prognosis.

Incidence
The annual incidence of PTL in the United States is less than one in two million
persons (62). It is 10-fold less frequent than gastrointestinal lymphoma, the most
common extranodal non-Hodgkin’s lymphoma, and only slightly more common
than breast lymphoma (62). The likelihood that a thyroid nodule is a lymphoma
is ⬍1 per 1000, whereas secondary involvement of the thyroid with lymphoma
occurs in approximately 10% of patients who die of the disease (62). Despite
the relative rarity of PTL, its incidence has been rising, and it now constitutes
approximately 5% of all thyroid malignancies, with estimates ranging from 2%
to 8% (62). Its frequency may be increasing for several reasons. Many PTLs
were incorrectly diagnosed in the past as anaplastic small-cell thyroid carcinoma
(62). Also, pathologists now have become more adept at separating lymphoma
from advanced Hashimoto’s thyroiditis, which is important, since PTL typically
develops in the setting of preexisting lymphocytic thyroiditis (62). Some of the
increase seems related to more aggressive diagnostic evaluation of thyroid nodules.
Finally, the increasing frequency of PTL has paralleled the rising incidence of
Hashimoto’s thyroiditis in the United States.

Age and Sex Distribution


Contrary to other lymphomas in which males predominate, female preponderance
is the rule for PTL (62), probably because it originates from active lymphoid cells
in chronic lymphocytic thyroiditis, which occurs more often in women. Among
Medullary Thyroid Carcinoma 317

812 PTL patients, females outnumbered males almost 3:1, and the mean age was
62.7 years (62). However, the female-to-male ratio for those younger than 60 years
was 1.5:1 compared with 5:1 in those older than 60 years (62). Although primarily
a disease of older women, about one-third of these patients are younger than
60 years (62).

Hashimoto’s Thyroiditis (Chronic Lymphocytic Thyroiditis)


and Thyroid Lymphoma
An association exists between PTL and Hashimoto’s thyroiditis, but how this
occurs is unclear. In one study, the relative risk of PTL among people in Sweden
with Hashimoto’s thyroiditis, after an average follow-up of 8.5 years, was 67-
fold greater than expected (63). Another study (64) from Japan found an 80-
fold increased frequency of PTL among 5592 women aged 25 years or older
with chronic thyroiditis. The average interval between the diagnosis of chronic
thyroiditis and PTL was 9.2 years.

Pathology
Lymphoma Cell Types
Virtually all PTLs are B-cell types, which can be identified by monoclonal anti-
bodies (62). Many extranodal lymphomas, including those in the thyroid, arise
from mucosa-associated lymphoid tissue (MALT); they are a special group of B-
cell lymphomas (62). Most are diffuse (as opposed to nodular) lymphomas, with
variable cellular features typical of low-grade malignancies. Cases classified as
immunoblastic B-cell lymphoma are uncommon but are typically quite aggressive.
Histologic Features of Hashimoto’s Disease
The histologic features of Hashimoto’s disease and PTL are often difficult to dif-
ferentiate. Hashimoto’s features include well-differentiated lymphocytes, plasma
cells, macrophages, and lymphoid follicles with germinal centers and scattered
Langerhans-type giant cells. The reactive lymphoid follicles are largely composed
of B lymphocytes that mainly produce polyclonal IgG. The inflammatory cells are
associated with both damage and stimulation of the thyroid follicular cells, causing
atrophy, metaplasia, and hyperplasia, which leads to considerable loss of thyroid
colloid. Most of the inflammatory cells lie in the interstitial tissue between thy-
roid follicles and are typically confined to the thyroid parenchyma, with minimal
infiltration of the thyroid capsule.
Histologic Features of Lymphoma
The normal thyroid tissue is extensively infiltrated with abnormal lymphoid cells
that often penetrate the thyroid capsule, extending into adjacent soft tissues. The
lymphoma is usually composed of small cells monotonously similar to one another,
which represents a distinct difference from those seen in autoimmune thyroiditis.
318 Sipos and Mazzaferri

The border between the two may be sharply defined, or there may be a transi-
tional zone in which elements of both are mixed. The lymphoma cells tend to
displace, distort, and replace the thyroid epithelium. Most thyroid follicles are
packed and distended with lymphoma cells and loose colloid (62). Lymphoma
cells infiltrate blood vessel walls in at least one-fourth of the cases and lym-
phoma cells may undergo necrosis. Leukocyte common antigen can usually be
found and substantial proportions of lymphomas contain both monoclonal heavy
and light chain immunoglobulins, characteristics that allow their differentiation
from small-cell carcinomas. Also, identification of monoclonal light chains usu-
ally allows differentiation of a malignant proliferation from a polyclonal benign
lymphocyte inflammation, although monoclonal gammopathy occasionally can be
demonstrated in Hashimoto’s disease.

Clinical Features
Symptoms and Signs
Duration of Symptoms
Most patients with PTL have compressive symptoms caused by a rapidly expand-
ing goiter that is invading and compressing neck structures. Symptoms typically
are short in duration, averaging less than five months (62), which contrasts sharply
with the indolent course of Hashimoto’s thyroiditis.

Compression Symptoms
Patients characteristically have a subacute onset of symptoms caused by tumor
pressing and growing into vital neck organs, particularly the trachea, laryngeal
nerve, esophagus, and neck muscles. The most common complaints are hoarseness
(21%); dysphagia, dyspnea, and stridor (19%); neck pressure (5%); and neck
pain (5%) (62). These are important signals of extrathyroidal tumor extension
that should alert the clinician to the malignant nature of the goiter. Stridor and
hoarseness often occur together, and when they do, laryngeal nerve paralysis
is almost invariably seen on laryngoscopy. Symptoms of invasion, particularly
dyspnea and dysphagia, are ominous and portend a poor outcome.

Goiter
The most common presenting feature is recent growth of goiter, which is experi-
enced by about half the patients (62). A rapidly enlarging goiter typically becomes
apparent to the patient over several weeks to a few months before the diagnosis
of lymphoma is established, although longer periods have been reported. A goiter
was present for a year or more in 16% of 556 patients reported in nine large stud-
ies (62). Only one thyroid lobe may enlarge or a discrete nodule may be palpable
within the gland, or there may be diffuse thyromegaly. The goiter is nearly always
firm or hard, is usually nontender and is often fixed to adjacent tissues, a sign that
tumor is invading neck structures. Other signs of extrathyroidal spread include
ill-defined thyroid borders with extension laterally or retrosternally.
Medullary Thyroid Carcinoma 319

Lymph Node Involvement and Retrosternal Extension


Large (⬎2 cm), nontender, and matted cervical lymph nodes are found on phys-
ical examination in up to half the patients (62). The chest roentgenogram may
show retrosternal involvement or mediastinal widening, although lymph nodes
elsewhere are seldom enlarged.

Coexisting Hashimoto’s Thyroiditis


Almost all patients with PTL have clinical or histologic evidence of lymphocytic
thyroiditis at the time of diagnosis (62). In 10 large series (62), antithyroglobulin
and antimicrosomal antibodies are found in almost 75% of 539 patients, and 83%
had histologic evidence of Hashimoto’s thyroiditis.

Thyroid Dysfunction
Although laboratory evidence of autoimmune thyroiditis is common, few patients
have overt hypothyroidism. Most are euthyroid or only mildly hypothyroid, show-
ing minimal elevation in serum thyroid-stimulating hormone (TSH) and otherwise
normal thyroid function tests at the time PTL is discovered. When lymphoma was
diagnosed, overt hypothyroidism was seen in only approximately 8% of 366
patients from seven large studies (62). However, despite its relative infrequency,
a goiter that suddenly enlarges during appropriate long-term levothyroxine ther-
apy is an important clinical clue to the diagnosis of lymphoma. Thyrotoxicosis is
extremely uncommon but can occur in patients with lymphoma.

Diagnosis
Differential Diagnosis
The diagnoses that should be considered in a patient with a rapidly enlarging
thyroid mass are PTL, ATC, MTC, multinodular goiter, or colloid nodule with
acute hemorrhage, and various inflammatory disorders including acute, subacute,
and Hashimoto’s thyroiditis.

Early Diagnosis
It is important to diagnose PTL at an early stage. A German study showed that
non-Hodgkin’s lymphomas with a low-grade histology can, over time, undergo
transformation into high-grade tumors (62). In one large study from Japan (65),
the frequency of high-grade PTLs decreased significantly after 1982, which the
authors attributed to early diagnosis and therapy. Although the diagnosis is usually
made in response to a rapidly enlarging and symptomatic thyroid mass, PTL can
sometimes be discovered at an earlier stage by investigating more subtle clues in
patients with known Hashimoto’s thyroiditis. PTL should be considered in a patient
with Hashimoto’s thyroiditis with a discrete hypofunctional area on 123 I scanning,
whether or not there is a palpable nodule (62). PTL also should be suspected when
the entire thyroid gland with Hashimoto’s disease or any portion of it enlarges
320 Sipos and Mazzaferri

Table 1 Indications for Fine-Needle Aspiration or Open Surgical Biopsy in Patients


Suspected of Having Thyroid Lymphoma

Thyroid mass with


Rapid enlargement
Hoarseness
Laryngeal paralysis
Dysphagia
Stridor
Dyspnea
Pain
Hashimoto’s thyroiditis
Hypofunctional (cold) nodule
Hypofunctional area on scan without nodule goiter or nodule enlargement despite
thyroid hormone therapy
Monoclonal gammopathy

Source: From Ref. 62.

during thyroid hormone therapy or when palpable cervical or supraclavicular


lymph nodes develop.
Routine Laboratory Studies
Serum Chemistries and Immunoglobulins
Routine serum chemistries and hematologic studies are usually normal. The serum
immunoglobulins are occasionally abnormal, showing a monoclonal gammopathy,
and—very rarely—the bone marrow is involved by lymphoma (62).
Thyroid Tests
Thyroid function testing may disclose subclinical or overt hypothyroidism, and
serum antimicrosomal and antithyroglobulin antibodies are often positive.
FNA Biopsy
Selection of Patients
FNA should be considered in patients with features summarized in Table 1 (62). A
diagnosis of lymphoma may be established by FNA cytology or by open surgical
biopsy. Immunophenotyping with flow cytometry or Southern blot analysis for
lymphocytic clones may also be helpful (66).

Imaging Studies
Timing of Studies
Thyroid imaging done with radionuclides, CT scan, MRI, and ultrasonography
usually demonstrate nonspecific abnormalities and are not first-line tests. However,
once the diagnosis is established by FNA, imaging is useful in defining the extent
of disease.
Medullary Thyroid Carcinoma 321

Ultrasonography
PTL usually appears as a well-delineated hypoechoic sold mass intermingled with
echogenic structures. It is easily distinguishable from the surrounding thyroid
tissue even though the latter often shows low echogenicity and nodularity due
to coexisting Hashimoto’s thyroiditis. Most PTLs are discrete solid nodules in a
diffuse goiter. Ultrasonography may disclose contiguous tumor spread into both
thyroid lobes and is sensitive in detecting cervical lymph nodes. It is an important
adjunct to physical examination that is useful both on initial evaluation with FNA
and in follow-up.

Computed Tomography
PTL usually is manifest as one or more areas of low thyroid density. CT scan
appearances are of three types: solitary nodules (80%), multiple nodules (13%),
and diffuse goiter (7%) (62). Both lobes are usually involved with advanced dis-
ease. The tumors have a strong tendency to compress (80%) or infiltrate (53%)
surrounding structures and less often show calcification (7%) and necrosis (7%)
(62). Lymphomas often completely encircle the trachea, which is a characteris-
tic feature of malignancy sometimes termed the “donut sign.” although CT scan
and ultrasonography are both highly sensitive in detecting thyroidal abnormalities
caused by lymphoma, CT scan is better at demarcating intrathoracic tumor exten-
sion and laryngeal invasion and is the preferred radiologic technique for staging
(62).

Magnetic Resonance Imaging


Lymphomas appear as homogeneous iso- or high-intensity areas on T1-weighted
images compared with uninvolved thyroid tissue, which appears homogeneously
high-intensity on T2-weighted images (62). The distinction between tumor and
uninvolved thyroid gland is sometimes more apparent by MRI than CT scan,
but the two are comparable in identifying extrathyroidal extension, and cervical
lymphadenopathy and in staging of lymphoma (67). Hashimoto’s thyroiditis often
shows homogeneous signal intensities on MRI that are indistinguishable from
those of lymphoma.

Radionuclide Scanning
Various radionuclides, including compounds labeled with radioactive iodine, 201 Tl,
67
Ga, 111 In-octreotide, FDG-PET, MIBG, and 99m Tc may be used to study patients
with PTL, but none is specific for the diagnosis. Lymphoma appears as a hypofunc-
tional lesion with 123 I or 99m Tc-pertechnetate thyroid scanning. Like 99m Tc-labeled
compounds, 201 Tl may demonstrate uptake in PTL. Perhaps the best-studied agent
is 67 Ga, which demonstrates uptake not only in malignant lymphoma (86%) but
also in anaplastic carcinoma (90%) and other high-grade malignancies (62).
322 Sipos and Mazzaferri

Staging
Initial Disease Stage
The disease is confined to the thyroid gland (stage IE) in almost half the patients
and within regional lymph nodes (stage IIE) in 43% at the time of diagnosis (62).
In a study of 245 patients (68), ⬍2% had disease outside of regional lymph nodes
when they presented, but 10 eventually developed involvements of bone marrow,
retroperitoneal nodes, and other tissues outside the neck.

Gastrointestinal Lymphoma
Although opinions differ, some suggest that there is an association between PTL
and gastrointestinal lymphomas. Among the approximately 500 cases of PTL
reported from the mid-1930s to 1986, gastrointestinal involvement was found
in 62% of those dying with metastatic PTL—a frequency more than twofold
greater than that reported for all non-Hodgkin’s lymphomas (69). Accordingly,
gastrointestinal infiltration by PTL does not appear to occur by serendipity, which
poses certain practical questions regarding the initial staging evaluation of patients
with thyroid lymphoma.

Staging Workup
Since site-directed radiation alone is often used for lymphoma, accurate staging
of the disease is important, although there is little agreement regarding the extent
of staging evaluation (62). We believe that all patients should have a complete
blood count, blood chemistries, and chest roentgenogram in addition to studies to
investigate the extent of local disease, including regional lymph node involvement
and tumor extension into surrounding neck structures and disease in the chest
and abdomen. Thyroid ultrasonography may be helpful but CT scan or MRI of
the neck, chest, and abdomen should be done routinely. Whether a bone marrow
examination should be done in every patient is even less certain. Several studies
suggest that almost no patients with PTL have bone marrow involvement, and
they conclude that such an examination is usually unnecessary (62,70). There
is also lack of consensus about gastrointestinal evaluation, and at our institution
dedicated gastrointestinal studies are not done routinely. Others routinely examine
the gastrointestinal tract in almost all patients with PTL, although when this was
done in one study only 2.5% had stage IV disease (65).

Prognostic Features
Long-Term Survival
Five-year survival with PTL was almost 60% in 368 patients reported in seven large
series (62). However, survival is dependent upon several variables, particularly
tumor stage and grade.
Medullary Thyroid Carcinoma 323

Stage
In a large study in 1966 from the Mayo Clinic (71), when tumor was limited to the
thyroid (stage IE), only one death occurred; however, involvement of local lymph
nodes (stage IIE) or infiltration of adjacent soft tissues reduced average survival
rate to only 18 months. A more recent study (72) found 5-year survival rates were
91% in stage IE and only 62% in stage IIE patients. Over the past several decades,
5-year survival rates have been at least 75% to 100% when tumor is confined to the
thyroid compared with only 35% to 60% when it is outside the gland (62,73,74).

Tumor Grade
In one study (65), 5-year survival rates were 92% for patients with low-grade
lymphomas, 79% for intermediate-grade lymphomas, and 13% for immunoblastic
type (high-grade) tumors.

Other Prognostic Factors


Reported favorable prognostic factors include longer duration (⬎six months) of
goiter, age ⬍60 years at the time of diagnosis, goiter smaller than 10 cm, and preex-
isting Hashimoto’s thyroiditis (65). Reported negative prognostic factors include
large tumor size and fixation, extracapsular extension, retrosternal involvement,
large-cell lymphoma, unresectable tumor, possibly male sex, advanced age (⬎65),
hoarseness, stridor, hepatosplenomegaly, and axillary lymph nodes (73,75,76).

Therapy
Treatment is not standard because large, randomized, multicenter trials for this
uncommon disorder have not yet been done. Although external radiation has
become the treatment of choice in many centers, controversy exists about the
optimal extent of surgery and the use of chemotherapy. Although survival has
improved over the past few decades, it is not clear which factors are responsible.
There are indications that PTL is being diagnosed at an earlier stage and at a lower
grade, that adequate staging is being performed more regularly, and that radiation
therapy is being used in nearly all cases (62).

Surgery
Some (70) report stated that patients undergoing total macroscopic tumor removal
fare considerably better than those with persistent tumor after surgery; 5-year
survival rates in the two groups, respectively, were approximately 65% and 22%.
Similar results were reported from the Mayo Clinic (75), where 5-year survival
rates were lower (49 vs. 75%) when patients had obvious residual disease postop-
eratively. Perhaps the least controversial reasons for surgery are for diagnosis and
tumor staging and to relieve airway obstruction. Since the cytologic diagnosis of
lymphoma may be difficult to establish with certainty by FNA, open biopsy may
be necessary. In our institution, this is frequently the case, and surgeons excise as
324 Sipos and Mazzaferri

much malignant tissue as possible without performing radical surgery. In addition,


a tracheostomy is done if there is any question about the integrity of the airway.

Airway Protection
Patients commonly present with a rapidly enlarging thyroid mass. On occasion—
and sometimes with striking swiftness—this leads to severe tracheal compression
with respiratory distress that may necessitate emergency surgery (62). This com-
plication should be anticipated and an elective tracheostomy performed whenever
airway compromise is a possibility, because death from laryngeal obstruction can
occur in the immediate postoperative period or later with uncontrolled disease
(62). Almost 25% of the patients in a large series from the Mayo Clinic (77)
required an elective tracheostomy; however, there was a 10% rate of infection,
sepsis, or bleeding.

Radiotherapy
This therapy is employed in most centers, especially for patients with stage IE or
IIE thyroid lymphoma. Control of disease in the neck is related both to radiation
dosage and selection of radiation fields and may also be dependent upon the
degree of surgical debulking prior to radiotherapy. The thyroid, bilateral neck,
and mediastinum are treated with at least 40 Gy (4000 rad) given in divided
doses over four to five weeks (62). There are differences of opinion, however,
concerning the radiation fields with some (70) recommending radiation only to
the neck and others (75) advocating irradiation of the axilla and mediastinum.
One group (75) achieved a 59% disease-free survival with about 40 (24–60) Gy in
38 patients, most of whom had intermediate grade histology and stage IE or IIE
disease. None experienced substantial side effects. Despite important differences
in patient cohorts and radiation techniques, overall 5-year survival rates range
from 55% to 70% following external radiation given either alone or with surgery
(62). Results are best for patients with stage IE and IIE disease. For example, one
group (72) reported 5-year survival rates of 91% in patients with stage IE disease
who were treated with 40 Gy.

Chemotherapy
Although chemotherapy is often used as salvage therapy, PTL recurrence rates
as high as 50% and the predominance of diffuse histiocytic lymphoma (which
responds particularly well to chemotherapy) have led some to propose using
chemotherapy either alone or as an adjuvant to surgery or radiotherapy in most
patients (62). Others recommend chemotherapy only for patients with poor prog-
nostic factors (62). Most chemotherapy regimens consist of cyclophosphamide,
adriamycin, vincristine, and prednisone, with or without bleomycin (CHOP ±
bleomycin) or minor alterations of this combination (62). A review of the pub-
lished literature suggests that the addition of chemotherapy to radiation signifi-
cantly lowered distant and overall recurrence (78).
Medullary Thyroid Carcinoma 325

Failure Patterns
Recurrence rates vary dramatically among various series because authors do not
always explicitly distinguish between persistent and recurrent disease, but most
(75%) recurrences are detected within the first year following therapy and com-
monly involve lymph nodes (68,75,79). The gastrointestinal tract, lung, liver,
pancreas, and kidney are much less frequently involved.
Local Disease Failure
Local disease failure usually occurs in 25% to 35% of patients after 40 Gy of
external radiation, although results vary (62,70,72,78). The rate of local failure
is related to the amount of residual disease after initial thyroidectomy and to the
placement of the radiation fields and the use of chemotherapy. One study (75)
reported no failures within the treatment fields in patients without residual disease
when radiation therapy was started, but otherwise there were twice as many failures
following radiotherapy to the neck alone compared with radiotherapy to both the
neck and mediastinum (60 vs. 36%).
Distant Recurrence
Distant recurrences occur to lung, gastrointestinal tract, liver, central nervous
system, and kidneys. One study (68) of 245 patients found that the gastrointestinal
tract was infrequently involved, but others (69) reported its involvement in 62%
of patients dying with metastatic thyroid lymphoma. An autopsy study (80) found
the most common sites of involvement were the gastrointestinal tract (100%), lung
and kidney (each 63%), and liver and pancreas (each 50%).
Survival Following Relapse
Salvage therapy has little impact upon the disease once relapse occurs (62). One
group (70) reported that disease-free survival was almost identical to overall actual
survival, emphasizing the poor response of recurrent disease to therapy. Another
study (79) reported an eight month (1–21 months) mean survival in six patients
following relapse.

REFERENCES
1. Hundahl SA, Fleming ID, Fremgen AM, et al. A National Cancer Data Base report
on 53,856 cases of thyroid carcinoma treated in the US, 1985–1995. Cancer 1998;
83:2638–2648.
2. Sizemore GW. Medullary carcinoma of the thyroid gland. Semin Oncol 1987; 14:306–
314.
3. Mulligan LM, Kwok JBJ, Healey CS, et al. Germ-line mutations of the RET proto-
oncogene in multiple endocrine neoplasia type 2A. Nature 1993; 363:458–460.
4. Brandi ML, Gagel RF, Angeli A, et al. CONSENSUS: Guidelines for Diagnosis and
Therapy of MEN Type 1 and Type 2. J Clin Endocrinol Metab 2001; 86(12):5658–
5671.
326 Sipos and Mazzaferri

5. Verga U, Fugazzola L, Cambiaghi S, et al. Frequent association between MEN 2A


and cutaneous lichen amyloidosis. Clin Endocrinol (Oxf) 2003; 59(2):156–161.
6. Eng C. The RET proto-oncogene in multiple endocrine neoplasia type 2 and Hirsch-
prung’s disease. N Engl J Med 1996; 335:943–951.
7. Raue F, Zink A. Clinical features of multiple endocrine neoplasia type 1 and type 2.
Horm Res 1992; 38(suppl 2):31–35.
8. Gagel RF, Robinson MF, Donovan DT, et al. Medullary thyroid carcinoma: Recent
progress. J Clin Endocrinol Metab 1993; 76:809–814.
9. Eng C, Clayton D, Schufenecker I, et al. The relationship between specific RET
proto-oncogene mutations and disease phenotype in multiple endocrine neoplasia
type 2: International RET mutation consortium analysis. JAMA 1996; 276:1575–
1579.
10. Machens A, Niccoli-Sire P, Hoegel J, et al. Early malignant progression of hereditary
medullary thyroid cancer. N Engl J Med 2003; 349(16):1517–1525.
11. Howe JR, Norton JA, Wells SA Jr. Prevalence of pheochromocytoma and hyper-
parathyroidism in multiple endocrine neoplasia type 2A: Results of long-term follow-
up. Surgery 1993; 114:1070–1077.
12. Machens A, Brauckhoff M, Holzhausen HJ, et al. Codon-Specific Development of
Pheochromocytoma in Multiple Endocrine Neoplasia Type 2. J Clin Endocrinol Metab
2005; 90(7):3999–4003.
13. Gimm O, Neuberg DS, Marsh DJ, et al. Over-representation of a germline RET
sequence variant in patients with sporadic medullary thyroid carcinoma and somatic
RET codon 918 mutation. Oncogene 1999; 18:1369–1373.
14. Russo D, Arturi F, Chiefari E, et al. A case of metastatic medullary thyroid carcinoma:
Early identification before surgery of an RET proto-oncogene somatic mutation in
fine-needle aspirate specimens. J Clin Endocrinol Metab 1997; 82:3378–3382.
15. Elisei R, Cosci B, Romei C, et al. Prognostic significance of somatic RET oncogene
mutations in sporadic medullary thyroid cancer: A 10 years follow up study. J Clin
Endocrinol Metab 2008; 93(3):682–687.
16. Mendelsohn G. Markers as prognostic indicators in medullary thyroid carcinoma. Am
J Clin Pathol 1991; 95:297–298.
17. Elisei R, Bottici V, Luchetti F, et al. Impact of routine measurement of serum
calcitonin on the diagnosis and outcome of medullary thyroid cancer: Experience
in 10,864 patients with nodular thyroid disorders. J Clin Endocrinol Metab 2004;
89(1):163–168.
18. Cooper DS, Doherty GM, Haugen BR, et al. Management Guidelines for Patients
with Thyroid Nodules and Differentiated Thyroid Cancer. Thyroid 2006; 16(2):109–
141.
19. Cohen R, Buchsenschutz B, Estrade P, et al. Causes of death in patients suffering from
medullary thyroid carcinoma: Report of 119 cases. Presse Med 1996; 25:1819–1822.
20. Dottorini ME, Assi A, Sironi M, et al. Multivariate analysis of patients with medullary
thyroid carcinoma—Prognostic significance and impact on treatment of clinical and
pathologic variables. Cancer 1996; 77:1556–1565.
21. Modigliani E, Cohen R, Campos JM, et al. Prognostic factors for survival and for
biochemical cure in medullary thyroid carcinoma: Results in 899 patients. The GETC
Study Group. Groupe d’etude des tumeurs a calcitonine. Clin Endocrinol (Oxf) 1998;
48(3):265–273.
22. Wells SA, Baylin SB, Leight Geal. The importance of early diagnosis in patients with
hereditary medullary thyroid carcinoma. Ann Surg 1982; 195:204.
Medullary Thyroid Carcinoma 327

23. Lips CJM, Landsvater RM, Höppener JWM, et al. Clinical screening as compared
with DNA analysis in families with multiple endocrine neoplasia type 2A. N Engl J
Med 1994; 331:828–835.
24. Iler MA, King DR, Ginn-Pease ME, et al. Multiple endocrine neoplasia type 2A: A
25-year review. J Pediatr Surg 1999; 34:92–96.
25. Dralle H, Gimm O, Simon D, et al. Prophylactic thyroidectomy in 75 children
and adolescents with hereditary medullary thyroid carcinoma: German and Austrian
experience. World J Surg 1998; 22:744–751.
26. Moley JF, DeBenedetti MK, Dilley WG, et al. Surgical management of patients with
persistent or recurrent medullary thyroid cancer. J Intern Med 1998; 243:521–526.
27. Chen HB, Roberts JR, Ball DW, et al. Effective long-term palliation of symptomatic,
incurable metastatic medullary thyroid cancer by operative resection. Ann Surg 1998;
227:887–893.
28. Fromigue J, De Baere T, Baudin E, et al. Chemoembolization for liver metastaes
from medullary thyroid carcinoma. J Clin Endocrinol Metab 2006; 91(7):2496–2499.
29. Frank-Raue K, Ziegler R, Raue F. The use of octreotide in the treatment of medullary
thyroid carcinoma. Horm Metab Res 1993; 27(suppl):44–47.
30. Machens A, Behrmann C, Dralle H. Chemoembolization of liver metastases from
medullary thyroid carcinoma. Ann Intern Med 2000; 132(7):596–597.
31. de Groot JW, Zonnenberg BA, Quarles vU-M, et al. A phase-II trial of imatinib
therapy for metastatic medullary thyroid carcinoma. J Clin Endocrinol Metab 2007;
92(9):3466–3469.
32. Frank-Raue K, Fabel M, Delorme S, et al. Efficacy of imatinib mesylate in advanced
medullary thyroid carcinoma. Eur J Endocrinol 2007; 157(2):215–220.
33. Giraudet AL, Vanel D, Leboulleux S, et al. Imaging medullary thyroid carcinoma
with persistent elevated calcitonin levels. J Clin Endocrinol Metab 2007; 92(11):4185–
4190.
34. Laure GA, Al Ghulzan A, Auperin A, et al. Progression of medullary thyroid car-
cinoma: Assessment with calcitonin and carcinoembryonic antigen doubling times.
Eur J Endocrinol 2008; 158(2):239–246.
35. Skinner MA, DeBenedetti MK, Moley JF, et al. Medullary thyroid carcinoma in
children with multiple endocrine neoplasia types 2A and 2B. J Pediatr Surg 1996;
31:177–182.
36. Wells SA Jr, Skinner MA. Prophylactic thyroidectomy, based on direct genetic testing,
in patients at risk for the multiple endocrine neoplasia type 2 syndromes. Exp Clin
Endocrinol Diabetes 1998; 106:29–34.
37. Mazzaferri EL. Undifferentiated thyroid carcinoma and unusual thyroid malignan-
cies. In: Mazzaferri EL, Samaan N, eds. Endocrine Tumors. Boston, MA: Blackwell
Scientific Publishing Co., 1993:378–398.
38. Belfiore A, La Rosa GL, Padova G, et al. The frequency of cold thyroid nodules
and thyroid malignancies in patients from an iodine-deficient area. Cancer 1987;
60:3096–3102.
39. Wallin G, Backdahl M, Tallroth E, et al. Co-existent anaplastic and well differentiated
thyroid carcinomas: A nuclear DNA study. Eur J Surg Oncol 1989; 15:43–48.
40. Mooradian AD, Allam CK, Khalil MF, et al. Anaplastic transformation of thyroid
cancer: Report of two cases and review of the literature. J Surg Oncol 1983; 23:95–98.
41. Aldinger KA, Samaan NA, Ibanez ML, et al. Anaplastic carcinoma of the thyroid:
A review of 84 cases of spindle and giant cell carcinoma of the thyroid. Cancer 1978;
41:2267–2275.
328 Sipos and Mazzaferri

42. Carcangiu ML, Steeper T, Zampi G, et al. Anaplastic thyroid carcinoma. A study of
70 cases. Am J Clin Pathol 1985; 83:135–158.
43. LiVolsi VA, Brooks JJ, Arendash Durand B. Anaplastic thyroid tumors. Immuno-
histology. Am J Clin Pathol 1987; 87:434–442.
44. Fagin JA, Matsuo K, Karmakar A, et al. High prevalence of mutations of the p53
gene in poorly differentiated human thyroid carcinomas. J Clin Invest 1993; 91:179–
184.
45. Venkatesh YS, Ordonez NG, Schultz PN, et al. Anaplastic carcinoma of the thyroid.
A clinicopathologic study of 121 cases. Cancer 1990; 66:321–330.
46. Nel CJ, van Heerden JA, Goellner JR, et al. Anaplastic carcinoma of the thyroid: A
clinicopathologic study of 82 cases. Mayo Clin Proc 1985; 60:51–58.
47. Higashi T, Ito K, Mimura T, et al. Clinical evaluation of 67 Ga scanning in the
diagnosis of anaplastic carcinoma and malignant lymphoma of the thyroid. Radiology
1981; 141:491–497.
48. Takashima S, Morimoto S, Ikezoe J, et al. CT evaluation of anaplastic thyroid
carcinoma. AJNR 1990; 11:361–367.
49. Woolner LB, Beahrs OH, Black BM, et al. Classification and prognosis of thyroid
carcinoma. Am J Surg 1961; 102:354–387.
50. Kebebew E, Greenspan FS, Clark OH, et al. Anaplastic thyroid carcinoma. Treatment
outcome and prognostic factors. Cancer 2005; 103(7):1330–1335.
51. Tallroth E, Wallin G, Lundell G, et al. Multimodality treatment in anaplastic giant
cell thyroid carcinoma. Cancer 1987; 60:1428–1431.
52. Asakawa H, Kobayashi T, Komoike Y, et al. Chemosensitivity of anaplastic thy-
roid carcinoma and poorly differentiated thyroid carcinoma. Anticancer Res 1997;
17:2757–2762.
53. Lu WT, Lin JD, Huang HS, et al. Does surgery improve the survival of patients
with advanced anaplastic thyroid carcinoma? Otolaryngol Head Neck Surg 1998;
118:728–731.
54. Mitchell G, Huddart R, Harmer C. Phase II evaluation of high dose accelerated
radiotherapy for anaplastic thyroid carcinoma. Radiother Oncol 1999; 50:33–38.
55. Tennvall J, Lundell G, Hallquist A, et al. Combined doxorubicin, hyperfractionated
radiotherapy, and surgery in anaplastic thyroid carcinoma: Report on two protocols.
Cancer 1994; 74:1348–1354.
56. Ribechini A, Bottici V, Chella A, et al. Interventional bronchoscopy in the treatment
of tracheal obstruction secondary to advanced thyroid cancer. J Endocrinol Invest
2006; 29(2):131–135.
57. Simpson WJ. Anaplastic thyroid carcinoma: A new approach. Can J Surg 1980;
23:25–27.
58. Ahuja S, Ernst H. Chemotherapy of thyroid carcinoma. J Endocrinol Invest 1987;
10:303–310.
59. Kim JH, Leeper RD. Treatment of anaplastic giant and spindle cell carcinoma of the
thyroid gland with combination Adriamycin and radiation therapy. A new approach.
Cancer 1983; 52:954–957.
60. Tennvall J, Tallroth E, el Hassan A, et al. Anaplastic thyroid carcinoma. Doxorubicin,
hyperfractionated radiotherapy and surgery. Acta Oncol 1990; 29:1025–1028.
61. Schlumberger M, Parmentier C, Delisle MJ, et al. Combination therapy for anaplastic
giant cell thyroid carcinoma. Cancer 1991; 67:564–566.
Medullary Thyroid Carcinoma 329

62. Mazzaferri EL, Oertel YC. Primary malignant lymphoma and related lymphopro-
liferative disorders. In: Mazzaferri EL, Samaan N, eds. Endocrine Tumors. Boston,
MA: Blackwell Scientific Publishing, 1993:348–377.
63. Holm LE, Blomgren H, Lowenhagen T. Cancer risks in patients with chronic
lymphocytic thyroiditis. N Engl J Med 1985; 312:601–606.
64. Kato I, Tajima K, Suchi T. Chronic thyroiditis as a risk factor of B-cell lymphoma
in the thyroid gland. Jpn J Cancer Res (Amsterdam) 1985; 76:1085–1090.1985.
65. Aozasa K, Inoue A, Tajima K, et al. Malignant lymphomas of the thyroid gland.
Analysis of 79 patients with emphasis on histologic prognostic factors. Cancer 1986;
58:100–104.
66. Wozniak R, Beckwith L, Ratech H, et al. Maltoma of the thyroid in a man with
Hashimoto’s thyroiditis. J Clin Endocrinol Metab 1999; 84:1206–1209.
67. Takashima S, Nomura N, Noguchi Y, et al. Primary thyroid lymphoma: Evaluation
with US, CT, and MRI. J Comput Assist Tomogr 1995; 19:282–288.
68. Compagno J, Oertel JE. Malignant lymphoma and other lymphoproliferative disorders
of the thyroid gland. A clinicopathologic study of 245 cases. Am J Clin Pathol 1980;
74:1–11.
69. Stone CW, Slease RB, Brubaker D, et al. Thyroid lymphoma with gastrointestinal
involvement: Report of three cases. Am J Hematol 1986; 21:357–365.
70. Tupchong L, Hughes F, Harmer CL. Primary lymphoma of the thyroid: Clinical
features, prognostic factors, and results of treatment. Int J Radiat Oncol Biol Phys
1986; 12:1813–1821.
71. Woolner LB, McConahey WM, Beahrs OH, et al. Primary malignant lymphoma of
the thyroid: Review of forty-six cases. Am J Surg 1966; 111:502–523.
72. Vigliotti A, Kong JS, Fuller LM, et al. Thyroid lymphomas stages IE and IIE:
Comparative results for radiotherapy only, combination chemotherapy only, and mul-
timodality treatment. Int J Radiat Oncol Biol Phys 1986; 12:1807–1812.
73. Pedersen RK, Pedersen NT. Primary non-Hodgkin’s lymphoma of the thyroid gland:
A population based study. Histopathology 1996; 28:25–32.
74. Sasai K, Yamabe H, Haga H, et al. Non-Hodgkin’s lymphoma of the thyroid – A
clinical study of twenty-two cases. Acta Oncol 1996; 35:457–462.
75. Blair TJ, Evans RG, Buskirk SJ, et al. Radiotherapeutic management of primary
thyroid lymphoma. Int J Radiat Oncol Biol Phys 1985; 11:365–370.
76. Shaw JH, Dodds P. Carcinoma of the thyroid gland in Auckland, New Zealand. Surg
Gynecol Obstet 1990; 171:27–32.
77. Devine RM, Edis AJ, Banks PM. Primary lymphoma of the thyroid: A review of the
Mayo Clinic experience through 1978. World J Surg 1981; 5:33–38.
78. Doria R, Jekel JF, Cooper DL. Thyroid lymphoma: The case for combined modality
therapy. Cancer 1994; 73:200–206.
79. Makepeace AR, Fermont DC, Bennett MH. Non-Hodgkin’s lymphoma of the thyroid.
Clin Radiol 1987; 38:277–281.
80. Souhami L, Simpson J, Carruthers JS. Malignant lymphoma of the thyroid gland.
Int J Radiat Oncol Biol Phys 1980; 6:1143–1147.
8
Pediatric Thyroid Disorders

Rosalind S. Brown
Harvard Medical School, Boston, Massachusetts, U.S.A.

INTRODUCTION
Because of the important maturational effects of thyroid hormone, abnormalities
in thyroid function in the pediatric age range may have profound effects on linear
growth and skeletal development. In addition, both hypothyroidism and hyperthy-
roidism in infants and children younger than 3 years can cause permanent cognitive
deficits, reflecting the pivotal role of thyroid hormone on brain development at this
time. The practitioner caring for infants and children with thyroid disorders should
be familiar with these unique aspects of the growing child, and also with differ-
ences in clinical presentation, etiology, consequences, and treatment of thyroid
disease in the pediatric age range.

MATURATION OF THYROID FUNCTION AND REGULATION


Thyroid Gland Development
The thyroid gland is derived from the fusion of a medial out-pouching from
the floor of the primitive pharynx, the precursor of the T4-producing follicular
cells, and bilateral evaginations of the fourth pharyngeal pouch, which give rise
to the parafollicular or calcitonin (C-) secreting cells. Commitment toward a
thyroid-specific phenotype as well as the growth and descent of the thyroid anlage
into the neck results from the coordinate action of a number of transcription
factors, including thyroid transcription factor (TTF) 1 (also called TTF1, Nkx2-1,
or T/ebp), TTF2 (also called FOXE1), and PAX8 (1). When the pharyngeal region

331
332 Brown

The timing of fetal thyroid gland


organogenesis and function

Fetus partially dependent


on maternal T4

Maturation of hypo-
thalamic-pituitary axis
Increasing thyroid
sensitivity to TSH
Increasing TSH
T4, T3,TSH accompanied by
Embryogenesis
levels low ↑T4,fT4,TBG
and descent of
thyroid rT3 high T3 remains low, rT3 high

10 20 30 40
Weeks gestation
Figure 1 The timing of fetal thyroid gland organogenesis and function. See text for details.
Abbreviations: T4, thyroxine; T3, triiodothyronine; TSH, thyrotropin; fT4, free T4; rT3,
reverse T3.

of the thyroid anlage contracts, a narrow stalk, known as the thyroglossal duct,
remains and subsequently atrophies. Usually the median anlage grows caudally so
that no lumen is left in the tract of its descent. An ectopic thyroid and persistent
thyroglossal duct or cyst form as a consequence of abnormalities of the thyroid
descent. Caudal migration of the thyroid anlage is also related to the descent of
the heart and great vessels (2).

Fetal Thyroid Function


Embryogenesis is largely complete by 10 to 12 weeks gestation (Fig. 1). At this
stage, tiny follicle precursors are first seen, thyroglobulin (Tg) can be detected in
follicular spaces, and evidence of iodine uptake and organification is first obtained.
Low concentrations of thyroxine (T4) and triiodothyronine (T3) are detectable in
fetal serum at 10 to 12 weeks (3), although it is likely that a fraction of the thyroid
hormone measurable at this early stage of development is maternal in origin (4).
Thyrotropin (TSH), first detectable in fetal serum at 12 weeks’ gestation, increases
from 18 weeks to term. This is accompanied by an increase in fetal thyroid
radioiodine uptake and a progressive increase in the serum concentrations of both
total T4 and free T4 (5). Pituitary responsiveness to both feedback inhibition and
to stimulation by thyrotropin releasing hormone (TRH) is observed by the end of
the second trimester.
Pediatric Thyroid Disorders 333

The maturation of fetal thyroid function in the second half of pregnancy is


due not only to increasing pituitary maturation and secretion of TSH, but to the
concomitant appearance and increasing expression of the TSH receptor, which
enables the fetal thyroid to respond (6). Thus, there is a progressive increase in the
ratio of free T4 to TSH that rises out of proportion to the increase in total T4 to
TSH ratio, strongly suggesting that there is a change in the thyroid follicular cell
sensitivity to TSH. The increase in total T4 is also due to an increase in thyroxine
binding globulin (TBG) as the liver matures.
Fetal T4 metabolism differs markedly from that of postnatal life. Activity of
Type 1 deiodinase (D1), the major activating deiodinase that converts T4 to T3, is
low throughout gestation. In contrast, D3, the major inactivating deiodinase that
converts T4 to reverse T3, a metabolically inactive compound, is highly expressed
in fetal tissues and in the placenta. As a result, circulating T3 concentrations in
the fetus are low, but reverse T3 and sulfated conjugates of T4 are markedly ele-
vated (5). Similarly, D2, an activating deiodinase that converts T4 to T3, is highly
expressed in fetal brain and pituitary as early as mid-gestation. As a consequence,
fetal brain T3 levels are 60% to 80% those of the adult by fetal age 20 to 26 weeks,
despite the low levels of circulating T3 (7). In the presence of fetal hypothy-
roidism, D2 increases while D3 decreases. These coordinate adjustments are of
critical importance and serve to preserve near-normal brain T3 levels providing
that maternal T4 levels are maintained at normal levels (8).

The Role of the Placenta and of Maternal T4


Under normal circumstances, the placenta has only limited permeability to T4 and
the fetal hypothalamic-pituitary-thyroid system develops relatively independently
of maternal influence. This relative barrier to thyroid hormone transport results
primarily from the high placental content of D3, which serves to inactivate most of
the thyroid hormone presented from the maternal circulation. The iodide released
in this way can then be used for fetal thyroid hormone synthesis. When a significant
T4 gradient between the maternal and fetal compartment exists, however, there is
an increased net flux of maternal thyroid hormone to the fetus. Thus, babies with
severe congenital hypothyroidism (CH) nonetheless have cord T4 concentrations
between 25% and 50% of normal (9). Maternal-fetal T4 transfer is also thought to
be important in the first half of pregnancy before the onset of mature fetal thyroid
function (4).
The transplacental passage of maternal T4, coupled with the adjustments
in brain deiodinase activity discussed earlier, plays a critical role in minimizing
the adverse effects of fetal hypothyroidism. Not only may it help to explain the
normal or near-normal cognitive outcome of hypothyroid fetuses as long as post-
natal treatment is early and adequate, it may also provide a partial explanation
of the relatively normal clinical appearance at birth of over 90% of infants with
CH. In contrast, when both maternal and fetal hypothyroidism occurs, patients are
symptomatic at birth and there is a significant impairment in neuro-intellectual
334 Brown

development despite the initiation of early and adequate postnatal thyroid replace-
ment (10). Isolated maternal hypothyroidism and/or hypothyroxinemia have also
been reported to cause significant cognitive and/or motor delay in the offspring,
although the magnitude of the deficit is not as great as when both fetal and maternal
hypothyroidism are present (11,12).

Other Hormones and Factors


In contrast to thyroid hormone, the placenta is freely permeable to TRH and
to iodide, the latter being essential for fetal thyroid hormone synthesis (13).
The placenta is also permeable to certain drugs and to immunoglobulins of the
immunoglobulin (Ig)G class. Thus, the administration to the mother of excess
iodide, certain drugs [especially propylthiouracil (PTU) or methimazole (MMI)]
or the transplacental passage of TSH receptor Abs from mothers with severe
Graves’ disease or primary myxedema may have significant effects on fetal and
neonatal thyroid function. On the other hand, maternal TSH does not cross the pla-
centa. Similarly, Tg is undetectable in the serum of athyreotic infants, indicating
the absence of any transplacental passage of this large protein.

Neonatal Thyroid Function


Within minutes after birth, a dramatic release of TSH occurs in the newborn infant,
reaching peak levels at 30 minutes of age and persisting with decreasing intensity
for the next 6 to 24 hours (3). The acute increase of TSH may be stimulated in
part by the drop in body temperature of the fetus at birth. In response to the early
postnatal TSH surge, T4, free T4, and T3 levels increase progressively during
the first hours of extrauterine life and peak by 48 hours. The marked increase in
T3 is not only due to the increase in TSH but also due to the maturation of D1
activity and the loss of placental D3 at the time of delivery. In contrast, the elevated
concentrations of the other substrates of D1, reverse T3 and T3 sulfate, decrease
relatively rapidly during the newborn period.

Thyroid Function in Premature Infants


Thyroid function in the premature infant reflects the relative immaturity of the
hypothalamic-pituitary-thyroid axis found in comparable gestational age infants in
utero. Following delivery, there is a surge of T4 and TSH analogous to that observed
in term infants, but the magnitude of the increase is less. In very premature babies
(≤30 weeks, approximately equivalent to 1.5 kg), there may be a fall rather than
an increase in the T4 concentration over the first 1 to 2 weeks of life (14). Reasons
for this fall in total T4 include the clearance of maternal T4 from the neonatal
circulation, relatively immature hypothalamic-pituitary axes, decreased thyroidal
iodide stores, inability to regulate iodide balance, and increased sensitivity to the
thyroid-suppressive effects of excess iodide found in certain skin antiseptics and
drugs to which these babies may be exposed. Also, premature infants are frequently
Pediatric Thyroid Disorders 335

sicker than their more mature counterparts and may be treated by drugs that affect
neonatal thyroid function (particularly dopamine and steroids). In most cases, the
total T4 is more affected than the free T4 (15), a consequence of abnormal protein
binding and/or the decreased TBG in these babies with immature liver function.
Despite the reduced total T4, the TSH concentration is not usually abnor-
mal. The occasional finding of an elevated TSH may reflect true primary hypothy-
roidism in some cases but probably reflects recovery from the sick euthyroid
syndrome in most infants, analogous to the situation in adults who are recovering
from severe illness.

Infants and Children


After the acute perturbations of the neonatal period, there is a slow and progressive
decline in the concentrations of T4, free T4, T3, and TSH during infancy and
childhood age- and gender-specific normative values of thyroid function in a large
population of infants and children have been published and should be referred
to (16–19), although precise values may vary somewhat, depending upon the
commercial method employed. Both the serum TSH and T3 concentrations, in
particular, tend to be higher in children than in adults, especially in the first 2
years of life. Use of adult values, provided by many hospital laboratories, should
be interpreted with caution as it may result in inappropriate concern and referral.
The 24-hour thyroidal 123 I uptake is also much greater in the neonate, although
similar to adult values after the first month. The T4 turnover rate is also higher
and its serum half-life reduced.

THYROID DISEASE IN INFANCY


Congenital Hypothyroidism
CH, that is, hypothyroidism at birth, is the commonest treatable cause of mental
retardation. Worldwide, the most common cause of CH is iodine deficiency, a prob-
lem that continues to afflict almost one billion people despite international efforts
aimed at its eradication. In such cases, CH is endemic (“endemic cretinism”). In
iodine-sufficient areas and in areas of borderline iodine deficiency, CH is usually
sporadic and occurs in 1 in 3000 to 1 in 4000 infants. Females are affected twice
as often as males. In the United States, CH is less frequent in African-Americans
and more common among Hispanics and Asians. There is an increased incidence
of mild, transient CH in patients with Down syndrome (20).
Because optimal outcome treatment must be initiated soon after birth before
affected infants are recognizable clinically, neonatal screening programs have
been introduced in most industrialized areas of the world. There continues to
be some disagreement as to whether minor neuro-intellectual sequelae remain in
the most severely affected infants, but there is no doubt that the main objective
of screening, the eradication of mental retardation, has been achieved. Newborn
screening has also permitted an elucidation of the prevalence of the various causes
336 Brown

of CH, including a series of transient disorders found predominantly in premature


infants.

Screening Strategies for CH


Three screening strategies for the detection of CH are used. In much of North
America, primary T4 screening is performed with TSH reserved for those speci-
mens with a low T4 (usually the lowest 3rd-20th percentile). In other programs,
a primary TSH is employed, or, alternately, both T4 and TSH are measured. In
practice, the choice of strategy used depends on the preference of the individ-
ual screening program. Although measurement of both T4 and TSH is optimal,
this approach is more expensive and so most screening programs opt for pri-
mary measurement of one or the other analyte. Both primary T4/backup TSH
and primary TSH strategies have their advantages and disadvantages, but the two
approaches appear to be equivalent in the detection of babies with permanent
forms of CH (21). Both strategies will miss the rare infant whose T4 and TSH
levels on initial screening are normal but who later develop low T4 and ele-
vated TSH concentrations (“atypical” CH or “delayed TSH rise”) (22). In both
strategies there is the possibility of human error in failing to identify affected
infants.
Newborn screening was performed initially at between three and four days
of life, and the normal values that were derived reflected this postnatal age. The
practice of early discharge from the hospital of otherwise healthy full-term infants
has resulted in a greater proportion of babies being tested before this time. Because
of the neonatal TSH surge and the dynamic changes in T4 and T3 concentrations
that occur within the first few days of life, early discharge increases the num-
ber of false-positive results, particularly in primary TSH screening programs.
Another complicating factor is the dramatically increased survival of very prema-
ture infants.

Causes of CH
Thyroid Dysgenesis
The causes of permanent nonendemic CH are listed in Table 1. Thyroid dysgenesis
is the most common cause, accounting for 85% to 90%. It is almost always
a sporadic disease. Thyroid dysgenesis may result in the complete absence of
thyroid tissue (agenesis), or it may be partial (hypoplasia); the latter is often
accompanied by a failure to descend into the neck (ectopia).
Both genetic and environmental factors have been implicated in the etiology
of thyroid dysgenesis, but the cause is unknown in most cases. Genetic abnormal-
ities in the transcription factors TTF1, TTF2, and PAX8 have been identified rarely
in isolated CH (23,24). Heterozygous deletions of TTF1 have been reported in a
few patients with CH, unexplained neonatal respiratory distress, and neurological
manifestations (25,26). Usually the CH is mild. Similarly a homozygous missense
Pediatric Thyroid Disorders 337

Table 1 Causes of Permanent Congenital Hypothyroidism


Primary hypothyroidism
Thyroid dysgenesis (agenesis/dysgenesis)
-Syndromic
-Non syndromic
Thyroid dyshormonogenesis
-Iodide concentrating defect
-Defective organification
-Thyroid peroxidase mutation
-Abnormal H2 O2 generation
-Defective thyroglobulin synthesis or transport
TSH resistance
Secondary or tertiary hypothyroidism
Isolated TSH deficiency
-TRH deficiency or resistance
-TSH␤ mutation
Multiple pituitary hormone deficiencies
-Septo-optic dysplasia
-Pituitary transcription factor mutation, e.g., PROP1, PIT1

mutation in the TTF2 gene has been associated with the syndrome of thyroid
agenesis, bifid epiglottis, cleft palate, kinky hair, and choanal atresia (27).

Inborn Errors of Thyroid Hormonogenesis


Decreased T4 synthesis due to an inborn error of thyroid hormonogenesis is
responsible for most of the remaining cases (10–15%) of permanent CH (24,28).
A number of different defects have been characterized and include (1) failure to
concentrate iodide, (2) defective organification of iodide due to an abnormality
in the thyroid peroxidase (TPO) enzyme or in the H2 O2 generating system, (3)
defective Tg synthesis or transport, and (4) abnormal iodotyrosine deiodinase
activity. A partial organification defect can also be due to mutations in the gene for
pendrin, an iodine transport protein located in the apical border of the thyrocyte
(29). Affected patients may have the classical triad of goiter, may have partial
organification defect and congenital sensorineural defect, or they may have con-
genital sensorineural deafness alone. Thyroid function is usually not affected in
the newborn period. All the inborn errors of thyroid hormonogenesis are associ-
ated with a normally placed thyroid gland of normal or increased size, and this
feature forms the basis for the clinical distinction from thyroid dysgenesis.
Unlike thyroid dysgenesis, a sporadic condition, the inborn errors of thyroid
hormonogenesis tend to have an autosomal-recessive form of inheritance, consis-
tent with a single gene mutation. It is not surprising, therefore, that a molecular
basis for many of these abnormalities has now been identified.
338 Brown

TSH Resistance
Decreased T4 synthesis resulting from resistance to the action of TSH is a rare
cause of CH. Babies with this abnormality have a normal or hypoplastic gland in
the normal location in the neck; in rare cases, no thyroid gland at all is discernible
on thyroid imaging. TSH resistance may be caused by a loss of function mutation
of the TSH receptor and is usually autosomal recessive (30,31). In other patients
with TSH resistance, a postreceptor defect has been hypothesized. Rarely, TSH
resistance may result from an inactivating mutation of the stimulatory guanine
nucleotide-binding protein (GSa) gene (Albright’s hereditary osteodystrophy).
The hypothyroidism at birth is usually mild.
Decreased TSH Synthesis or Secretion
CH resulting from TSH deficiency is only detected by newborn screening programs
that use a primary T4 strategy; the reported incidence is 1 in 50,000 to 1 in
100,000. TSH deficiency may be isolated, or, more commonly, it is associated with
other pituitary hormone deficiencies. Reported causes of isolated TSH deficiency
have included to an abnormality in TSH␤, TRH deficiency, and TRH resistance.
Congenital hypopituitarism is suggested by the presence of additional clinical
features, such as hypoglycemia, microphallus, abnormal midline facial and brain
structures, and/or nystagmus. An important cause of congenital hypopituitarism is
septo-optic dysplasia (De Morsier syndrome); abnormalities in PIT-1, PROP-1,
and TTFs important in pituitary gland development may also cause congenital
hypopituitarism.
Abnormal T4 Action
Abnormal T4 action (thyroid hormone resistance) can result from a mutation
in the T4 receptor or in a postreceptor signaling abnormality. Affected babies
usually do not present in the newborn period. A novel, recently described, cause
of abnormal T4 action is a mutation in the MCT8 thyroid hormone transporter,
which is essential for the transport of T4 into the cell (32).
Transient Congenital Hypothyroidism
The true frequency of transient CH is unknown but is probably greater than
the earlier estimate of 10% of babies detected on newborn screening in view
of the greater incidence and survival of increasingly premature infants in recent
years. The most common causes of transient neonatal hypothyroidism are iodine
deficiency or excess, maternal antithyroid medication, and maternal TSH receptor
blocking Abs (Table 2).
Transient hypothyroidism due to both iodine deficiency and iodine excess
is more common in relatively iodine-deficient areas of Europe than in North
America, an iodine-sufficient region (33,34). Premature infants are particularly
at risk of iodine-induced hypothyroidism, whether administered to the mother
during pregnancy or delivery or directly to the baby. Reported sources of iodine
Pediatric Thyroid Disorders 339

Table 2 Causes of Other Thyroid Hormone


Abnormalities in the Newborn Period

Tansient primary hypothyroidism


Iodine deficiency or excess
Maternal antithyroid drug therapy
Maternal TSH receptor–blocking antibodies
Transient secondary or tertiary hypothyroidism
Drugs (dopamine, steroids)
Maternal hyperthyroidism
Other
Prematurity
Illness
Undernutrition
Isolated hyperthyrotropinemia

have included drugs (e.g., potassium iodide, amiodarone, or radiocontrast agents)


and antiseptic solutions (e.g., povidone-iodine) used for skin cleansing or vaginal
douches.
Transient neonatal hypothyroidism may develop in babies whose mothers are
being treated with antithyroid medication for the treatment of Graves’ disease. The
fetus appears to be particularly sensitive to the effects of these drugs, even when
the dosage used in the mother is within currently recommended guidelines (35).
Babies with antithyroid drug-induced hypothyroidism characteristically develop
an enlarged thyroid gland. Replacement therapy is usually not required because
the hypothyroidism tends to be short lived.
Maternal TSH receptor blocking Abs may be transmitted to the fetus in
sufficient titer to cause transient CH. Transient CH has been reported most often
in babies born to mothers who have been treated earlier for Graves’ disease or
who have the nongoitrous form of chronic lymphocytic thyroiditis (CLT) (primary
myxedema) (36). Occasionally, these mothers are not aware that they are hypothy-
roid and the diagnosis is made in them only after CH has been recognized in
their infants. Because TSH-induced growth is blocked, these babies do not have a
goiter. Affected babies have been misdiagnosed with thyroid agenesis because the
blocking Abs inhibit TSH-stimulated radioactive iodine (RAI) uptake. In contrast
to findings on scintiscan, a normally placed thyroid gland can be visualized on
ultrasound. The hypothyroidism generally resolves in three or four months when
Ab is cleared from the neonatal circulation. Unlike babies with thyroid dysgenesis
in whom a normal cognitive outcome is found if postnatal therapy is early and
adequate, babies with maternal blocking Ab-induced hypothyroidism may have
a permanent deficit in intellectual development if feto-maternal hypothyroidism
was present in utero (10).
Transient central hypothyroidism may be due to drugs (particularly steroids
or dopamine) commonly used in the newborn intensive care unit or may be
340 Brown

observed in neonates born to mothers with Graves’ disease (see later). A low
free T4 associated with a normal TSH may also be observed in premature infants
whose hypothalamic-pituitary axis is immature and in babies who are sick or
undernourished.

Other Abnormalities of Thyroid Function Discovered on


Newborn Screening
Isolated Hyperthyrotropinemia
Isolated hyperthyrotropinemia may be detected in screening programs that use
a primary TSH method or postnatally, particularly in premature infants. While
some of these babies represent cases of “compensated” hypothyroidism, in other
instances the etiology is not clear. In one study, babies diagnosed with hyperthy-
rotropinemia in infancy had a higher serum TSH compared to control children
when reexamined in early childhood (37). Also, these infants have a higher preva-
lence of both thyroid morphological abnormalities, antithyroid antibodies (Abs)
and mutations in the thyroperoxidase and TSH receptor genes than do normal
babies (38). In other cases, the serum TSH normalized without treatment within
the first year of life (39).

Clinical Manifestations
Clinical evidence of hypothyroidism is usually difficult to appreciate in the new-
born period. Many of the classic features (large tongue, hoarse, cry, facial puffiness,
umbilical hernia, hypotonia, mottling, cold hands and feet and lethargy) are subtle
and develop only with the passage of time (Fig. 2). Nonspecific signs that sug-
gest the diagnosis of CH include prolonged, unconjugated hyperbilirubinemia;
gestation longer than 42 weeks; feeding difficulties; delayed passage of stools;
hypothermia; or respiratory distress in an infant weighing more than 2.5 kg. A
large anterior fontanelle and/or a posterior fontanelle ⬎0.5 cm is frequently present
in affected infants, but may not be appreciated.
Babies with CH are of normal size at birth. However, if diagnosis is delayed,
subsequent linear growth is impaired. The finding of palpable thyroid tissue sug-
gests that the hypothyroidism is due to an abnormality in thyroid hormonogenesis
or that it will be transient.

Laboratory Evaluation
Infants detected by newborn screening should be evaluated without delay, prefer-
ably within 24 hours. The diagnosis of primary CH is confirmed by the demonstra-
tion of a decreased concentration of free T4 and an elevated TSH level in serum.
Most infants with permanent abnormalities of thyroid function have a serum TSH
concentration ⬎40 mU/L. A bone age X-ray may be performed as a reflection of
the duration and severity of the hypothyroidism in utero. Thyroid imaging provides
information about the location and size of the thyroid gland. A radionuclide scan
Pediatric Thyroid Disorders 341

Figure 2 Baby with congenital hypothyroidism diagnosed clinically at the age of three
months. Baby was born in Puerto Rico prior to the advent of newborn screening and moved
to the United States shortly after birth. Note the dull facies and large tongue.

(either 123 I or 99 m pertechnetate) is the classical approach, but ultrasonography,


which avoids the potential risk of radiation exposure, is a suitable alternative.
Thyroid ultrasound is not as sensitive as scintiscan in identifying ectopic tissue
(40) and, unlike 123 I, cannot evaluate iodine transport defects or abnormalities in
thyroid oxidation. Nonetheless, it provides information as to whether a normally
located thyroid is present, and, therefore, whether the condition is likely to be
transient or permanent. It is also helpful in genetic counseling as thyroid dys-
genesis is almost always a sporadic condition, whereas abnormalities in thyroid
hormonogenesis are usually autosomal recessive.
TSH receptor blocking Abs should be measured in babies with a normally
located thyroid gland. Autoimmune thyroid disease in the mother or a history
of a previously affected sibling should alert the physician to this diagnosis, but
such information is not always known. A binding assay (TSH Receptor Abs,
TRAbs, or TSH Binding Inhibitory IgGs, TBII, discussed further under Graves’
342 Brown

Abnormal thyroid screen

Serum free T4, TSH, Tg,


ultrasound (BA optional)

No thyroid visible Thyroid eutopic Thyroid ectopic

Tg nl or ↑ Tg ↓ or ND TSH receptor Abs


Positive Blocking
Probably Probably Ab-
Negative
ectopic agenesis Induced
CH
Tg ND or ↓ Tg ↑ or normal
Can confirm with
thyroid uptake/scan
(optional) Tg synthetic defect Trapping defect
TSH insensitivity Organification defect
Iodine excess Maternal ATDs

Figure 3 Suggested strategy for the investigation of a child with an abnormal newborn
thyroid screen. Abbreviations: Tg, thyroglobulin; ND, not detected; nl, normal; ATDs,
antithyroid drugs; CH, congenital hypothyroidism; Ab, antibody.

disease) is appropriate for screening; bioassay can be done later, if desired, to


demonstrate the biological action of the Abs, that is, whether it is a stimulating or a
blocking antibody. In cases of TSH receptor Ab-induced CH, the blocking activity
is extremely potent (36); a weak or borderline result should cause a reconsideration
of this diagnosis. TPO Abs, although frequently detectable in babies with blocking
Ab-induced CH, are neither sensitive nor specific in predicting the presence of
transient CH.
Measurement of urinary iodine is helpful if the diagnosis of iodine-induced
hypothyroidism is suspected. The serum Tg concentration reflects the amount of
thyroid tissue and is low or undetectable in patients with agenesis. It can also be
used to distinguish a defect in Tg synthesis or secretion (low serum level) from
other causes of thyroid dyshormonogenesis in which a high value is found.
In babies in whom hypothyroxinemia unaccompanied by TSH elevation is
found, free T4 should be measured, preferably by a direct dialysis method, and the
TBG concentration should be evaluated as well. The finding of a low free T4 in
the presence of a normal TBG may suggest the diagnosis of secondary or tertiary
hypothyroidism, particularly if the patient has a microphallus, hypoglycemia,
or a midline facial abnormality. In premature, low birth weight or sick babies
in whom a low T4 and “normal” TSH are found, the free T4 when measured
by a direct dialysis method is frequently not as low as the total T4 (15). In
these infants, T4 (and/or free T4) and TSH should be repeated every one to two
Pediatric Thyroid Disorders 343

weeks until the T4 normalizes because of the rare occurrence of delayed TSH
rise (22). Thyroid function should also be monitored in other infants at risk of
a delayed TSH rise, such as severely ill babies in an intensive care setting (41)
and monozygotic twins (42) in whom fetal blood mixing may initially mask the
presence of CH. Even though in many ill babies the hypothyroidism will be
transient, treatment should be considered if values do not normalize within one
to two weeks. In any infant, if signs or symptoms suggestive of hypothyroidism
are present, thyroid function testing should be repeated because of the possibility
of delayed onset of hypothyroidism and because of rare errors in the screening
program.

Therapy
Replacement therapy with thyroxine should begin as soon as the diagnosis of
CH is confirmed (43). Parents should be counseled regarding the causes of CH,
the importance of compliance, and the excellent prognosis in most babies if
therapy is initiated early. Educational material should be provided. Treatment
need not be delayed in anticipation of performing a thyroid scan as long as this
is done within before suppression of the serum TSH. Until recently, an initial
daily thyroxine dosage of 10 to 15 ␮g/kg (equivalent to 37.5 ␮g in a full-term
infant) was recommended to normalize the T4 as soon as possible, but a slightly
higher dose (50 ␮g, equivalent to 12–17 ␮g/kg/day), which results in more rapid
normalization of the serum T4 and TSH concentrations, may be even better (44).
Babies with compensated hypothyroidism may be started on the lower dosage,
while those with severe CH (e.g., T4 ⬍ 5 ␮g/dL), such as those with thyroid
agenesis, should be started on the higher dosage. Thyroid hormone may be crushed
and administered with juice or water, but care should be taken that all the medicine
has been swallowed. Medication should be administered at least 30 minutes to an
hour before feeding and never with substances that interfere with its absorption,
such as iron, soy, or fiber. Many babies will swallow the pills whole or chew the
tablets with their gums even before they have teeth. A brand name preparation of
levothyroxine is preferred. Liquid preparations are unstable and should be used
with caution. If a dose is missed or thought to be missed, a double dose should be
given the next day.
The aims of therapy are to normalize the serum T4 and TSH as soon as
possible preferably within three days and two weeks, respectively. Subsequent
adjustments in the dosage of medication are made according to the results of
thyroid function tests and the clinical picture. Some infants develop supraphysio-
logical serum T4 values, but the serum T3 concentration usually remains normal,
most affected infants are not symptomatic, and these short-term T4 elevations have
not been reported to be associated with adverse effects on growth, bony matura-
tion, or cognitive development. Both initial dosage and timing of onset of therapy
are independent prognostic variables of cognitive development (45). Combined
therapy with T4 and T3 offers no advantage over T4 alone.
344 Brown

Current recommendations are to repeat T4 and TSH at two and four weeks
after the initiation of thyroxine treatment, every 1 to 2 months during the first
year of life, every 2 to 3 months between 1 and 3 years of age, and every 3 to 12
months thereafter until growth is complete (43). In hypothyroid babies in whom
an organic basis was not established and in whom transient disease is suspected,
a trial of replacement therapy can be initiated after the age of three years when
most thyroid hormone-dependent brain maturation has occurred. Alternatively,
recombinant hTSH may prove to be a suitable alternative to thyroid hormone,
although it is not yet FDA approved for use in children (46).
Normalization of the TSH concentration may sometimes be delayed because
of relative pituitary resistance. In such cases, characterized by a normal or
increased serum T4 and an inappropriately high TSH level, the T4 value is used
to titrate the dosage of medication, but noncompliance is the most common cause
and should be excluded.
Whether or not premature infants with hypothyroxinemia should be treated
remains controversial and prospective clinical trials are needed (47). At the very
least, every effort should be made to avoid the use of iodine and drugs that can
suppress thyroid function (e.g., dopamine, steroids) in these infants (48). The
initial finding of an 18-point increase in the Bayley Mental Development Index
score in T4-treated infants ⬍27 weeks gestation (49) was not sustained when these
children were reevaluated at 10 years of age (50).

Neonatal Hyperthyroidism
Transient Neonatal Hyperthyroidism
Unlike CH, which is usually permanent, neonatal hyperthyroidism is almost
always transient, secondary to the transplacental passage of maternal TSH
receptor-stimulating Abs. Hyperthyroidism develops only in babies born to moth-
ers with the most potent stimulatory activity in serum (51,52), corresponding to
2% to 3% of mothers with Graves’ disease, or 1 in 50,000 newborns. The incidence
is approximately four times higher than that for transient neonatal hypothyroidism
due to maternal TSH receptor-blocking Abs (36). Both TSH receptor potency, the
severity, and duration of in utero hyperthyroidism, and the dose and duration of
maternal antithyroid medication are important determinants of neonatal thyroid
status. In one study of 230 infants born to mothers with Graves’ disease, 83.5%
had normal function while 10% had transient hypothyroidism (7.8% subclini-
cal, 2.2% overt) due to increased fetal sensitivity to the maternally administered
antithyroid drugs (ATD). Transient hyperthyroidism was a much less common
sequela, occurring in 5.6%, only 2.6% being clinically symptomatic. Transient
central hypothyroidism occurred in two infants (0.9%) (53). Rarely, mothers have
mixtures of stimulating and blocking Abs in their circulation, resulting in a more
complicated clinical picture in the infant (54). It is important to appreciate that
neonatal hyperthyroidism may occur in infants born to hypothyroid mothers whose
thyroid was destroyed by prior radioablation, surgery, or destructive autoimmune
Pediatric Thyroid Disorders 345

processes since in these mothers potent stimulating Abs might still be present but
would not be apparent clinically.
Iodine-Induced Hyperthyroidism
Iodine exposure is a rare cause of transient neonatal hyperthyroidism (55).

Clinical Manifestations
Although maternal TSH receptor Ab-mediated hyperthyroidism may present in
utero, the onset is often toward the end of the first week of life. This is both due
to the clearance of maternally administered antithyroid drugs from the infant’s
circulation and due to the increased conversion of T4 to T3 after birth. The onset
of neonatal hyperthyroidism may be delayed if higher affinity blocking Abs are
also present, but this is rare.
Fetal hyperthyroidism is suspected in the presence of fetal tachycardia (pulse
⬎160/min), especially if a goiter is seen on ultrasound and if there is evidence
for intrauterine growth retardation. In the newborn infant, characteristic signs and
symptoms include tachycardia, irritability, poor weight gain, and prominent eyes.
Infants may be born prematurely. Goiter may be related to maternal antithyroid
drug treatment, to consequent neonatal hypothyroidism, or to neonatal Graves’
disease. Rarely, infants with neonatal Graves’ disease present with thrombocy-
topenia, hepatosplenomegaly, jaundice, and hypoprothrombinemia, a picture that
may be confused with congenital infections. Dysrhythmias and cardiac failure may
develop and may cause death, particularly if treatment is delayed or inadequate.
In addition to a significant mortality rate that approximates 20% in some older
series, untreated fetal and neonatal hyperthyroidism is associated with deleteri-
ous long-term consequences, including premature closure of the cranial sutures
(cranial synostosis), failure to thrive, and developmental delay (56).
The half-life of TSH receptor Abs in the blood is one to two weeks. The
duration of neonatal hyperthyroidism, a function of Ab potency and metabolic
clearance rate, is usually two to three months, but may be longer.

Laboratory Evaluation
Because of the importance of early diagnosis and treatment, fetuses and infants at
risk for neonatal hyperthyroidism should undergo both clinical and biochemical
assessment. A high index of suspicion is necessary in babies of women who have
had thyroid ablation because a high titer of TSH receptor Abs would not be evident
clinically. Similarly, women with persistently elevated TSH receptor Abs and with
a high requirement for antithyroid medication during pregnancy are at an increased
risk of having an affected child.
The diagnosis of hyperthyroidism is confirmed by the demonstration of an
increased concentration of T4, free T4, and T3 accompanied by a suppressed TSH.
Fetal ultrasonography may help to detect a goiter and to monitor fetal growth. If
necessary, blood can be obtained by fetal umbilical cord catheterization and results
346 Brown

can be compared with normal values during gestation. Demonstration of a high


titer of TSH receptor Abs in the baby or mother will confirm the etiology of the
hyperthyroidism, and in babies whose thyroid function testing is normal initially,
it indicates the degree to which the baby is at risk. In general, babies likely to
become hyperthyroid have the highest TSH receptor Ab titer, whereas, if TSH
receptor Abs are not detectable, the baby is most unlikely to become hyperthyroid
(57,58). The sensitivity of different TSH receptor Ab assays varies greatly, so
specific values that are recommended in the literature should be interpreted with
caution. Close follow-up of all babies with abnormal thyroid function tests or
detectable TSH receptor Abs is mandatory.

Therapy
Treatment of the fetus is accomplished by maternal administration of antithyroid
medication. The minimal dosage of PTU or MMI necessary to normalize the
fetal heart rate and render the mother euthyroid or slightly hyperthyroid is usu-
ally chosen. In the neonate, treatment is expectant. Either PTU (5–10 mg/kg/day)
in three divided doses or MMI (0.5–1 mg/kg/day) twice a day can be used ini-
tially. If the hyperthyroidism is severe, a strong iodine solution (Lugol’s solu-
tion or saturated solution of potassium iodine, one drop every eight hours) is
added to block the release of thyroid hormone immediately because the effect of
PTU and MMI may be delayed for several days. Therapy with both antithyroid
drug and iodine is adjusted subsequently, depending on the response. Propranolol
(2 mg/kg/day) in two or three divided doses is added if sympathetic overstimulation
is severe, particularly in the presence of pronounced tachycardia. If cardiac failure
develops, treatment with digoxin should be initiated, and propranolol should be
discontinued.
Rarely, prednisone (2 mg/kg/day) is added for immediate inhibition of thy-
roid hormone secretion and is added for decreased generation of T3 from T4
in peripheral tissues. Measurement of TSH receptor Abs in treated babies may
be helpful in predicting when antithyroid medication can be safely discontinued.
Lactating mothers on antithyroid medication can continue nursing as long as the
dosage of PTU or MMI does not exceed 400 mg or 40 mg, respectively. At higher
dosages, close supervision of the infant is advisable.

Permanent Neonatal Hyperthyroidism


Rarely, neonatal hyperthyroidism is permanent and is due to a germline mutation
in the TSH receptor, resulting in its constitutive activation (59). A gain-of-function
mutation of the TSH receptor should be suspected if persistent neonatal hyper-
thyroidism occurs in the absence of detectable TSH receptor Abs in the maternal
circulation. An autosomal dominant inheritance has been noted in many of these
infants, but other cases have been sporadic, arising from a de novo mutation.
Early recognition is important because the thyroid function of affected infants
is frequently difficult to manage medically (60). When diagnosis and therapy
Pediatric Thyroid Disorders 347

Table 3 Causes of Hypothyroidism in Childhood


and Adolescence Primary hypothyroidism

Primary hypothyroidism
Chronic lymphocytic thyroiditis
Thyroid dysgenesis or dyshormonogenesis
Iodine deficiency/goitrogenes
Drugs
Miscellaneous
-Mantle irradiation
-Cystinosis
-Histiocytosis X
-Mitochondrial disease
-Hemangiomas (“consumptive hypothyroidism”)
Secondary or tertiary hypothyroidism
Brain irradiation/surgery
Brain tumor (craniopharyngioma)
Congenital hypopituitarism
Abnormal thyroid hormone action
Thyroid hormone resistance (TR␤ mutation)
Abnormal T4 transport (MCT8 mutation)

are delayed, irreversible sequelae, such as cranial synostosis and developmental


delay, may result (61). For this reason, early, aggressive therapy with either thy-
roidectomy or radioablation has been recommended.

THYROID DISEASE IN CHILDHOOD AND ADOLESCENCE


Hypothyroidism
Chronic Lymphocytic Thyroiditis
The causes of hypothyroidism after the neonatal period are listed in Table 3.
The most frequent cause is CLT, an autoimmune disease that is closely related to
Graves’ disease. Both a goitrous (Hashimoto’s disease) and a nongoitrous (primary
myxedema) variant of thyroiditis have been distinguished. The disease has a strik-
ing predilection for females, and a family history of autoimmune thyroid disease
(both CLT and Graves’ disease) is found in 30% to 40% of patients. Adolescence
is the most common age at presentation during childhood, but it can occur at any
age after the neonatal period, even in infants as young as six months of age (62).
Patients with insulin-dependent mellitus have an increased prevalence of
CLT. CLT may also occur as part of autoimmune polyglandular syndrome (APS).
In APS-1, also called APECED (Autoimmune Polyendocrinopathy Candidiasis
Ectodermal Dystrophy) syndrome, CLT is found in 10% of patients (63); CLT
is much more common in APS types 2 (Addison’s disease, CLT with or with
T1DM) and 3 (CLT and another organ-specific autoimmune disease, not Addison’s
348 Brown

disease). Unlike APS-1, which usually presents in early childhood, APS-2 tends to
occur later in childhood or in the adult. There is an increased prevalence of CLT in
patients with Down, Turner, Klinefelter and Noonan syndromes. CLT may also be
associated with chronic urticaria and with immune complex glomerulonephritis.
Antibodies to Tg and TPO, the thyroid Abs measured in routine clinical
practice, are detectable in a majority of patients with CLT. They are useful
as markers of underlying autoimmune thyroid damage, TPO Abs being more
sensitive. TSH receptor Abs are also found in a small proportion of patients. When
stimulatory TSH receptor Abs are present they may give rise to a clinical picture
of hyperthyroidism, the coexistence of CLT and Graves’ disease being known
as “Hashitoxicosis”. TSH receptor blocking Abs, on the other hand, have been
postulated to underlie both the hypothyroidism and the absence of goiter in some
patients with primary myxedema, but are detectable in only a minority of children
(64). Occasionally, the disappearance of blocking Abs has been associated with
normalization of thyroid function in previously hypothyroid patients (65).
Goiter is present in approximately two-thirds of children with CLT. Most
patients are euthyroid. Although the most common functional disturbance is
hypothyroidism, less commonly, as noted earlier, patients may be hyperthyroid.
When present, an initial thyrotoxic phase is usually not due to true hyperthy-
roidism, but is due to the discharge of preformed T4 and T3 from the damaged
gland, similar to “silent” thyroiditis in adults.

Thyroid Dysgenesis and Inborn Errors of Thyroid Hormonogenesis


Occasionally, patients with mild thyroid dysgenesis will escape detection by new-
born screening and present later in childhood with nongoitrous hypothyroidism
or with an enlarging mass at the base of the tongue, or along the course of the
thyroglossal duct. Similarly, children with milder inborn errors of thyroid hor-
monogenesis may only be recognized later in childhood because of the detection
of a goiter.

Iodine and Other Micronutrient Deficiency: Natural Goitrogens


In areas of endemic iodine deficiency, hypothyroidism may be exacerbated by the
coincident ingestion of goitrogen-containing foods, such as cassava, soybeans,
broccoli, cabbage, sweet potatoes, and cauliflower, or by certain water pollutants.
Iodine deficiency may be endemic or can be due to dietary restriction (for multiple
food allergies) or the result of dietary fadism. Iodine deficiency may also be exac-
erbated by lack of selenium, a component of the selenocysteine thyroid hormone
deiodinases.

Drugs
A number of drugs used in childhood may affect thyroid function. These include
antithyroid medication, certain anticonvulsants, lithium, aminosalicylic acid,
amiodarone, and aminoglutethimide.
Pediatric Thyroid Disorders 349

Miscellaneous Causes of Acquired Hypothyroidism


Mantle irradiation for Hodgkin’s disease or lymphoma may result in hypothy-
roidism. The thyroid gland may be involved in generalized infiltrative (cystinosis),
granulomatous (histiocytosis X), or infectious disease processes that are of suf-
ficient severity to result in a disturbance in thyroid function. Hypothyroidism
may also occur in patients with mitochondrial disease. Rarely in infancy a large
hemangioma with high D3 activity can be associated with rapid inactivation of T4
and severe hypothyroidism (66). Extremely high T4 replacement doses may be
required in this syndrome.

Secondary or Tertiary Hypothyroidism


Like primary hypothyroidism, milder forms of congenital hypopituitarism often
present after the newborn period. Acquired damage to the pituitary or hypothala-
mus is most commonly a consequence of brain tumors (particularly craniopharyn-
gioma) or their therapy (surgery or irradiation). Other causes include granuloma-
tous disease, infection, and trauma.

Thyroid Hormone Resistance


Children with thyroid hormone resistance usually come to attention when thyroid
function tests are performed because of poor growth, hyperactivity, a learning
disability, or other nonspecific signs or symptoms. A small goiter may be present.
The presentation is highly variable, and some individuals may be completely
asymptomatic. Thyroid hormone resistance is most frequently caused by a point
mutation in the hinge region or ligand-binding domain of the thyroid hormone
receptor (TR) ␤ gene (67). As a consequence, there is a dramatic reduction in
T3 binding. Less frequently, it results from impaired interaction with one of the
cofactors involved in the mediation of thyroid hormone action. Because these
mutant TRs interfere with the function of the normal TRs, a dominant pattern of
inheritance is seen.
Rarely, thyroid hormone resistance may be found in patients with cystinosis
(68).

Clinical Manifestations
The onset of hypothyroidism in childhood is insidious. Affected children are usu-
ally recognized either because of the detection of a goiter on routine examination
or because of poor growth, sometimes for several years prior to diagnosis (Fig. 4).
Because linear growth tends to be more affected than weight, patients are relatively
overweight for their height, although they are rarely significantly obese. Dental
and skeletal maturation are delayed, the latter often significantly. Patients with
secondary or tertiary hypothyroidism tend to be even less symptomatic than those
with primary hypothyroidism.
350 Brown

A B

Dx
F

Figure 4 Adolescent female with severe primary hypothyroidism who presented with sec-
ondary amenorrhea at the age of 16 years. Note the progressive change in facial appearance
over time (A) and the decreased growth velocity (B) prior to diagnosis. With treatment,
growth rate increased but weight decreased. Heights of mother (M) and father (F) are indi-
cated at the right. Despite adequate thyroid hormone replacement, adult height was only
47, well below her genetic potential.

The classical clinical manifestations of hypothyroidism (lethargy, cold intol-


erance, constipation, dry skin, or hair texture) can be elicited on careful evaluation;
however, they are often not the presenting complaints. School performance is not
usually affected, in contrast to the severe irreversible sequelae that occur in inad-
equately treated babies with CH. A delayed relaxation time of the deep tendon
reflexes may be appreciated in more severe cases.
The typical thyroid gland in CLT is diffusely enlarged and has a rubbery
consistency. Occasionally asymmetric enlargement occurs and must be distin-
guished from thyroid neoplasia. An enlarged pyramidal lobe or Delphian lymph
node superior to the isthmus can be found and may be confused with a thyroid nod-
ule. The absence of goiter suggests the diagnosis of primary myxedema, thyroid
dysgenesis, or secondary/tertiary hypothyroidism.
Puberty tends to be delayed in hypothyroid children, although sexual pre-
cocity has been described in long-standing, severe hypothyroidism. Females may
menstruate but commonly have breast development with little sexual hair. Ovarian
Pediatric Thyroid Disorders 351

cysts may be demonstrated on ultrasonography due to follicle stimulating hor-


mone secretion. Galactorrhea due to hyperprolactinemia may occur occasionally.
In boys, isolated testicular enlargement may be found. There is an increased inci-
dence of slipped femoral capital epiphyses in hypothyroid children. In patients
with severe hypothyroidism of long-standing duration, the sella turcica may be
enlarged due to thyrotrope hyperplasia.

Laboratory Evaluation
Measurement of TSH is the best initial screening test for the presence of primary
hypothyroidism. If the TSH is elevated, measurement of free T4 will distin-
guish whether the child has compensated (normal free T4) or overt (low free T4)
hypothyroidism.
Measurement of TSH is not helpful in secondary or tertiary hypothyroidism.
Hypothyroidism in these cases is demonstrated by the presence of a low free T4
with an inappropriately low or normal TSH. Occasionally, mild TSH elevation
is seen in individuals with hypothalamic hypothyroidism, a consequence of the
secretion of a TSH molecule with impaired bioactivity but normal immunoreac-
tivity. Thyroid hormone resistance is characterized by elevated levels of free T4
and T3 and an inappropriately normal or elevated TSH concentration.
CLT is diagnosed by elevated titers of Tg and/or TPO Abs. Ancillary inves-
tigations (thyroid ultrasonography and/or thyroid scintigraphy) may be performed
if thyroid Ab tests are negative or if a nodule is palpable, but are rarely necessary.
If thyroid Ab tests are negative and no goiter is present, thyroid ultrasonogra-
phy and/or scan identify the presence and location of thyroid tissue and thereby
distinguish primary myxedema from thyroid dysgenesis.

Therapy
In contrast to CH, rapid replacement is not essential in the older child. This is
particularly true in children with long-standing, severe thyroid underactivity in
whom rapid normalization may result in unwanted side effects (deterioration in
school performance, short attention span, hyperactivity, insomnia, and behavior
difficulties). Replacement doses should be increased very slowly as tolerated over
several weeks to months. Severely hypothyroid children should also be observed
closely for complaints of severe headaches when therapy is initiated because of
the rare development of pseudotumor cerebri. In contrast, full replacement can be
initiated at once without much risk of adverse consequences in children with mild
hypothyroidism.
The typical replacement dose of thyroxine in childhood is approximately 4
to 6 ␮g/kg for children 1 to 5 years of age, 3 to 4 ␮g/kg for those 6 to 10 years
and 2 to 3 ␮g/kg for those 11 years of age and older, but treatment should be
individualized. In patients with a goiter, a somewhat higher thyroxine dosage is
used in order to keep the TSH in the low normal range (0.3–1 mU/L) and thereby
352 Brown

minimize its goitrogenic effect. Whether and how patients with thyroid hormone
resistance should be treated is controversial.
After the child has received the recommended dosage for at least six to
eight weeks, T4 and TSH should be measured. Once a euthyroid state has been
achieved, patients should be monitored every 6 to 12 months. Some children with
severe, long-standing hypothyroidism at diagnosis may not achieve their genetic
potential for height even with optimal therapy, emphasizing the importance of
early diagnosis and treatment. Treatment is usually continued indefinitely.
Thyroid suppression in children with a euthyroid goiter causes a small
reduction in the size of the gland that is detectable on ultrasonography (69) but is
not appreciable clinically (70). Whether this small effect is sufficient to warrant
lifetime therapy is not clear. Similarly, whether or not children with subclinical
hypothyroidism should be treated is controversial. Some physicians treat all such
patients while others choose to reassess thyroid function in three to six months
before initiating therapy because thyroid function will normalize spontaneously
in approximately 30% to 50% of children and adolescents (71,72). Treatment has
been advocated both for symptom relief and because of the risk of progression to
overt hypothyroidism, a risk seen particularly in older individuals with positive
thyroid Abs. An expert U.S. panel recently recommended observation without
treatment of adult patients whose TSH level was ⬍10 mU/L regardless of Ab
titer (73). When signs and symptoms suggestive of hypothyroidism are present, a
trial of thyroxine therapy can be tried. Whether or not therapy is initiated, regular
follow-up of thyroid function is important.

Painful Thyroid
Painful thyroid enlargement is rare in pediatrics and suggests the probability
of either acute (suppurative) or subacute thyroiditis. Occasionally, CLT may be
associated with intermittent pain and be confused with these disorders.

Acute Suppurative Thyroiditis


In acute thyroiditis, the thyroid is very tender and the child appears toxic from the
infection, but is not thyrotoxic. Serial thyroid ultrasound examinations are very
important to perform to detect abscess formation early, so that proper incision
and drainage can be performed. Intravenous antibiotics and fluids are essential.
Recurrent attacks and involvement of the left lobe suggest a pyriform sinus fistula
between the oropharynx and the thyroid as the route of infection (74). In the latter
case, surgical extirpation of the pyriform sinus will frequently prevent further
attacks. No residual thyroid disease is expected.

Subacute Thyroiditis
Subacute thyroiditis is much less common in children and adolescents than in older
individuals. The disease presumably results from a viral infection of the thyroid
gland that typically is tender. During the early phase of the disease, the child may
Pediatric Thyroid Disorders 353

Table 4 Causes of Hyperthyroidism in Childhood


and Adolescence

Hyperthyroidism
Graves’ disease
Functioning thyroid adenoma
Gain of function mutation in TSH receptor
McCune Albright syndrome
Iodine-induced hyperthyroidism
Miscellaneous
-TSH secreting adenoma
-Hydatidiform mole
Thyrotoxicosis without hyperthyroidism
Toxic phase of chronic lymphocytic thyroiditis
Subacute thyroiditis
Thyrotoxicosis factitia

have mild thyrotoxicosis, the result of the release of preformed T4 and T3 into the
circulation. Either negative or low titers of thyroid Abs may be found, and, unlike
in Graves’ disease, RAI is low or absent. The erythrocyte sedimentation rate
is usually markedly elevated, but may be normal to slightly elevated whenever
the thyroid is not tender. Mild thyrotoxicosis is controlled with low-dose beta-
adrenergic blockade. The clinical course of subacute thyroiditis is variable, but
often progresses through three phases (thyrotoxicosis, mild hypothyroidism, and
euthyroid goiter) before the patient finally recovers with completely normal thyroid
function. The transient phase of hypothyroidism during recovery will vary in
length and severity that usually does not require a course of thyroxine therapy.
Full recovery is expected; late recurrences may occur, but are rare.

Hyperthyroidism
Graves’ Disease
Graves’ disease is by far the most common cause of hyperthyroidism in childhood
and adolescence, responsible for ⬎95% of cases (Table 4). It is an autoimmune
disorder that, like CLT, occurs in a genetic predisposed population. Graves’ disease
is caused by TSH receptor Abs that mimic the action of TSH, causing increased
thyroid hormonogenesis and growth (75). There is a strong female predisposition.
As in adults, there is an increased frequency of Graves’ disease in children with
other autoimmune diseases, such as type 1 diabetes mellitus, Addison’s disease,
and myasthenia gravis, as well as in patients with Down syndrome.

Rarer Causes of Hyperthyroidism


Hyperthyroidism may be caused by a functioning thyroid adenoma, by constitutive
activation of the TSH receptor, or may be part of the McCune–Albright syndrome.
354 Brown

Hyperthyroidism due to a TSH-secreting pituitary adenoma or a hydatidiform mole


(76) is exceedingly rare in childhood. In the rare patient with thyroid autonomy
(e.g., multinodular goiter), hyperthyroidism may be induced by iodides.
Thyrotoxicosis (elevated thyroid hormone levels with or without increased
thyroid hormonogenesis), but not hyperthyroidism (increased thyroid hormono-
genesis), may also be seen in patients with thyroiditis or after thyroid hormone
ingestion (thyrotoxicosis factitia), for example, in adolescents who are trying to
lose weight or as a suicide gesture. Accidental thyroid hormone poisoning has
also been described in infants and toddlers (77).

Clinical Manifestations
All but a few children with Graves’ disease present with some degree of thyroid
enlargement and most have symptoms and signs of excessive thyroid activity,
such as tremors, inability to fall asleep, weight loss despite an increased appetite,
proximal muscle weakness, heat intolerance, headache, and tachycardia. Often
the onset is insidious. Shortened attention span and emotional lability may lead
to severe behavioral and school difficulties. Occasionally, the diagnosis is made
accidentally during investigation of unexplained tachycardia or a heart murmur.
Acceleration in linear growth may occur, often accompanied by advancement in
skeletal maturation, but adult height is not affected. In the adolescent child, puberty
may be delayed. If menarche has occurred, secondary amenorrhea is common. If
sleep is disturbed, the patient may complain of fatigue.
Physical examination reveals a diffusely enlarged, soft, or “fleshy” thyroid
gland; smooth skin and fine hair texture; excessive activity; and a fine tremor of
the tongue and fingers. A thyroid bruit may be audible. The finding of a thyroid
nodule suggests the possibility of a toxic adenoma. The hands are often warm
and moist. Tachycardia, a wide pulse pressure, and a hyperactive precordium are
common. Café-au-lait spots, particularly in association with precocious puberty,
on the other hand, suggest a possible diagnosis of McCune–Albright syndrome
rather than Graves’ disease. If a goiter is absent, thyrotoxicosis factitia should be
considered. Severe ophthalmopathy is considerably less common in children than
in adults, although a stare and mild proptosis are frequently observed.

Laboratory Evaluation
The clinical diagnosis of hyperthyroidism is confirmed by the finding of increased
concentrations of circulating free T4 and total T3 associated with a suppressed
TSH. If the TSH is inappropriately “normal,” thyroid hormone resistance or a
TSH-secreting adenoma should be considered. An elevated total T4 associated
with a normal free component would suggest an abnormality of TBG (either
familial or acquired) or familial dysalbuminemic hyperthyroxinemia (78). An
important acquired cause of TBG excess that is often not considered in pediatric
patients is elevated estrogen, for example, secondary to oral contraceptives or
Pediatric Thyroid Disorders 355

pregnancy. Thyrotoxicosis factitia can be distinguished from hyperthyroidism by


the demonstration of a low serum Tg (79).
If the diagnosis of Graves’ disease is unclear, TSH receptor Abs should
be evaluated. Current commercially available radioreceptor or enzyme-linked
immunosorbent assay kits that measure the binding of Abs to the TSH recep-
tor (called TSH Receptor Abs, TRAbs or TSH Binding Inhibitory IgGs, TBII)
are highly sensitive and specific for Graves’ disease in children as in adults, and
are technically simple, rapid, and reproducible (80–84). Measurement of thyroid-
stimulating activity by bioassay (Thyroid-Stimulating IgGs, TSI), although the-
oretically preferable and sensitive when performed in a research setting (82), is
more expensive and technically demanding and its results tend to be more vari-
able, depending on the sensitivity of the assay employed by the individual clinical
laboratory. Measurement of TSH receptor Abs may be particularly useful in dis-
tinguishing the toxic phase of CLT and subacute thyroiditis (TSH receptor Ab
negative) from patients with Graves’ disease. Tg Abs and/or TPO Abs are also
frequently found in children and adolescents with Graves’ disease but their mea-
surement is less sensitive and less specific than TSH receptor Abs measurement
(82). RAI uptake and scan can also be used to distinguish Graves’ disease from
other causes of thyrotoxicosis (e.g., the thyrotoxic phase of either CLT or suba-
cute thyroiditis, thyrotoxicosis factitia, or a functioning thyroid nodule), but this
procedure is much more expensive and involves exposure to radioactivity. In prac-
tice, scintigraphy is necessary to confirm the diagnosis of Graves’ disease only
in atypical cases, particularly if measurement of TSH receptor Abs is negative or
borderline.

Therapy
In the absence of specific therapy for the immunological abnormality, treatment
of hyperthyroidism secondary to Graves’ disease is aimed at preventing the thy-
roid gland from responding to the stimulation by TSH receptor Abs. The choice
of which of the three therapeutic options (antithyroid drugs, radioactive iodine,
or surgery) to use should be individualized and discussed with the patient and
his/her family. Each approach has its advantages and disadvantages with respect
to efficacy, short- and long-term complications, time required to control the hyper-
thyroidism, and the requirement for compliance. Medical therapy with one of the
thionamide derivatives (propylthiouracil, PTU; methimazole, MMI; or carbima-
zole, converted to MMI) is the initial choice of most pediatricians, although RAI
is gaining increasing acceptance.

Medical Therapy
The traditional first-line approach to therapy is pharmacological blockade of thy-
roid hormone synthesis. It is hoped that the immunological disease will remit
spontaneously and that permanent destruction of the thyroid can be avoided. An
additional use of medical therapy in younger children is to postpone definitive
356 Brown

therapy with RAI to a later age when the theoretic risks of irradiation are not as
great.
Of the ATDs, MMI is preferred by many pediatric endocrinologists because,
for an equivalent dose, it requires taking fewer tablets and has a longer duration of
action, an advantage in noncompliant adolescents. In addition, MMI is associated
with a more rapid resolution of the hyperthyroidism and has a better safety profile
(see later). On the other hand, PTU, but not MMI, inhibits the conversion of T4
to the more active isomer T3, a potential advantage if the thyrotoxicosis is severe.
The usual initial dosage of MMI is 0.5 to 1 mg/kg/day (given once or twice a
day) and that of PTU is 5 to 10 mg/kg/day given three times daily. In adults a
low initial MMI dose (15 mg daily) is almost as effective as a high dose (30 mg
daily) in normalizing thyroid function tests within three to six months (85) but
whether or not a similar approach would be effective in children who tend to have
more severe, persistent disease has not been evaluated. In view of the apparent
relationship between dose of MMI and serious side effects (86), use of the smallest
effective dose necessary to control the hyperthyroidism would appear prudent. A
long-acting beta-adrenergic blocker (e.g., metoprolol or atenolol) can be added to
control adrenergic symptoms and the cardiovascular overactivity until a euthyroid
state is obtained.
The serum concentrations of T4 and T3 normalize in most patients in three
to six weeks, but the TSH concentration may not return to normal for several
months. Therefore, measurement of TSH is useful as a guide to therapy only after
it has normalized, but not initially. Once the T4 and T3 concentrations have fallen
by ⬎50% and/or normalized, one can either decrease the dose of thionamide
drug by 30% to 50% or, alternatively, wait until the TSH begins to rise and
add a supplementary dose of thyroxine in a block-replace regimen. Advocates
of the block-replace regimen cite the fewer visits, but a larger MMI dose is
required, perhaps exposing the patient to a greater risk of side effects. Initial
studies suggesting that combined therapy might be associated with an improved
rate of remission have not been confirmed.
As long as patients are compliant, hyperthyroidism is readily controlled with
ATDs in 90% of individuals. Thus, the major difficulty relates to the persistence
of the disease in pediatric patients as compared with their adult counterparts.
Unlike most adults in whom TSH Receptor Abs disappear from the circulation
within six months of treatment initiation, in children TSH receptor Abs remain
elevated in ⬎80% of children and adolescents with Graves’ disease even after
one to two years of treatment (83). The median time to remission is three to four
years (87,88) and in one study only 25% of pediatric patients remitted after two
years, with an additional 25% remitting every two years for up to six years of ATD
therapy (89). Prepubertal children, especially those younger than 4 years, appear
to have particularly severe and persistent disease (90,91). Thus, treatment for a
fixed duration of time, for example, one to two years, as recommended in adults is
likely to result in relapse in many children, and therapy needs to be individualized.
However, in patients who are compliant and in whom the hyperthyroidism can
Pediatric Thyroid Disorders 357

be controlled readily and tapered to a relatively small dose of ATD, medical


therapy is a reasonable approach and is well tolerated. Lack of eye signs, small
goiter, and a small drug requirement suggest that drug therapy can be tapered and
withdrawn (86). Persistence of TSH Receptor Abs, on the other hand, predicts a
high likelihood of relapse (92). Currently, it is not possible to predict at diagnosis
which patient is likely to undergo a sustained remission. Initial disease severity and
TSH receptor Ab titer have been of limited prognostic assistance, being predictive
of failure to remit in some (93) but not in most other studies (94,95). Usually,
drug withdrawal is done during a school vacation so as to minimize any potential
interference with school performance. After cessation of ATD therapy, relapses
usually occur within six months.
Toxic drug reactions (erythematous rashes, urticaria, arthralgias, and tran-
sient granulocytopenia—⬍1500 granulocytes/mm3 ) may be slightly more com-
mon in children than in adults but the precise frequency varies greatly, depending
on the series. These side effects, considered to be allergic reactions, are almost
always seen in the first three months of therapy and in patients treated with higher
doses of medication. Usually they subside spontaneously or with substitution of an
alternative thionamide drug. Urticaria can be treated with antihistamine therapy,
and thus may not warrant discontinuation of the drug. An estimate of transient
abnormalities in liver enzymes and mild leukopenia in 20% to 30% of patients
was quoted in one recent review (96). In contrast, in a review of 651 children from
10 centers treated with ATD, the incidence of granulocytopenia was 5.0% and the
mean incidence of abnormal liver function was 1.9% (97). A transient increase
in liver enzymes, with levels less than three times the upper limit of normal, may
occur with PTU therapy (98). Routine monitoring of the white blood cell count and
liver enzymes is not usually recommended because of lack of cost-effectiveness,
although some authors favor initial evaluation of these parameters prior to starting
therapy.
The overall frequency of agranulocytosis (usually defined as a granulocyte
count ⬍500 cells/␮L) has been estimated to be 0.1% to 0.5%. There is some
evidence that this life-threatening complication is more common in the first three
months of therapy, in adults over the age of 40 years, and in patients given
larger doses of MMI (⬎40 mg/day) (86). Hepatitis, on the other hand, has been
reported to occur almost exclusively with PTU and may be more common in
childhood, whereas a cholestatic reaction is typically seen with MMI. Other rarer
side effects of ATDs include a lupus-like syndrome, polyarteritis and antineutrophil
cytoplasmic antibodies, positive vasculitis, thrombocytopenia, aplastic anemia,
and nephrotic syndrome. Like hepatitis, the latter very rare side effects are more
common in patients treated with PTU. It is important to caution all patients to stop
their medication immediately and consult their physician should they develop
unexplained fever, sore throat, gingival sores, or jaundice.
Approximately 10% of children treated medically will develop long-term
hypothyroidism later in life, a consequence of coincident cell- and cytokine-
mediated destruction and/or the development of TSH receptor blocking Abs.
358 Brown

Radioactive Iodine
Definitive therapy with radioiodine ablation is usually reserved for patients who
have failed to remit or who relapse on drug therapy, developed a toxic drug reac-
tion, are noncompliant, or choose this modality. RAI is being favored increasingly
in some centers, even as the initial approach to therapy, particularly in older adoles-
cents (99). The advantages are the relative ease of administration, the reduced need
for medical follow-up, and the lack of demonstrable long-term adverse effects.
On the other hand, as the goal of therapy is thyroid ablation, one is substituting
daily medication with thyroxine for MMI, and continued medical follow-up is
nonetheless necessary.
RAI therapy in younger children, particularly in those younger than 10
years, is controversial because of the increased susceptibility of the thyroid gland
in the young to the proliferative effects of ionizing radiation. For this reason, an
ablative dose of RAI is preferred to eliminate the possibility of neoplasia in thyroid
remnant tissue. Some advocates employ a fixed radioiodine dose while others use
a formula that considers both the approximate thyroid weight and the 24-hour RAI
uptake. Pretreatment with ATDs prior to RAI therapy is not necessary unless the
hyperthyroidism is severe. Thyroid hormone concentrations may rise transiently
4 to 10 days after RAI administration owing to the release of preformed hormone
from the damaged gland. Beta-blockers may be useful to prevent worsening of
symptoms. Other acute complications of RAI therapy (nausea, significant neck
pain or swelling) are rare. One usually sees a therapeutic effect within six weeks
to three months.
Worsening of ophthalmopathy, described in adults after RAI, does not appear
to be common in childhood, but if significant ophthalmopathy is present, RAI ther-
apy should be used with caution and treatment with corticosteroids for six to eight
weeks after RAI administration may be wise. Alternatively, another permanent
treatment modality (surgery) should be considered.
In approximately 1000 children with Graves’ disease treated with RAI and
followed for ⬍5 to ⬎20 years to date, there did not appear to be any increased rate
of congenital anomalies in offspring, leukemia or thyroid cancer (96). However,
in most of these series, only a few were younger than or equal to 11 years and
even fewer were younger than 5 years, the population of patients who are most
at risk (100). Furthermore, given the rarity of thyroid cancer in childhood (1 in a
million) and even in the adult (1 in 100,000) as well as the long latency period,
small increases would not be detectable in such a small series and much more data
are needed.

Surgery
Surgery, the oldest form of therapy, is performed less frequently now than in
the past. An advantage of surgery is the rapid resolution of the hyperthyroidism.
Surgery is appropriate for patients who have failed medical management, those
in whom rapid control of the hyperthyroidism is important, those who have a
Pediatric Thyroid Disorders 359

markedly enlarged thyroid (⬎60–80 g), those who refuse RAI, and for the rare
patient with significant eye disease in whom RAI therapy is contraindicated.
Near-total thyroidectomy is preferred to minimize the risk of recurrence.
Pretreatment, usually with antithyroid medication for four to six weeks until the
hyperthyroidism is controlled followed by iodide (Lugol’s solution, 5–10 drops
daily) for one to two weeks to decrease the vascularity of the gland, is usually
recommended, although successful surgery has been reported after pretreatment
with ␤-adrenergic antagonist drug alone or in combination with iodide for only
10 to 14 days. Surgery is associated with a higher morbidity than the other thera-
peutic modalities, greatly limiting its popularity. When the results of six separate
studies involving more than 2000 children treated with surgery were pooled, the
most common complication (aside from temporary pain and discomfort, present
in all patients) was transient hypocalcemia (10%). Keloid formation occurred
in 2.8% of patients. Other less common side effects were recurrent laryngeal
nerve paralysis (2%), permanent hypoparathyroidism (2%), and, very rarely, death
(0.08%) (96). However, in a recent review of 82 children and adolescents from
one institution, no instances of either recurrent laryngeal paralysis or permanent
hypoparathyroidism were recorded and no patients died (101). Thus, when an
experienced thyroid surgeon is available and modern methods of anesthesia and
pain control are used, this therapeutic option is a safe and effective alternative.
Unfortunately, with the increased use of RAI, there has been a reduction in the
number of experienced surgeons.
After both medical and surgical thyroid ablation, most patients become
hypothyroid and require lifelong thyroid replacement therapy. On the other hand,
if therapy is inadequate, hyperthyroidism may recur.

Treatment of Other Causes of Thyrotoxicosis


Usually, treatment of other causes of thyrotoxicosis, for example, thyroiditis or
thyroid hormone ingestion, is not necessary as the signs and symptoms are self-
limited and well tolerated. Rarely, ingestion of massive doses of thyroid hormone
has resulted in severe thyrotoxicosis. In these cases, beta-blockade and iopanoic
acid (to block T4 to T3 conversion) may be useful (77).

THYROID NODULES
Thyroid carcinoma is rare in children, with an incidence of 0.5 to 1 case per
million children per year (102). In the older literature, children and adolescents
were reported to have an increased risk of malignancy (30%) in thyroid but other
studies suggest that the risk of malignancy is the same as it is in adults, 5% to 10%.
Follicular adenomas and colloid cysts account for the majority of benign thyroid
nodules. Other causes of benign nodular enlargement include CLT and embryolog-
ical defects, such as intrathyroidal duct cysts or unilateral thyroid agenesis (usually
left-sided), which can mimic a nodule. Nonthyroidal masses include teratomas,
360 Brown

branchial cleft and thyroglossal duct cysts, hemangiomas, lymphangiomas, and


neurofibromas.
Papillary thyroid carcinoma is the most common form of cancer, but other
histological types found in the adult, such as follicular carcinoma and, less often,
the oxyphil (Hurthle cell) variant, may also occur. Patients with Cowden syn-
drome, and Bannayan–Riley–Ruvalcaba syndrome are at increased risk of thyroid
carcinoma. Children whose thyroid has been exposed to therapeutic irradiation
to the head and neck area comprise a particularly high-risk group. Although
most malignancies of the thyroid are carcinomas, other malignant tumors, such as
lymphoma and sarcoma, may rarely occur. Anaplastic thyroid carcinoma is not
seen during the first 2 decades of life. Medullary thyroid carcinoma (MTC), a rare
carcinoma of the calcitonin producing or C-cells of the thyroid, is usually seen as
part of the multiple endocrine neoplasia syndromes (MEN-2a and MEN-2b).

Clinical Evaluation
A high index of suspicion is necessary if the nodule is painless; of firm or hard
consistency; or if it is fixed to surrounding tissues, especially if it has undergone
rapid growth; or if there is cervical adenopathy, hoarseness, or dysphagia. Occa-
sionally, unexplained, persistent cervical adenopathy can be the first clinically
evident manifestation of thyroid cancer. MTC should be considered if there is a
family history of thyroid cancer, pheochromocytoma or unexplained death on the
operating table, and/or if the child has findings suggestive of MEN-2b (multiple
mucosal neuromata and a marfanoid habitus).

Laboratory Evaluation
The initial investigation includes evaluation of thyroid function with a serum
TSH. A suppressed serum TSH concentration, with or without an elevation in the
circulating T4 and/or T3 levels, suggests the possibility of a functioning nodule.
Positive Ab, although indicating the presence of underlying CLT, does not exclude
the possibility of coexistent thyroid cancer. Serum calcitonin should be measured
if MTC is a concern but its routine measurement in the evaluation of thyroid
nodules is controversial. Genetic screening for a mutation of the Rearranged
during Transfection (RET) proto-oncogene should be performed if MEN-2a or
-2b is suspected (103).

Other Screening Tools


Ultrasound examination has replaced thyroid scintiscan as the preferred imaging
procedure to confirm and evaluate the morphological characteristics of a thy-
roid nodule. Although nodules that are cystic or homogeneously hyperechoic are
reputed to carry a lower risk of malignancy, ultimately there are no sonographic
findings that reliably predict the likelihood of malignancy. Therefore, a biopsy
is indicated for all thyroid nodules ≥1 cm in diameter. Fine-needle aspiration
Pediatric Thyroid Disorders 361

biopsy (FNAB), popular in the investigation of thyroid carcinoma in adults, has


gained increasing acceptance in the pediatric population, particularly in older
children. Guidelines developed for the evaluation of nodules in adults should be
followed (104). Ultrasound guidance improves the diagnostic accuracy and safety
of this procedure. If the child is very young or very anxious, sedation and/or open
excisional biopsy are suitable alternatives. Ultrasonography, FNAB, and cytology
should be performed by individuals experienced in these procedures. Due to the
relative rarity of thyroid nodules in children, this often means referral to adult
thyroidology centers.

Therapy
Surgery
Therapy of differentiated thyroid cancer is similar in childhood as in the adult,
although controversies exist, particularly in children younger than 10 years who
tend to have more persistent disease and who are at a higher risk of recurrence
(105–107). Total or near-total thyroidectomy with preservation of the parathyroid
glands and recurrent laryngeal nerves is the optimal therapy for malignant thyroid
tumors, as it facilitates RAI ablation and subsequent monitoring for recurrence
and disease progression. However, one can reserve initial bilateral surgery for
patients at high risk for malignancy, such as those whose cytology predicts a
⬎50% likelihood of differentiated thyroid cancer or who have bilateral nodules
with abnormal cytology. For other patients, thyroid lobectomy can be performed
initially, followed by completion thyroidectomy only if lobectomy confirms the
diagnosis of cancer.
Radioactive Iodine
Even after total thyroidectomy, RAI uptake usually persists in the thyroid bed
as a result of residual normal thyroid tissue. Ablation of this thyroid remnant
with RAI has been shown to lower recurrence rates and, in some series, to reduce
cancer mortality. Similar to completion thyroidectomy, RAI also facilitates disease
surveillance by increasing the specificity of Tg measurements and the sensitivity of
diagnostic whole body scans. It is important to note that for any given administered
dose, the absorbed radiation dose to normal tissues will be higher in young children
secondary to their smaller organ volumes and increased cross-radiation due to
the shorter distances between organs. Formulae for the estimation of relative
pediatric doses should be consulted. In patients with diffuse pulmonary metastases,
pulmonary fibrosis is another potential consequence of RAI.
The efficacy of radiation therapy is enhanced by clinical interventions that
increase thyroidal iodine uptake, such as withdrawal from thyroid hormone therapy
to produce an elevated circulating TSH concentration and a low-iodine diet for
one to two weeks before the procedure. The use of rhTSH to stimulate radioiodine
uptake for ablation is FDA approved in adults. Prepubertal children are more likely
to experience nausea and vomiting with 131 I therapy so antiemetic medications
362 Brown

should be available. After RAI therapy, the dosage of thyroxine replacement is


adjusted to keep the serum TSH concentration suppressed (between 0.05 and
0.1 mU/L in sensitive assays). Measurement of serum Tg, a thyroid follicular
cell-specific protein, is used to detect evidence of metastatic disease.

Follow-up Treatment and Surveillance


After RAI therapy, the dosage of thyroxine is adjusted to keep the serum TSH
concentration suppressed. Measurement of Tg on suppression and after thyroid
hormone withdrawal or rhTSH stimulation is used to detect recurrent disease. The
presence of Tg Abs, present in 15% to 30% of patients, interferes with Tg assess-
ment, so serum should be screened for Abs whenever Tg is measured. Surveillance
should also include annual neck imaging with ultrasound since the majority of
thyroid cancer recurrences are local (cervical or mediastinal lymph nodes).
Children with thyroid nodules ⬍1 cm or with benign cytology are followed
by serial ultrasound every 6 to 12 months, with ultrasound-guided FNAB if there
is any evidence of growth.

Prognosis
In comparison with adults, pediatric thyroid cancers are characterized by high rates
of regional lymph node involvement and distant metastases, especially pulmonary
micrometastases. Rates of recurrence are also higher in children, particularly those
younger than 10 years. Despite this, long-term survival is common. Overall, there
appears to be a modest decline in life expectancy, but accurate data are limited by
both the rarity of the disease in the pediatric population and the longer duration of
follow-up that is necessary. Careful lifetime follow-up is mandatory.

Medullary Thyroid Carcinoma


The clinical presentation, diagnostic evaluation, and management of children and
adolescents with MTC are the same as described for adults. Where there is a
family history of familial MTC or MEN-2a or -2b syndromes, or a positive test for
a mutation in the RET proto-oncogene in a proband, every member of the family
is tested for the RET proto-oncogene mutation, including infants (103). Genetic
screening permits the identification of affected individuals during the preclinical
stage of the disease when C-cell hyperplasia without macroscopic cancer is seen,
offering the best opportunity for cure. Because different mutations are associated
with varying degrees of aggressiveness, optimal timing of prophylactic thyroidec-
tomy varies. In MEN-2b, MTC has been described in infancy, so early detection
and surgery is necessary in affected individuals.

REFERENCES
1. De Felice M, Di Lauro R. Thyroid development and its disorders: Genetics and
molecular mechanisms. Endocr Rev 2004; 5:722–746.
Pediatric Thyroid Disorders 363

2. Fagman H, Grande M, Edsbagge J, et al. Expression of classical cadherins in thyroid


development: Maintenance of an epithelial phenotype throughout organogenesis.
Endocrinology 2003; 144(8):3618–3624.
3. Fisher DA, Klein AH. Thyroid development and disorders of thyroid function in the
newborn. N Engl J Med 1981; 304(12):702–712.
4. Morreale de Escobar G, Obregon MJ, Escobar del Rey F. Is neuropsychological
development related to maternal hypothyroidism or to maternal hypothyroxinemia?
J Clin Endocrinol Metab 2000; 85(11):3975–3987.
5. Burrow GN, Fisher DA, Larsen PR. Maternal and fetal thyroid function. N Engl J
Med 1994; 331(16):1072–1078.
6. Brown RS, Shalhoub V, Coulter S, et al. Developmental regulation of thyrotropin
receptor gene expression in the fetal and neonatal rat thyroid: Relation to thyroid
morphology and to thyroid-specific gene expression. Endocrinology 2000; 141(1):
340–345.
7. Costa A, Arisio R, Benedetto C, et al. Thyroid hormones in tissues from human
embryos and fetuses. J Endocrinol Invest 1991; 14(7):559–568.
8. Ruiz de Ona C, Obregon MJ, Escobar del Rey F, et al. Developmental changes in rat
brain 5’-deiodinase and thyroid hormones during the fetal period: The effects of fetal
hypothyroidism and maternal thyroid hormones. Pediatr Res 1988; 24(5):588–594.
9. Vulsma T, Gons MH, de Vijlder JJ. Maternal-fetal transfer of thyroxine in congenital
hypothyroidism due to a total organification defect or thyroid agenesis. N Engl J Med
1989; 321(1):13–16.
10. Matsuura N, Konishi J. Transient hypothyroidism in infants born to mothers
with chronic thyroiditis—A nationwide study of twenty-three cases. The Transient
Hypothyroidism Study Group. Endocrinol Jpn 1990; 37(3): 369–379.
11. Haddow JE, Palomaki GE, Allan WC, et al. Maternal thyroid deficiency during
pregnancy and subsequent neuropsychological development of the child. N Engl J
Med 1999; 341(8):549–555.
12. Pop VJ, Kuijpens JL, van Baar AL, et al. Low maternal free thyroxine concentrations
during early pregnancy are associated with impaired psychomotor development in
infancy. Clin Endocrinol (Oxf) 1999; 50(2):149–155.
13. Roti E, Gnudi A, Braverman LE. The placental transport, synthesis and metabolism
of hormones and drugs which affect thyroid function. Endocr Rev 1983; 4(2):131–
149.
14. Mercado M, Yu VY, Francis I, et al. Thyroid function in very preterm infants. Early
Hum Dev 1988; 16(2–3):131–141.
15. Deming DD, Rabin CW, Hopper AO, et al. Direct equilibrium dialysis compared with
two non-dialysis free T4 methods in premature infants. J Pediatr 2007; 151(4):404–
408.
16. Zurakowski D, Di Canzio J, Majzoub JA. Pediatric reference intervals for serum
thyroxine, triiodothyronine, thyrotropin, and free thyroxine. Clin Chem 1999;
45(7):1087–1091.
17. Williams FL, Simpson J, Delahunty C, et al. Developmental trends in cord and
postpartum serum thyroid hormones in preterm infants. J Clin Endocrinol Metab
2004; 89(11): 5314–5320.
18. Carrascosa A, Ruiz-Cuevas P, Potau N, et al. Thyroid function in seventy-five healthy
preterm infants thirty to thirty-five weeks of gestational age: A prospective and
longitudinal study during the first year of life. Thyroid 2004; 14(6):435–442.
364 Brown

19. Frank JE, Faix JE, Hermos RJ, et al. Thyroid function in very low birth weight infants:
Effects on neonatal hypothyroidism screening. J Pediatr 1996, 128(4);548–554.
20. Von Trotsenburg ASP, Vulsma T, Van Santen HM, et al. Lower neonatal screening
thyroxine concentrations in Down syndrome newborns. J Clin Endocrinol Metab
88(4):1512–1515.
21. Dussault JH, Morissette J. Higher sensitivity of primary thyrotropin in screening for
congenital hypothyroidism: A myth? J Clin Endocrinol Metab 1983; 56(4): 849–852.
22. Mandel SJ, Hermos RJ, Larson CA, et al. Atypical hypothyroidism and the very low
birthweight infant. Thyroid 2000; 10(8): 693–695.
23. Brown RS, Demmer LA. The etiology of thyroid dysgenesis—Still an enigma after
all these years. J Clin Endocrinol Metab 2002; 87(9):4069–4071.
24. Djemli A, Van Vliet G, Delvin EE. Congenital hypothyroidism: From paracelsus to
molecular diagnosis. Clin Biochem 2006; 39(5):511–518.
25. Devriendt K, Vanhole C, Matthijs G, et al. Deletion of thyroid transcription factor-1
gene in an infant with neonatal thyroid dysfunction and respiratory failure. N Engl
J Med 1998; 338(18):1317–1318.
26. Krude H, Schutz B, Biebermann H, et al. Choreoathetosis, hypothyroidism, and
pulmonary alterations due to human NKX2-1 haploinsufficiency. J Clin Invest 2002;
109(4):475–480.
27. Clifton-Bligh RJ, Wentworth JM, Heinz P, et al. Mutation of the gene encoding
human TTF-2 associated with thyroid agenesis, cleft palate and choanal atresia. Nat
Genet 1998; 19(4):399–401.
28. Knobel M, Medeiros-Neto G. An outline of inherited disorders of the thyroid hor-
mone generating system. Thyroid 2003; 13(8):771–801.
29. Kopp P, Pesce L, Solis-S JC. Pendred syndrome and iodide transport in the thyroid.
Trends Endocrinol Metab 2008; 19(7):260–268.
30. Sunthornthepvarakui T, Gottschalk ME, Hayashi Y, et al. Brief report: Resistance
to thyrotropin caused by mutations in the thyrotropin-receptor gene. N Engl J Med
1995; 332(3):155–160.
31. Biebermann H, Schoneberg T, Krude H, et al. Mutations of the human thyrotropin
receptor gene causing thyroid hypoplasia and persistent congenital hypothyroidism.
J Clin Endocrinol Metab 1997; 82(10):3471–3480.
32. Dumitrescu AM, Liao XH, Best TB, et al. A novel syndrome combining thyroid
and neurological abnormalities is associated with mutations in a monocarboxylate
transporter gene. Am J Hum Genet 2004; 74(1):168–175.
33. Delange F. Neonatal screening for congenital hypothyroidism: Results and perspec-
tives. Horm Res 1997; 48(2):51–61.
34. Brown RS, Bloomfield S, Bednarek FJ, et al. Routine skin cleansing with povidone-
iodine is not a common cause of transient neonatal hypothyroidism in North America:
A prospective controlled study. Thyroid 1997; 7(3):395–400.
35. Cheron RG, Kaplan MM, Larsen PR, et al. Neonatal thyroid function after propylth-
iouracil therapy for maternal Graves’ disease. N Engl J Med 1981; 304(9):525–528.
36. Brown RS, Bellisario RL, Botero D, et al. Incidence of transient congenital hypothy-
roidism due to maternal thyrotropin receptor-blocking antibodies in over one million
babies. J Clin Endocrinol Metab 1996; 81(3):1147–1151.
37. Daliva AL, Linder B, DiMartino-Nardi J, et al. Three-year follow-up of borderline
congenital hypothyroidism. J Pediatr 2000; 136(1):53–56.
Pediatric Thyroid Disorders 365

38. Calaciura F, Motta RM, Miscio G, et al. Subclinical hypothyroidism in early


childhood: A frequent outcome of transient neonatal hyperthyrotropinemia. J Clin
Endocrinol Metab 2002; 87(7):3209–3214.
39. Miki K, Nose O, Miyai K, et al. Transient infantile hyperthyrotrophinaemia. Arch
Dis Child 1989; 64(8):1177–1182.
40. Ohnishi H, Sato H, Noda H, et al. Color Doppler ultrasonography: Diagnosis of
ectopic thyroid gland in patients with congenital hypothyroidism caused by thyroid
dysgenesis. J Clin Endocrinol Metab 2003; 88(11):5145–5149.
41. Larson C, Hermos R, Delaney A, et al. Risk factors associated with delayed thy-
rotropin elevations in congenital hypothyroidism. J Pediatr 2003; 143(5):587–591.
42. Perry R, Heinrichs C, Bourdoux P, et al. Discordance of monozygotic twins for
thyroid dysgenesis: Implications for screening and for molecular pathophysiology.
J Clin Endocrinol Metab 2002; 87(9):4072–4077.
43. Rose SR, Brown RS, Foley T, et al. Update of newborn screening and therapy for
congenital hypothyroidism. Pediatrics 2006; 117(6);2290–2303.
44. Selva KA, Harper A, Downs A, et al. Neurodevelopmental outcomes in congenital
hypothyroidism: Comparison of initial T4 dose and time to reach target T4 and TSH.
J Pediatr 2005; 147(6):775–780.
45. Bongers-Schokking JJ, Koot HM, Wiersma D, et al. Influence of timing and dose of
thyroid hormone replacement on development in infants with congenital hypothy-
roidism. J Pediatr 2000; 136(3):292–297.
46. Osborn DA, Hunt RW. Postnatal thyroid hormones for preterm infants with transient
hypothyroxinaemia Cochrane Database Syst Rev 2007; (1):CD005945.
47. Williams FL, Ogston SA, van Toor H, et al. Serum thyroid hormones in preterm
infants: Associations with postnatal illnesses and drug usage. J Clin Endocrinol
Metab 2005; 90(11):5954–5963.
48. Tiosano D, Even L, Shen Z, et al. Recombinant Thyrotropin in the diagnosis of
congenital hypothyroidism. J Clin Endocrinol Metab 2007; 92(4):1434–1437.
49. van Wassenaer AG, Kok JH, de Vijlder JJ, et al. Effects of thyroxine supplementation
on neurologic development in infants born at less than thirty weeks’ gestation. N
Engl J Med 1997; 336(1):21–26.
50. van Wassenaer AG, Westera J, Houtzager BA, et al. Ten-year follow-up of children
born at ⬍30 weeks’ gestational age supplemented with thyroxine in the neonatal
period in a randomized, controlled trial. Pediatrics 2005; 116(5):e613–e618.
51. Zakarija M, McKenzie JM. Pregnancy-associated changes in the thyroid-stimulating
antibody of Graves’ disease and the relationship to neonatal hyperthyroidism. J Clin
Endocrinol Metab 1983; 57(5):1036–1040.
52. Skuza KA, Sills IN, Stene M, et al. Prediction of neonatal hyperthyroidism in infants
born to mothers with Graves disease. J Pediatr 1996; 128(2):264–268.
53. Mitsuda N, Tamaki H, Amino N, et al. Risk factors for developmental disorders in
infants born to women with Graves disease. Obstet Gynecol 1992; 80(3 pt 1):359–
364.
54. Zakarija M, McKenzie JM, Munro DS. Immunoglobulin G inhibitor of thyroid-
stimulating antibody is a cause of delay in the onset of neonatal Graves’ disease. J
Clin Invest 1983; 72(4):1352–1356.
55. Bryant WP, Zimmerman D. Iodine-induced hyperthyroidism in a newborn. Pediatrics
1995; 95(3):434–436.
366 Brown

56. Daneman D, Howard NJ. Neonatal thyrotoxicosis: Intellectual impairment and cran-
iosynostosis in later years. J Pediatr 1980; 97(2):257–259.
57. Matsuura N, Konishi J, Fujieda K, et al. TSH-receptor antibodies in mothers with
Graves’ disease and outcome in their offspring. Lancet 1988; 1(8575–8576):14–17.
58. Tamaki H, Amino N, Aozasa M, et al. Universal predictive criteria for neonatal overt
thyrotoxicosis requiring treatment. Am J Perinatol 1988; 5(2):152–158.
59. de Roux N, Polak M, Couet J, et al. A neomutation of the thyroid-stimulating
hormone receptor in a severe neonatal hyperthyroidism. J Clin Endocrinol Metab
1996; 81(6):2023–2026.
60. Holzapfel HP, Wonerow P, von Petrykowski W, et al. Sporadic congenital hyperthy-
roidism due to a spontaneous germline mutation in the thyrotropin receptor gene. J
Clin Endocrinol Metab 1997; 82(11):3879–3884.
61. Gruters A, Schoneberg T, Biebermann H, et al. Severe congenital hyperthyroidism
caused by a germ-line neo mutation in the extracellular portion of the thyrotropin
receptor. J Clin Endocrinol Metab 1998; 83(5):1431–1436.
62. Foley TP, Jr., Abbassi V, Copeland KC, et al. Brief report: Hypothyroidism caused
by chronic autoimmune thyroiditis in very young infants. N Engl J Med 1994;
330(7):466–468.
63. Betterle C, Greggio NA, Volpato M. Clinical review 93: Autoimmune polyglandular
syndrome type 1. J Clin Endocrinol Metab 1998; 83(4):1049–1055.
64. Matsuura N, Konishi J, Yuri K, et al. Comparison of atrophic and goitrous auto-
immune thyroiditis in children: Clinical, laboratory and TSH-receptor antibody
studies. Eur J Pediatr 1990; 149(8):529–533.
65. Takasu N, Yamada T, Takasu M, et al. Disappearance of thyrotropin-blocking anti-
bodies and spontaneous recovery from hypothyroidism in autoimmune thyroiditis.
N Engl J Med 1992; 326(8):513–518.
66. Huang SA, Tu HM, Harney JW, et al. Severe hypothyroidism caused by type
3 iodothyronine deiodinase in infantile hemangiomas. N Engl J Med 2000;
343(3):185–189.
67. McDermott MT, Ridgway EC. Thyroid hormone resistance syndromes. Am J Med
1993; 94(4):424–432.
68. Bercu BB, Orloff S, Schulman JD. Pituitary resistance to thyroid hormone in cysti-
nosis. J Clin Endocrinol Metab 1980; 51(6):1262–1268.
69. Svensson J, Ericsson UB, Nilsson P, et al. Levothyroxine treatment reduces thy-
roid size in children and adolescents with chronic autoimmune thyroiditis. J Clin
Endocrinol Metab 2006; 91(5):1729–1734.
70. Rother KI, Zimmerman D, Schwenk WF. Effect of thyroid hormone treatment on
thyromegaly in children and adolescents with Hashimoto disease. J Pediatr 1994;
124(4):599–601.
71. Rallison ML, Dobyns BM, Keating FR, et al. Occurrence and natural history of
chronic lymphocytic thyroiditis in childhood. J Pediatr 1975; 86(5):675–682.
72. Maenpaa J, Raatikka M, Rasanen J, et al. Natural course of juvenile autoimmune
thyroiditis. J Pediatr 1985; 107(6):898–904.
73. Surks MI, Ortiz E, Daniels GH, et al. Subclinical thyroid disease: Scientific review
and guidelines for diagnosis and management. JAMA 2004; 291(2):228–238.
74. Mali VP, Prabhakaran K. Recurrent acute thyroid swelling because of pyriform sinus
fistula. J Pediatr Surg 2008; 43(4):e27–e30.
Pediatric Thyroid Disorders 367

75. Rapoport B, Chazenbalk GD, Jaume JC, et al. The thyrotropin (TSH) receptor:
Interaction with TSH and autoantibodies. Endocr Rev 1998; 19(6):673–716.
76. Misra M, Levitsky LL, Lee MM. Transient hyperthyroidism in an adolescent with
hydatidiform mole. J Pediatr 2002; 140(3):362–366.
77. Brown RS, Cohen JH Jr, Braverman LE. Successful treatment of massive thyroid
hormone poisoning with iopanoic acid. J Pediatr 1998; 132(5):903–905.
78. Ruiz M, Rajatanavin R, Young RA, et al. Familial dysalbuminemic hyperthyrox-
inemia: A syndrome that can be confused with thyrotoxicosis. N Engl J Med 1982;
306(11):635–639.
79. Mariotti S, Martino E, Cupini C, et al. Low serum thyroglobulin as a clue to the
diagnosis of thyrotoxicosis factitia. N Engl J Med 1982; 307(7):410–412.
80. Costagliola S, Morgenthaler NG, Hoermann R, et al. Second generation assay for
thyrotropin receptor antibodies has superior diagnostic sensitivity for Graves’ dis-
ease. J Clin Endocrinol Metab 1999; 84(1):90–97.
81. Foley TP Jr, White C, New A. Juvenile Graves disease: Usefulness and limita-
tions of thyrotropin receptor antibody determinations. J Pediatr 1987; 110(3):378–
386.
82. Botero D, Brown RS. Bioassay of thyrotropin receptor antibodies with Chinese ham-
ster ovary cells transfected with recombinant human thyrotropin receptor: Clinical
utility in children and adolescents with Graves disease. J Pediatr 1998; 132(4):612–
618.
83. Smith J, Brown RS. Persistence of thyrotropin (TSH) receptor antibodies in children
and adolescents with Graves’ disease treated using antithyroid medication. Thyroid
2007; 17(11):103–107.
84. Bolton J, Sanders J, Oda Y, et al. Measurement of thyroid-stimulating hormone
receptor autoantibodies by ELISA. Clin Chem 1999; 45(12):2285–2287.
85. Reinwein D, Benker G, Lazarus JH, et al. A prospective randomized trial of antithy-
roid drug dose in Graves’ disease therapy. European Multicenter Study Group on
Antithyroid Drug Treatment. J Clin Endocrinol Metab 1993; 76(6):1516–1521.
86. Cooper DS. Antithyroid drugs in the management of patients with Graves’ disease:
An evidence-based approach to therapeutic controversies. J Clin Endocrinol Metab
2003; 88(8):3474–3481.
87. Vaidya VA, Bongiovanni AM, Parks JS, et al. Twenty-two years’ experience in the
medical management of juvenile thyrotoxicosis. Pediatrics 1974; 54(5):565–570.
88. Barnes HV, Blizzard RM. Antithyroid drug therapy for toxic diffuse goiter (Graves
disease): Thirty years experience in children and adolescents. J Pediatr 1977;
91(2):313–320.
89. Lippe BM, Landaw EM, Kaplan SA. Hyperthyroidism in children treated with
long-term medical therapy: Twenty-five percent remission every two years. J Clin
Endocrinol Metab 1987; 64(6):1241–1245.
90. Shulman DI, Muhar I, Jorgensen EV, et al. Autoimmune hyperthyroidism in prepu-
bertal children and adolescents: Comparison of clinical and biochemical features at
diagnosis and responses to medical therapy. Thyroid 1997; 7(5):755–760.
91. Segni M, Leonardi E, Mazzoncini B, et al. Special features of Graves’ disease in
early childhood. Thyroid 1999; 9(9):871–877.
92. Davies TF, Roti E, Braverman LE, et al. Thyroid controversy—Stimulating antibod-
ies. J Clin Endocrinol Metab 1998; 83(11):3777–3785.
368 Brown

93. Vitti P, Rago T, Chiovato L, et al. Clinical features of patients with Graves’ disease
undergoing remission after antithyroid drug treatment. Thyroid 1997; 7(3):369–375.
94. Ikenoue H, Okamura K, Sato K, et al. Prediction of relapse in drug-treated
Graves’ disease using thyroid stimulation indices. Acta Endocrinol (Copenh)
1991;125(6):643–650.
95. van Ouwerkerk BM, Krenning EP, Docter R, et al. Cellular and humoral immunity
in patients with hyperthyroid Graves’ disease before, during and after antithyroid
drug treatment. Clin Endocrinol (Oxf) 1987; 26(4):385–394.
96. Rivkees SA, Sklar C, Freemark M. Clinical review 99: The management of Graves’
disease in children, with special emphasis on radioiodine treatment. J Clin Endocrinol
Metab 1998; 83(11):3767–3776.
97. Zimmerman D, Lteif AN. Thyrotoxicosis in children. Endocrinol Metab Clin N Am
1998; 27(1):109–126.
98. Liaw YF, Huang MJ, Fan KD, et al. Hepatic injury during propylthiouracil therapy in
patients with hyperthyroidism. A cohort study. Ann Intern Med 1993; 118(6):424–
428.
99. Rivkees SA, Dinauer C. An optimal treatment for pediatric Graves’ disease is radioio-
dine. J Clin Endocrinol Metab 2007; 92(3):797–800.
100. Nikiforov Y, Gnepp DR, Fagin JA. Thyroid lesions in children and adolescents after
the Chernobyl disaster: Implications for the study of radiation tumorigenesis. J Clin
Endocrinol Metab 1996; 81(1):9–14.
101. Sherman J, Thompson GB, Lteif A, et al. Surgical management of Graves disease in
childhood and adolescence: An institutional experience. Surgery 2006; 140(6):1056–
1061.
102. Harach HR, Williams ED. Childhood thyroid cancer in England and Wales. Br J
Cancer 1995; 72(3):777–783.
103. Brandi ML, Gagel RF, Angeli A, et al. Guidelines for diagnosis and therapy of MEN
type 1 and type 2. J Clin Endocrinol Metab 2001; 86(12):5658–5671.
104. Mazzaferri EL, Kloos RT. Clinical review 128: Current approaches to primary
therapy for papillary and follicular thyroid cancer. J Clin Endocrinol Metab 2001;
86(4):1447–1463.
105. Hung W, Sarlis NJ. Current controversies in the management of pediatric patients
with well-differentiated nonmedullary thyroid cancer: A review. Thyroid 2002;
12(8):683–702.
106. Newman KD, Black T, Heller G, et al. Differentiated thyroid cancer: Determinants
of disease progression in patients ⬍21 years of age at diagnosis: A report from the
Surgical Discipline Committee of the Children’s Cancer Group. Ann Surg 1998;
227(4):533–541.
107. Dinauer C, Francis GL. Thyroid cancer in children. Endocrinol Metab Clin N Am
2007; 36:779–806.
9
Thyroid Disease and Pregnancy

Susan J. Mandel
University of Pennsylvania School of Medicine, Philadelphia,
Pennsylvania, U.S.A.

INTRODUCTION
When pregnancy is associated with alterations in maternal thyroid function, the
fetus can be affected in two ways: either directly by transplacental passage of
maternal thyroid hormone, antithyroid antibodies, or medications, or indirectly
by adverse influences on maternal physiology. It is important to recognize the
expected alterations in thyroid hormone levels during gestation. The clinician must
be able to differentiate normal physiologic changes from true thyroid disease.
Hyperthyroidism and hypothyroidism may first be detected during pregnancy
and patients with preexisting thyroid dysfunction require close monitoring, and
frequently need adjustment of therapy.

THYROID HORMONE PHYSIOLOGY DURING GESTATION


During normal gestation, there are changes in thyroid hormone physiology that
are reversible after delivery. Serum thyroxine binding globulin (TBG) levels begin
to increase within the first weeks after conception, usually more than doubling
in concentration, with peak levels reached at the middle of gestation (1). Levels
remain elevated until delivery and then normalize in the postpartum period. The
rise in serum TBG concentration results from an estrogen-induced increase in the
sialylation of the protein, which subsequently decreases its hepatic clearance and
prolongs its serum half-life (2).
Associated with the rise in TBG are increases in both serum total T4 and T3
levels, which also plateau at the midtrimester or slightly earlier (1,3). Because of

369
370 Mandel

Figure 1 Serum thyroxine (TT4) and free thyroxine (FT4) levels by trimester. Interquar-
tile ranges are shown by the shaded boxes, with the median value indicated by the line.
Serum TT4 levels rise to approximately 1.5 times the normal nonpregnant reference range.
Although serum FT4 ranges were method-dependent, as shown by the differences in mea-
surement by the Elecsys (Roche, Basel, Switzerland) and Tosoh (Fisher Scientific Inter-
national, Hampton, NH) methods, both methods show a consistent decrease in FT4 as
pregnancy progresses. Abbreviations: NP, nonpregnant (n = 62); 1st , first trimester (n =
105); 2nd , second trimester (n = 39); 3rd , third trimester (n = 64). Source: Photo courtesy
of Carole Spencer. From Ref. 59.

these TBG changes, normal serum T4 and T3 levels throughout pregnancy are pre-
dictably approximately 1.5 times the normal nonpregnant reference range (Fig. 1)
(4). Interestingly, it has been observed the serum T4 and T3 concentrations do not
increase as much during pregnancy as would be expected given the rise in TBG,
and this may be due to decreased TBG saturation (1). The increase in maternal
T4 production that occurs in normal gestation is most evident from observations
of levothyroxine replaced hypothyroid women, who require a 25% to 45% dosage
increase in order to maintain normal serum TSH levels in pregnancy (5–8). Fur-
thermore, the findings of relative hypothyroxinemia and slightly increased serum
TSH levels during pregnancy in women from areas of borderline iodine suffi-
ciency (⬍100 ␮g/day) support the view that pregnancy constitutes a stress for
the maternal thyroid by stimulating thyroidal production (9). There are several
possible explanations for this increased T4 requirement. Early in pregnancy, the
rise in serum TBG results in expansion of the extrathyroidal T4 pool. In addi-
tion, transplacental passage of T4 and placental T4 degradation may contribute
to the increased demand on maternal thyroidal production to maintain euthyroid
status. Lastly, renal clearance of iodide increases because of the higher glomerular
filtration rate in pregnancy (3).
Thyroid Disease and Pregnancy 371

Determination of free T4 (FT4) levels may reflect both methodological


differences as well as alterations due to gravid physiology. With their sensitivity
to binding proteins, however, automated assays appear to show a decrease in serum
FT4 levels as pregnancy progresses compared to their own nonpregnant reference
range. By the third trimester, serum FT4 levels are often lower than the normal
nonpregnant reference range (Fig. 1) (10,11).
Serum TSH levels also fluctuate during pregnancy. The hCG-mediated,
increased thyroid hormone synthesis coinciding with the first trimester peak in
hCG is reflected by a reciprocal fall in serum TSH levels. It has been hypoth-
esized that hCG has thyrotropic activity because of its structural similarity to
TSH, and that the high serum hCG levels stimulate the TSH receptor via a hor-
mone specificity “spillover” syndrome (12). In fact, a positive correlation between
individual FT4 and hCG levels in early gestation has been reported, consistent
with TSH-like activity of hCG (1). Recent studies have documented the 95%
confidence interval lower and upper limits for the first trimester median serum
TSH values in healthy pregnant women to be 0.02 to 0.03 mIU/L and 2.5 to
3.0 mIU/L respectively (13,14). It is critical for clinicians to recognize this appro-
priate decrease in the TSH range during normal pregnancy, since 9% of women
without thyrotoxic symptoms have first trimester TSH levels that are subnor-
mal but detectable compared to the nonpregnant reference range (greater than
0.05 mU/L but less than 0.4–0.5 mU/L) and an additional 9% have values that
are frankly suppressed (⬍0.05 mU/L). Median serum TSH levels then rise during
the second and third trimesters (95% CI 0.03–3.1 mIU/L and 0.13–3.4 mIU/L,
respectively) (13,14).
Renal iodine clearance increases as a result of an increase in glomerular
filtration rate and there is transplacental passage of iodine and iodothyronines as
the fetal-placental unit grows (3). In areas of borderline iodine sufficiency, such
as in many European countries, this loss of iodine, combined with the increase in
thyroid hormone pools from the marked increase in serum TBG levels, may result
in goiter formation. In a prospective study from Belgium, thyroid volume increased
on an average of 18% between the first and third trimesters and was associated with
the biochemical features of thyroid stimulation, a high T3/T4 ratio, and elevated
thyroglobulin levels (9). In areas where iodine intake is more than sufficient, such
as the United States, a palpable goiter should not occur during normal gestation and
if present, this should direct the clinician to investigate possible thyroid hormone
abnormalities (15).

THYROID AUTOIMMUNITY AND EUTHYROIDISM


Among euthyroid women in the reproductive years, up to 18% may have detectable
antithyroid antibodies (16). These asymptomatic euthyroid women with thyroid
autoimmunity have been reported to be at risk for three complications dur-
ing or after pregnancy: increased rates of spontaneous miscarriage and very
preterm delivery (⬍32 weeks gestation) (17), possible development of subclinical
372 Mandel

hypothyroidism during gestation, and risk of postpartum thyroiditis (PPT) (see


section “Postpartum Thyroiditis” of this chapter).
Several studies have reported a twofold increase in the spontaneous mis-
carriage rate early in pregnancy among those euthyroid women who have serum
antithyroid antibodies (either antithyroid peroxidase or antithyroglobulin) detected
in the first trimester (16,18). The majority of these antibody positive women who
miscarry have normal thyroid function. Furthermore, the presence of antithyroid
antibodies either prior to pregnancy or in the first trimester is not correlated with
anticardiolipin antibody positivity, which is also known to be associated with preg-
nancy loss (16,19). The mechanism linking thyroid autoimmunity and miscarriage
is not known. Thyroid autoimmunity may be a marker either for a more gener-
alized activation of the immune system or for subtle changes in maternal/fetal
thyroid metabolism. However, a recent, randomized, controlled trial of levothy-
roxine therapy in euthyroid women with positive serum antithyroid peroxidase
antibodies demonstrated that, compared to untreated women, levothyroxine ther-
apy is associated with decreased rates of both miscarriage (13.85 versus 3.5%, p
⬍ 0.05) and preterm delivery (22.4 versus 7%, p ⬍ 0.05); these lower rates are
similar to a control antithyroid antibody negative population (20).
Euthyroid women with detectable antithyroid antibodies may have slightly
higher first trimester serum TSH values, still remaining within the normal range,
compared to normal pregnant controls. Despite the decrease in antithyroid anti-
body titers with pregnancy progression in these antibody positive women, thyroid
function parameters have been reported to show a progressive deterioration toward
hypothyroidism (20). At term, up to 16% of previously euthyroid women devel-
oped mild subclinical hypothyroidism as indicated by an elevated serum TSH
level (21). However, this study was conducted in an area of borderline iodine
sufficiency, which may have further compromised maternal thyroid gland reserve.
There are no data on the possible development of subclinical hypothyroidism in
an iodine-replete antibody positive pregnant population.

HYPOTHYROIDISM
Overt hypothyroidism is reported to occur in approximately 1 in 1600 pregnancies
(22). However, the prevalence of subclinical hypothyroidism is reported to be
significantly higher, affecting 2.2% of American women screened at 16 to 18 weeks
gestation (23).
Hypothyroxinemia and increased serum TSH levels may occur if true
hypothyroidism is present, or if the mother has been overtreated with antithy-
roid drugs (ATD) for hyperthyroidism. The obstetric complications that have been
associated with hypothyroidism are linked to the decreased maternal thyroid hor-
mone levels that provide a less than optimal environment for both fetal and mater-
nal health. The majority of cases of hypothyroidism are caused by Hashimoto’s
thyroiditis and prior radioiodine or surgical treatment of Graves’ disease. Tran-
sient hypothyroidism may occur as part of autoimmune or PPT, especially if a
Thyroid Disease and Pregnancy 373

woman has had a recent miscarriage (see section “Postpartum Thyroiditis” of this
chapter).

Diagnosis
It is important to diagnose hypothyroidism because of its potential adverse impact
on pregnancy (see section “Pregnancy Outcome” of this chapter), and yet most
patients are relatively asymptomatic. Only 20% to 30% of women with overt
biochemical hypothyroidism (low T4 and high TSH) have symptoms (24) and
complaints of fatigue and weight gain are often attributed to the pregnancy itself.
The majority of patients with subclinical hypothyroidism are asymptomatic as
well. The diagnosis of hypothyroidism is confirmed by finding an elevated serum
TSH concentration, except in the rare instance when hypothyroidism is secondary
to pituitary or hypothalamic disease.
A cost-effectiveness analysis of universal screening for hypothyroidism dur-
ing pregnancy has not been done, in part because the true costs with respect to
fetal outcome are not known. The recently published Endocrine Society guide-
lines for the management of thyroid disorders during pregnancy acknowledged
that universal screening based upon the current published evidence could not
be justified but did recommend targeted case-finding in asymptomatic women
at high-risk for thyroid disease (25). These include those with evidence of thy-
roid autoimmunity because of either a past history of PPT, prior detection of
antithyroid antibodies, or previous treatment for hyperthyroidism, even if they are
euthyroid without levothyroxine therapy. In such women, autoimmune damage
may not affect basal thyroid hormone output in the nonpregnant state, but may
impair the thyroid’s ability to compensate for the increased production needed
during pregnancy. In addition, 15% of diabetic women with type 1 diabetes may
develop clinical hypothyroidism during pregnancy (26). Thyroid function testing
is also recommended for women with other autoimmune disorders, a family his-
tory of thyroid disease, prior miscarriage (see section “Thyroid Autoimmunity and
Euthyroidism” of this chapter), and potentially decreased thyroid gland synthetic
function because of a prior lobectomy or head and neck irradiation. In addition,
all levothyroxine-replaced hypothyroid women must be monitored during preg-
nancy (see section “Treatment” of this chapter) (25). Unfortunately, such targeted
screening will still miss up to 30% of women with hypothyroidism in the first
trimester (27).

Pregnancy Outcome
Overt hypothyroidism can be associated with anovulatory cycles and subse-
quent infertility. However, hypothyroid women may become pregnant and several
retrospective case series have investigated pregnancy outcome in these women
(Table 1). The likelihood of complications is correlated with the severity of the
hypothyroidism (overt versus subclinical) and the adequacy of maternal treatment
(5,17,24,28–32). The majority of women reported in these studies had less than
374 Mandel

Table 1 Pregnancy Complications Reported in Hypothyroid Women


Subclinical hypothyroidism Overt hypothyroidism
(%) (%)

Spontaneous abortion (5,32) 10–70 60


Pregnancy-induced 0–17 0–44
hypertension/Preeclampsia
(24,28,30,31)
Abruption (5,28–31) 0 0–19
Stillbirth (24,28–30) 0–3 0–12
Anemia (29,30) 0–2 0–31
Postpartum hemorrhage (29–31) 0–17 0–19
Preterm birth majority from 0–9 20–31
premature delivery due to
preeclampsia (low birth
weight) (5,17,29,30)

optimal, prenatal care as the average initial antenatal visit occurred between 16 and
20 weeks gestation. However, a recent prospective study investigated pregnancy
outcome in women with untreated subclinical hypothyroidism diagnosed prior to
20 weeks by gestational age-specific serum TSH and FT4 levels. Compared to
euthyroid women, the rates of preterm delivery (⬍34 weeks gestation) and pla-
cental abruption were significantly higher (relative risk 1.8 and 3.0 respectively,
p ⬍ 0.03) (14). Levothyroxine therapy, especially if optimized by the midgestation,
may ameliorate some of these complications (5,24).
In addition to pregnancy complications, does maternal hypothyroxinemia
pose a threat to fetal development? A recent report documented severe retarda-
tion of fetal development, with biparietal diameter and femur length of less than
the third centile in a woman with inadequately treated hypothyroidism and a
serum TSH level of 72 mU/L at 29 weeks gestation. With appropriate increase in
her levothyroxine dosage and normalization of her serum TSH level, these fetal
parameters normalized by 39 weeks gestation (33). Although the fetus had normal
thyroid function, this case illustrates the pivotal role of maternal thyroid hormone
for fetal somatic growth.
Thyroid hormone is also necessary for normal fetal neurologic development
and the fetal thyroid does not begin to function until 12 weeks of life. It is now
evident that maternal T4 crosses the placenta (34). As early as 7 weeks, gestation,
T3 is present in the fetal neurologic tissues, which originates from the transpla-
cental passage of maternal T4 that undergoes intracellular deiodination in the fetal
brain (35). The relative contribution of maternal thyroid hormone versus fetal
thyroid hormone to fetal neurologic development is unknown. In areas of iodine
deficiency where both maternal and fetal thyroid status are compromised, neuro-
logic cretinism occurs. In contrast, infants born with congenital hypothyroidism
Thyroid Disease and Pregnancy 375

in areas of iodine adequacy have normal neurologic function at birth and postnatal
levothyroxine therapy is required to continue normal neuronal maturation. There-
fore, it is presumed that the transplacental passage of maternal T4 is sufficient to
maintain normal fetal neurologic development in utero.
The contribution of maternal thyroid hormone to the brain maturation of a
fetus with intact thyroid function is inadequately understood. Cognitive function
is impaired and brain DNA and protein content are decreased in rats born to thy-
roidectomized mothers and thereby deprived of maternal thyroid hormone (36).
These deficits are not as severe if maternal hypothyroidism occurs only during
the second half of gestation, when the rat fetal thyroid gland activity is adequate
(37). Data in humans are less direct. Although, early studies of children born to
women who were hypothyroximeic during pregnancy reported impaired mental
development (38), they have been criticized for their lack of accurate biochemical
assessment of hypothyroidism. However, this issue has been addressed by a more
recent study. Haddow and colleagues performed neuropsychological testing of 62
children (of average age eight years) born to women who had elevated serum TSH
levels at 17 weeks gestation and compared these to results for 124 control children
matched for maternal educational level and age. Children born to hypothyroid
women (partially treated or untreated) scored on average four points lower than
control children (p = 0.06) on the full-scale IQ score of the Wechlser Intelligence
Scale for Children. This difference was even more marked, a decrease of seven IQ
points (p = 0.005), when the subset of 48 children born to untreated hypothyroid
mothers was compared to control children (39). Furthermore, the decrease in the
IQ score was inversely proportional to the degree of maternal serum TSH eleva-
tion (40). Although differences in the postnatal environment cannot be excluded
as etiologic factors, this study strongly suggests that untreated or inadequately
treated maternal hypothyroidism during pregnancy adversely affects fetal brain
development.
Lastly, a very small percentage of women with atrophic Hashimoto’s thy-
roiditis may have antibodies that block thyroidal stimulation by TSH. These
antibodies may be detected by assays for TSH receptor binding inhibitory
immunoglobulins (TBII), which assess the ability of maternal immunoglobulin
to block TSH binding to the TSH receptor in vitro. Transient congenital hypothy-
roidism may be caused by the transplacental passage of these antibodies that then
block TSH stimulation of the neonatal thyroid, analogous to but opposite of the
situation of neonatal Graves’ disease. The estimated prevalence of this disorder is
1 in 180,000 births, or 2% of infants with congenital hypothyroidism (41). The
antibodies can be measured in both the mother and the neonate, and if present, may
indicate that lifelong levothyroxine therapy may not be necessary for the infant.

Treatment
Several studies have documented that levothyroxine requirements increase in
many hypothyroid women during pregnancy (5–8). There are various possible
376 Mandel

explanations for this increased requirement and each may have relative impor-
tance at different times in gestation. In early pregnancy, the concentration of TBG
rapidly increases and more thyroid hormone may be needed to saturate binding
sites. Glomerular filtration rate increases resulting in increased iodide clearance.
Later, with placental growth, there is increased metabolism of T4 to its inactive
metabolite reverse T3 by the high levels of placental type 3 deiodinase (3). In addi-
tion, there is transplacental passage of T4 (34). Lastly, there may be alterations in
the volume of distribution of thyroid hormone because of both gravid physiology
and the fetal/placental unit.
For patients initially diagnosed with overt hypothyroidism during pregnancy,
a daily dose of 2 ␮g/kg/day should be started, which is higher than the full replace-
ment dose in the nonpregnant patient and accounts for the higher requirement in
pregnancy (22). If the serum TSH is first found to be only minimally elevated
(⬍10 mU/L) in pregnancy, a levothyroxine dose of 0.1 mg/day may be adequate.
In those with known hypothyroidism taking levothyroxine replacement, the need
for dosage adjustment may depend upon the etiology of hypothyroidism, with an
increase needed in 76% of women who have undergone prior radioiodine ablation
or surgery, but only in 47% who have Hashimoto’s thyroiditis (8). Levothyroxine
requirements generally increase in the first trimester and persist through gesta-
tion. A recent, prospective study documented the median time for levothyroxine
dosage increase was 8 weeks gestation. However, this was using the upper limit
of a nonpregnant TSH reference range (5.0 mIU/L) for dosage adjustment (7). If
a trimester-specific reference TSH range were used, then it is possible that the
increased requirement would be manifest earlier. It is also important to remember
that 25% of those with initial normal serum TSH levels in first trimester and 37%
of those with initial normal serum TSH concentrations in second trimester will
later require dosage increases (8). The increased dosage requirement appears to
plateau after 20 weeks gestation (7). Women with subclinical hypothyroidism,
who are taking less than replacement dosages of levothyroxine may not require
a dosage increase during gestation because the residual thyroid gland is able to
increase synthesis of thyroid hormone. These women may be at increased risk for
PPT, however (see section “Postpartum Thyroiditis” of this chapter).
Levothyroxine replaced hypothyroid women should have thyroid function
monitored as soon as they become pregnant and every four to six weeks in the
first half of pregnancy (Table 2) (25,42). There are two schools of recommen-
dations for adjustment of levothyroxine dosage requirements during pregnancy.
The first is to increase the dose only once the serum TSH is abnormal compared
to trimester-specific values; pragmatically, the upper limit of 2.5 mIU/L may be
used during gestation. Kaplan has proposed that the increment in levothyroxine
dosage can be based upon the initial degree of TSH elevation. For those with
serum TSH levels ⬍10 mU/L, the average increase was approximately 50 ␮g/day;
for those with serum TSH values between 10 and 20 mU/L, it was approximately
75 ␮g/day; and for those with serum TSH values ⬎20 mU/L, the average increase
was approximately 100 ␮g/day (8). The second is to recommend that women with
Thyroid Disease and Pregnancy 377

Table 2 Guidelines for Clinical Management of Maternal Hypothyroidism During


Pregnancy (25,42)

1. Optimize levothyroxine dosage prior to pregnancy (TSH 0.5–2.5 mIU/L)


2. Check serum TSH level as soon as pregnancy is confirmed
3. Adjust levothyroxine dosage to maintain a serum TSH level ⬍2.5 mIU/L. Increment
in dosage may depend upon etiology of hypothyroidism
Athyreosis (Graves’ after 131 I therapy, thyroid cancer) approximately 45%
increment
Hashimoto’s thyroiditis approximately 25% increment
Subclinical hypothyroidism may not require increment
4. TSH should be monitored every 4–6 wks in the first half of pregnancy;
subsequently, it can be checked every 8 wks, unless a dose adjustment is made.
5. Patients should be instructed to separate levothyroxine ingestion and prenatal
vitamins containing iron, iron or calcium supplements, or soy products by at least
4 hrs.
6. After delivery, the levothyroxine dose should reduced to the prepregnancy dosage
and the serum TSH level should be rechecked at 6-wk postpartum.

hypothyroidism should be instructed to increase their usual levothyroxine intake


by two additional doses each week immediately on confirmation of pregnancy
and to contact their health care provider so that a program of TSH guided dose
adjustments can be instituted (7). Patients should be instructed to separate levothy-
roxine ingestion from that of prenatal vitamins containing iron and especially iron
supplements, calcium supplements, and soy products which may interfere with
levothyroxine absorption (42,43). Thyroid function should be rechecked four to
six weeks after any dose change. The dose may be lowered to prepregnancy levels
at delivery and thyroid function should be measured at the 6-week postpartum
visit.

HYPERTHYROIDISM
HCG-Associated Thyrotoxicosis and Hyperemesis Gravidarum
A spectrum of hCG-induced hyperthyroidism occurs during pregnancy and this
entity has recently been referred to as “gestational thyrotoxicosis” (44,45). As
previously noted, it is postulated that hCG activates the TSH receptor by a spillover
mechanism because of the molecular similarity between these two glycoproteins
(12). Findings range from an isolated subnormal serum TSH concentration (up to
18% of pregnancies) to elevations of free thyroid hormone levels in the clinical
setting of hyperemesis gravidarum. In women without symptoms of thyrotoxicosis,
the serum TSH level may be subnormal but detectable in approximately 9% and
undetectable (⬍0.05 mU/L) in an additional 9% (46). Systematic screening of
1900 consecutive pregnant women at their initial antenatal visit demonstrated low
serum TSH and elevated FT4 levels in 2.4%, half of whom had weight loss, lack of
378 Mandel

weight gain, or unexplained tachycardia (45). In all these women, normalization


of the FT4 paralleled the decrease in hCG.
It has been observed that hyperemesis gravidarum, defined as severe nausea
and vomiting in pregnancy resulting in weight loss and fluid and electrolyte
disturbances, has been associated with abnormal thyroid function tests. Suppressed
serum TSH levels may occur in 60% of these patients, with elevated FT4 levels in
almost 50% (47). Serum hCG concentrations correlate positively with the serum
FT4 levels and inversely with serum TSH determinations. The magnitude of the
deviation from normal values increases with the severity of nausea and vomiting
(48). Furthermore, thyroid stimulating activity as measured by adenylate cyclase
activity per IU of hCG is reported to be greatest in women with hyperemesis
gravidarum when compared to those with occasional or no vomiting (44). The
vomiting may be related to the elevated, hCG-mediated estradiol production since
estradiol levels are higher in hyperemesis subjects than controls, rather than to the
thyroid stimulation itself (49).
Similar thyroid hormone changes and emetic symptoms may be present with
multiple gestations, which are associated with higher peak and more sustained
hCG levels (7,44). In addition, a recent case report further supports the concept of
hCG-induced thyrotoxicosis. A woman and her mother with recurrent gestational
thyrotoxicosis were found to have a missense mutation in the extracellular domain
of their TSH receptor that caused a two- to –threefold increase in activation (cAMP
generation) when exposed to hCG compared to wild type receptor (50).
Gestational thyrotoxicosis is transient and usually resolves within 10 weeks
of the diagnosis (47). In one study of 44 women with hyperemesis gravidarum,
serum FT4 levels normalized by 15 weeks gestation while serum TSH remained
suppressed until 19 weeks gestation (51). Clinically, this disorder differs from
Graves’ disease in several ways: (1) nonautoimmune origin, with negative antithy-
roid and anti-TSH receptor antibodies; (2) absence of goiter; (3) resolution in
almost all patients after 20 weeks gestation (52). Hyperthyroid symptoms such
as weight loss, or lack of normal pregnancy weight increase and tachycardia are
present in 50% of women with gestational thyrotoxicosis (45). However, ophthal-
mopathy, which is autoimmune in origin, is not seen with this disorder. Treatment
with ATD is not recommended unless coincident Graves’ disease is present (25).
Patients with hyperemesis who remain symptomatic after 20 weeks gestation with
elevated thyroid hormone concentrations and suppressed TSH levels may be con-
sidered for antithyroid drug therapy. More than likely, such patients probably have
mild Graves’ disease.

Graves’ Disease
Hyperthyroidism is reported to affect 1 in 500 pregnancies, with Graves’ disease
accounting for the vast majority of cases (85%), with less common causes being
toxic nodular disease in 10% and thyroiditis in 1% to 2% (53). Autoimmune
thyroiditis should also be considered as a possible cause, especially if a woman
Thyroid Disease and Pregnancy 379

has had a recent miscarriage, which has been reported to trigger “postpartum”
thyroiditis (see section “Postpartum Thyroiditis” of this chapter) (54).
The activity of Graves’ disease fluctuates through pregnancy with TSH
receptor antibody (TRAb) patterns generally reflecting the clinical course of the
disease (55). TRAb may be elevated in the first trimester, but values often decrease
over the second and third trimesters and may become undetectable before increas-
ing again postpartum (56,57). Clinically, patients may experience relapse or exac-
erbation of Graves’ disease by 10 to 15 weeks of gestation. However, Graves’
disease may remit in the late second and third trimesters, a time of known immune
tolerance (58). This disease pattern is thought to be due to decreases in TRAb, as
described above, rather than increases in inhibitory anti-TSH receptor antibodies
(56). Often, antithyroid drug dosage can be reduced or even discontinued late in
gestation, only to be followed by a worsening of the disease in the postpartum
period (58).
Graves’ disease may affect a pregnancy in three scenarios. First, women may
have active Graves’ disease (either treated or untreated) that can be exacerbated in
the first trimester. Second, women in remission may experience a relapse during
pregnancy. Third, Graves’ disease may occur for the first time during gestation
(59). In women who have been euthyroid throughout pregnancy, but have been
treated with ATD for Graves’ disease previously, hyperthyroidism may recur in
the postpartum period. However, this may represent either the thyrotoxic phase of
PPT in up to 25% or relapse of Graves’ disease. Even in those with PPT, Graves’
disease may recur after resolution of PPT (60).

Diagnosis
The clinical diagnosis of hyperthyroidism may be difficult because pregnancy is
itself a hypermetabolic state with symptoms of palpitations and heat intolerance.
In addition, patients will usually have increased irritability, decreased exercise
tolerance, and fatigue. Patients may describe an inability to control their emotions,
with otherwise small irritants culminating in what may be perceived as exag-
gerated emotional responses. They are aware of increasing shortness of breath
climbing stairs. The astute clinician must be cognizant of this constellation of
symptoms so that the patient can be appropriately screened for hyperthyroidism.
The examination usually reveals the presence of a diffuse goiter, sometimes
with a bruit or thrill. Other clinical signs may be present as described in
Chapter 2.
Laboratory studies reveal a serum TSH level below the trimester-specific
95% lower confidence limit, usually with elevated serum thyroid hormone con-
centrations. However, it must be remembered that up to 50% of women with
hyperemesis gravidarum may have a suppressed serum TSH level and/or elevated
FT4 (49). An elevated free T3 index or free T3 level may be the most clini-
cally useful test to distinguish hyperthyroid patients from those with hyperemesis
gravidarum as less than 15% of hyperemetic women with have elevations in these
380 Mandel

Table 3 Pregnancy Complications Reported in Hyperthyroid Women


Controlled hyperthyroidism Untreated
on ATD therapy hyperthyroidism

Miscarriage (62,65) 8–10% 21%


Preterm delivery (53,62,63) 3–14% 21–88%
Preeclampsia (53) 2% 11%
Heart failure (63) 3% 63%
Stillbirth (62,63) 0% 7–50%
Small for gestational age (64,65) less more
Thyroid storm less more

Abbreviation: ATD, antithyroid drug.

measures (49). TSH receptor antibodies are usually detectable and may also be of
diagnostic utility.

Pregnancy Outcome
Throughout the discussion of the risks and treatment of Graves’ disease during
pregnancy, it is important to remember that in reality there are two patients,
the mother and the fetus. Maternal hyperthyroidism is associated with increased
morbidity for both mother and fetus. Prior to the development of ATD, only about
50% of hyperthyroid women were even reported to be able to conceive. Of those
who conceived, spontaneous miscarriage and premature delivery occurred in half
(61).
The frequency of poor outcomes for both mother and fetus is correlated with
the degree and duration of hyperthyroidism, with the highest rates in those women
with uncontrolled disease and a decreased risk in those appropriately treated with
ATD (Table 3) (53,62–65). In addition, one study reported an increased incidence
of congenital malformations (imperforate anus, polydactyly, harelip) if maternal
hyperthyroidism is uncontrolled at the time of embryogenesis in the first trimester,
although ATD therapy itself is not associated with a higher incidence of structural
anomalies (66). These results highlight the importance of control of maternal
hyperthyroidism to ensure optimal pregnancy outcome. Significantly, subclinical
hyperthyroidism, defined as a serum TSH level below the 2.5th percentile for
gestational age and a normal serum FT4 level, has not been found to be associated
with adverse pregnancy outcomes (67).

Treatment
Antithyroid Drugs
ATD are the main treatment for Graves’ disease during pregnancy. Propylthiouracil
(PTU) and methimazole (MMI, Tapazole R
) have both been used during gestation.
They inhibit thyroid hormone synthesis via reduction in iodine organification
Thyroid Disease and Pregnancy 381

and iodotyrosine coupling (see chap. “Graves’ Disease”). Pregnancy itself does
not appear to alter the maternal pharmacokinetics of MMI, although serum PTU
levels may be lower in the latter part of gestation compared to the first and
second trimesters (68). PTU is more extensively bound to albumin at physiologic
pH, whereas MMI is less bound, which hypothetically might result in increased
transplacental passage of MMI relative to PTU. Historically, PTU was preferred
over MMI, partly due to early experimental data suggesting that PTU, which is
more highly protein bound than MMI, had more limited transplacental passage
than MMI (69). Since then, however, other studies have found that both drugs
readily cross the placenta (70,71). No such data evaluating simultaneous maternal
and cord levels are available for MMI.
The goals of treatment of Graves’ disease during pregnancy are to control
maternal hyperthyroidism with vigilant monitoring of maternal thyroid function
and to optimize fetal outcome with careful surveillance of fetal development.
Throughout gestation, it is critical that the endocrinologist and obstetrician com-
municate frequently so that biochemical and clinical parameters may be correlated.
Signs of clinical improvement include maternal weight gain and decrease in pulse
rate, as well as appropriate fetal growth. For example, if there is concern because
of lack of maternal weight gain in conjunction with mild elevations in thyroid
hormone levels, the initiation of a low dose of ATD should be discussed.
ATD: Effect on the Fetus
The clinician must assume that both PTU and MMI cross the placenta and may
decrease fetal thyroid hormone production. For women with Graves’ disease,
fetal thyroid status reflects the influence of two maternal factors, both of which
cross the placenta: maternal ATD dosage and maternal TRAb activity. Different
assays for maternal TRAb exist. The more commonly used radioreceptor assay
is the TBII. This assay does not distinguish between those antibodies that bind
to and block the TSH receptor versus those that stimulate the receptor, resulting
in increased thyroid production (72). However, in the majority of women with
Graves’ disease, TBII levels are reported to represent stimulating antibodies and
correlate with maternal disease activity (73). The currently available bioassay is
the thyroid stimulating immunoglobulin (TSI), which measures the generation
of cyclic adenosine monophosphate by cells that express TSH receptor when
incubated with the patient’s serum (73).
Therefore, given these two potential opposing influences on fetal thyroid
function, what are the data correlating fetal thyroid function with maternal ATD
dosage? There are seven published studies examining a dose-response relationship
between maternal ATD dose and neonatal thyroid function. Three have reported
a direct correlation (64,73,74) and four have not demonstrated this (70,74–76). In
fact, one study reported that even low daily ATD dosages (PTU 100 mg or less,
MMI 10 mg or less) at term may affect the fetal thyroid function; an elevated cord
TSH level was found in 23% of babies born to such PTU-treated mothers and in
14% of those treated MMI (76). The lack of correlation between maternal dosage
382 Mandel

and fetal thyroid function may also reflect maternal factors as well, because there
is individual variability in serum PTU levels after a standard oral dose (72,77).
The second factor influencing fetal thyroid function is the transplacental pas-
sage of maternal TRAb resulting in excessive fetal thyroid stimulation. Clinically,
this becomes relevant at 24 to 26 weeks, and maternal levels reflect the degree
of fetal exposure (78). There is a strong correlation between maternal and cord
TBII levels at term with development of neonatal hyperthyroidism (see section
“Fetal/Neonatal Hyperthyroidism” of this chapter). In contrast, the continued use
of maternal ATD therapy at term, in conjunction with low maternal TBII levels
may result in elevated serum TSH levels in approximately 50% to 60% of infants
(73). In this scenario, fetal thyroid function may reflect the relative importance
of maternal ATD dosage when maternal immune thyroid stimulation is low. It is
possible that in pregnant women with toxic nodules, a dose relationship may be
more likely to be seen, since there is no contribution of fetal thyroid stimulation
by the maternal immune system.
Therefore, given these varied influences on fetal thyroid function, coupled
with maternal individual differences in ATD pharmacology, it is not surprising
that fetal thyroid status is not strictly correlated with maternal ATD dosage. Based
on the literature, current maternal thyroid status rather than ATD dose, is the most
reliable marker for titration of ATD therapy to avoid fetal hypothyroidism (75).
A recently published abstract analyzed fetal cord FT4 and TSH levels at birth in
relation to maternal serum FT4 levels in 249 women with Graves’ disease who
continued ATD through delivery (79). The authors reported that low fetal cord
blood FT4 levels were avoided only when the maternal serum FT4 concentra-
tion was ⬎1.9 ng/dL, although one infant whose mother’s serum FT4 level was
2.1 ng/dL developed central congenital hypothyroidism. The normal nonpregnant
reference range for FT4 in this study was 0.8 to 1.9 ng/dL (10.3–24.5 pmol/L).
However, if the maternal serum FT4 is in the lower two-thirds of the nonpregnant
normal reference range, 36% of neonates have a decreased FT4 and a decreased
FT4 is found in all neonates if the maternal FT4 is below normal (75).
In addition, overdosage with ATD alone may result in fetal or neonatal
goiter, which may cause respiratory distress at birth if markedly enlarged. Goiter
is reported to have occurred more frequently in older reports where concomitant
iodide therapy was used. Because of either transplacental ATD or iodide-induced
inhibition of fetal thyroid hormone production, fetal serum TSH levels increase,
resulting in stimulation of thyroid growth. A fetal utrasound should be obtained
for all women who are still taking relatively high ATD doses at 26 to 28 weeks
(PTU ≥450 mg/day, MMI ≥30 mg/day) (55). If a fetal goiter is detected on a late
pregnancy ultrasound, the clinician must consider whether this represents “Fetal
Hyperthyroidism” (see below) or fetal “hypothyroidism” because of transplacental
passage of maternal ATD therapy. Intrauterine growth retardation may occur with
either condition, but fetal tachycardia (⬎160–180 beats per minute) and advanced
fetal bone age is highly suggestive of hyperthyroidism (55,80,81). In cases where
neonatal goiter has occurred because of maternal ATD use, resolution usually
Thyroid Disease and Pregnancy 383

occurs within the first two weeks of life with dissipation of the drug (74). Therefore,
one approach is to stop maternal ATD therapy and monitor the fetal goiter by
ultrasound.
There are several case reports of intra-amniotic levothyroxine injections for
treatment of the fetal goiter due to maternal ATD exposure. However, in two
recent reports (82,83), the injections occurred while the maternal PTU dose was
lowered. Therefore, it is difficult to distinguish the relative importance of each
factor on the resolution of the fetal goiter. A third recent case report demon-
strated that cessation of maternal ATD therapy alone resulted in decrease in the
fetal goiter documented ultrasonographically (84). In cases of fetal goiter where
hypothyroidism is suspected because of transplacental ATD, it may be prudent to
discontinue or substantially decrease maternal ATD, and follow the goiter with
sequential ultrasounds. If reduction in size does not occur within two to three
weeks, periumbilical blood sampling should be performed to determine fetal thy-
roid function. If still low, intra-amniotic levothyroxine therapy should be given.
Four studies have reported no defects in either the cognitive and somatic
development of children exposed to maternal ATD in utero (85–88), even after
accounting for higher dosage or first trimester exposure. These were cross-
sectional studies that measured cognitive development by intelligence quotient.
Therefore, it is unknown if transient, or more subtle developmental changes might
have been present. However, maternal thyroid hormone levels were not reported,
so it is unknown if maternal hypothyroxinemia, a possible risk factor for cognitive
impairment, was present.
There have been reports of an association of maternal MMI therapy with
aplasia cutis, a heterogeneous group of disorders in which localized or widespread
areas of skin are absent at birth. The lesions seen in infants born to MMI treated
mothers have all been all localized scalp defects. However, the reported fre-
quency of this disorder in infants born to MMI-treated mothers is not higher than
the expected sporadic frequency (68). No cases to date have been reported with
PTU therapy despite its more widespread use. For this reason, the recently pub-
lished Endocrine Society guidelines on management of thyroid disorders during
pregnancy recommend, if available, PTU to MMI for initial therapy of maternal
hyperthyroidism, at least in the first trimester when organogenesis occurs (25). In
addition, the rare disorder of absent or hypoplastic nipples associated with choanal
atresia has been reported in two infants born to mothers treated with MMI during
the first trimester (89).

ATD: Treatment Guidelines (Table 4)


ATD dosage should be titrated to maintain either maternal serum FT4 levels at or
up to 10% above the normal limit of the nonpregnant reference range reported by
that laboratory or maternal total T4 at the upper limit of the pregnancy appropriate
reference range (1.5 times the nonpregnant reference range) (25). Maternal serum
T3 levels may not be as helpful because there is no correlation with fetal thyroid
function (75). Practically this means that ATD dosage should be adjusted to
384 Mandel

Table 4 Guidelines for Clinical Management of Maternal Hyperthyroidism During


Pregnancy (25,59)

Treatment goal: Subclinical hyperthyroidism


FT4: Use nonpregnant reference range
Titrate to or 10% above upper normal limit
Total T4: Use pregnant reference range (1.5 times nonpregnant reference range)
Titrate to upper normal limit
TSH: May consider measuring TSH after 2–3 mo
Titrate at or just below lower trimester-specific limit

1. Use the lowest dose of ATD to maintain maternal thyroid hormone levels at above
targets. Because PTU has not been implicated in causing aplasia cutis, initiating
therapy with PTU is preferred.
2. Check maternal function tests monthly
3. ATD dose can usually be lowered or discontinued (30%) by 32–36 weeks gestation
4. If either high maintenance ATD doses are required (PTU ⬎450 mg/day, MMI
⬎40 mg/day or if a patient is nonadherent or allergic to ATD therapy, surgery
(subtotal thyroidectomy) should be considered.
5. Low doses of iodides may be used transiently, especially preoperatively.
6. Frequent communication between the endocrinologist and obstetrician is essential so
that ATD dose titration is done with monitoring of fetal growth.
7. TRAb measurement and fetal ultrasound should be considered as discussed in Table 6.

Abbreviations: ATD, antithyroid drug; PTU, propylthiouracil; MMI, methimazole; TRAb, TSH recep-
tor antiobodies.

maintain a serum FT4 of 1.7 to 2.0 ng/L. If a woman has mild Graves’ disease with
these values as her initial indices, treatment may be withheld and her thyroid status
monitored as long as she has satisfactory clinical progression of pregnancy. Later,
if the serum TSH becomes detectable, it should be kept at or just below at or just
below the 95% confidence interval trimester-specific lower limit. (25). Therefore,
the therapeutic treatment goal for Graves’ disease during pregnancy is actually
subclinical hyperthyroidism compared to normal pregnant physiology. However,
as mentioned before, there are no reported gestational adverse effects of maternal
subclinical hyperthyroidism (67,90) and this slight degree of undertreatment of
the mother optimizes fetal outcome.
The initial ATD dosage may vary depending upon the degree of hyper-
thyroidism. We prefer to begin therapy with PTU because there are no reported
cases of PTU-associated aplasia cutis. However, if a woman cannot tolerate PTU
or finds it difficult to take the prescribed number of pills (PTU usually requires
multiple daily dosages, whereas MMI can often be given once daily), MMI may
be substituted. It is our approach generally not to use more than 450 mg of PTU or
30 mg of MMI daily. The median time to normalization of the maternal FT4 index
is seven to eight weeks for both PTU and MMI (91), although improvement in
Thyroid Disease and Pregnancy 385

parameters may be seen earlier at three to four weeks. One should reassess mater-
nal FT4 or total T4 three to four weeks later and adjust ATD dosage based upon
the decrement in thyroid hormone levels. As in nonpregnant women with Graves’
disease, maternal serum TSH levels may remain suppressed for several weeks
after normalization of thyroid hormone levels, and it is not helpful to monitor the
serum TSH early in treatment. Graves’ disease may improve in the third trimester
and with progressive decreases in ATD dosage throughout pregnancy, therapy may
be stopped by 32 to 34 weeks gestation in 30% of women (92). Of course, the
same spectrum of adverse effects related to ATD therapy in the nonpregnant state
applies to use during gestation (see chap. “Graves’ Disease”).

Beta-Adrenergic Blockers
Beta-adrenergic blocking agents may be used transiently to control adrenergic
symptoms, until ATD therapy decreases thyroid hormone levels. There is a recent
report of a higher rate of spontaneous first trimester miscarriages in women who
were treated with combined ATD and propranolol therapy compared to ATD alone,
although both groups had similar levels of thyroid hormone (93). However, this
was a small series and propranolol was prescribed for 6 to 12 weeks, which may
be longer than would be typically necessary in most patients.

Iodides
Chronic use of iodides during pregnancy has been associated with hypothyroidism
and goiter in neonates, sometimes resulting in asphyxiation because of tracheal
obstruction (68). However, a recent report of low-dose potassium iodide (6–40 mg/
day) administered to selected pregnant hyperthyroid women with maintenance of
maternal FT4 levels in the upper half of the nonpregnant reference range did not
cause goiter, although 6% of newborns had an elevated serum TSH level (94).
Since the experience with iodides is more limited, iodides should not be used as a
first line therapy for women with Graves’ disease, but could be used transiently if
needed in preparation for thyroidectomy.

Surgery
Subtotal thyroidectomy is usually only considered during pregnancy as therapy for
maternal Graves’ disease if consistently high levels of ATD (PTU ⬎450 mg/day,
MMI ⬎40 mg/day) are required to control maternal hyperthyroidism, if a patient
is nonadherent or allergic to ATD therapy, or if compressvie symptoms exist
because of goiter size. If a woman has experienced severe ATD-related side-effects
such as agranulocytosis, she should receive transient therapy with supersaturated
potassium iodide solution (50–100 mg/day) for 10 to 14 days prior to surgery. The
timing of surgery is usually in latter half of the second trimester. The rationale for
not performing surgery in the first trimester is that this is the time of the highest
spontaneous abortion rate and surgery could possibly further increase the risk.
However, there are no definitive data supporting an increased miscarriage rate
386 Mandel

related to the surgical procedure and anesthesia if performed in the first trimester
(95) and subtotal thyroidectomy may be done if clinically indicated.
131
I Therapy
Because of its adverse effects on the fetus, the use of 131 I therapy is completely
contraindicated in pregnancy, especially after 12 weeks gestation when the fetal
thyroid begins to concentrate radioiodine with an even greater avidity than the
maternal thyroid. In addition, other fetal tissues are generally more radiosensitive
(96). A pregnancy test should be performed in all women prior to radioiodine
therapy. However, inadvertent administration of 131 I in early pregnancy may occur
and a survey was sent to endocrinologists regarding their experience with this
situation. Of 237 cases, therapeutic abortion was advised and performed for 55
patients. In the remaining 182 pregnancies, the risk of stillbirths, spontaneous
abortions, and fetal abnormalities was not higher than the general population (97),
perhaps because the fetal whole-body irradiation from a therapeutic 131 I dose for
Graves’ disease at this time is calculated to be below the threshold associated with
increased congenital defects (96). However, six infants were hypothyroid at birth,
four of them were mentally retarded. 131 I therapy had been given after 12 weeks
in three of the mothers of the hypothyroid infants, at a time when the fetal thyroid
had already begun to concentrated iodine (97). Therefore, fetal hypothyroidism
is more likely to occur if 131 I treatment is given after 12 weeks and congenital
defects would be the major concern after 131 I therapy in the early first trimester.
Dosimetry studies could quantitate the actual fetal exposure, but experts have
suggested that the “relatively low fetal whole-body irradiation is probably not
sufficient to justify termination of pregnancy.” (96). If inadvertent 131 I therapy
has been administered to a pregnant woman, PTU therapy may be initiated within
seven days of 131 I, which may reduce 131 I recycling by the fetal thyroid, thereby
lowering radiation exposure (96). All infants exposed to maternal 131 I therapy need
to be evaluated immediately at birth with institution of levothyroxine therapy if
congenital hypothyroidism is documented. The therapeutic administration of 131 I
to a nursing mother is contraindicated and lactation should be stopped immediately
if this occurs.

Lactation
Traditionally, many texts have advised against breast-feeding in women treated
with ATDs because of the presumption that the ATD was present in breast milk in
concentrations sufficient to affect the infant’s thyroid. However, over the last two
decades, several studies have prospectively monitored thyroid function in infants
nursed by mothers taking ATD therapy. PTU is more tightly protein bound than
MMI; consequently, the ratio of milk to serum levels is lower for PTU (0.67)
(98) than for MMI (1.0) (99). In addition, the amount of ingested drug secreted
in breast milk is approximately six times higher for MMI than for PTU (0.14 vs.
0.025% of the ingested dose) (98).
Thyroid Disease and Pregnancy 387

Several studies reported no alteration in thyroid function in a total of 56


newborns breast-fed by mothers treated with daily doses of PTU (50–300 mg),
MMI (5–20 mg), or carbimazole (5–15 mg) for periods ranging from three weeks to
eight months Even in women who were overtreated and developed elevated serum
TSH levels, the babies’ thyroid function tests remained normal (100). Therefore,
ATD therapy (PTU ⬍300 mg/day, MMI ⬍20 mg/day) may be considered during
lactation, although the number of reported infants is small. PTU would be preferred
because of its decreased appearance in breast milk, and the drug should be taken
by the mother after a feeding. In addition, the theoretic possibility of the infant
developing ATD side-effects via ATD ingestion through lactation has not been
reported (101).

Fetal/Neonatal Hyperthyroidism
In either women with active Graves’ disease or those with radioiodine ablated,
surgically treated Graves’ disease, fetal or neonatal hyperthyroidism is reported to
occur in 1% of pregnancies (102). However, there are no data evaluating whether
the incidence rate is higher in those with active versus “thyrodectomized” Graves’
disease. This disorder is caused by the transplacental passage of maternal TRAb
that stimulate fetal thyroid hormone production and may cause goiter. Maternal to
fetal IgG transport becomes clinically significant at the end of the second trimester,
which is when fetal hyperthyroidism usually becomes apparent. Measurement of
maternal TSI or TBII at 26 to 28 weeks provides prognostic information about
the development of fetal Graves’ disease (55,64). However, the measurement of
these antibody levels is not standardized and may depend upon the individual
laboratory’s reference range. Therefore, the magnitude of the TRAb elevation
compared to that assay’s normal range may be more practical for prediction of
fetal/neonatal hyperthyroidism (Table 5) (55,57,64,73,81,103–106).

Table 5 Prediction of Fetal/Neonatal Graves’ Disease


TRAb methodology Fold (x) above upper normal limit

Zakarija 1983 (57) TSI 5


Matsuura 1988 (104) TBII 3
TSI 3.5
Mortimer 1990 (73) TBII 5
Mitsuda 1990 (64) TBII 3
Smith 2001 (105) TBII 3.5
Peleg 2002 (106) TSI 5
Nachum 2003 (81) TSI 2
Luton 2005 (55) TBII 3

Abbreviations: TRAb, TSH receptor antibody; TSI, thyroid stimulating immunoglobuin; TBII, thyroid
binding inhibitory immunoglobulin.
388 Mandel

Table 6 TRAb Measurement and Fetal Ultrasound to Predict Fetal Thyroid


Dysfunction(55,103)

Graves’ disease
Graves’ disease on Graves’ disease after 131 I in remission on
ATD therapy therapy or thyroidectomy no therapy

Maternal TRAb YES YES NO


measurement
Timing of TRAb Early 3rd trimester End of 1st trimester
measurement if negative: no repeat
if positive: repeat
early 3rd trimester

If mother is TRAb + or taking ATD → Fetal ultrasound at 28–32 wks to check for fetal goiter,
tachycardia, and bone growth.
Abbreviations: TRAb, TSH receptor antibody; ATD, antithyroid drug.

Fetal thyroid ultrasound at 32 weeks in screening for clinically relevant fetal


thyroid dysfunction has a reported sensitivity of 92% and a specificity of 100%
(55). However, if a fetal goiter is detected, fetal hyper- and hypothyroidism must
be differentiated. Signs suggestive of fetal hyperthyroidism include intrauterine
growth retardation, arrhythmias, congestive heart failure, advanced bone age,
craniosynostosis, and hydrops (80,81). Another suspicious feature is a diffuse
Doppler ultrasound signal throughout the thyroid gland (55). Tachycardia (⬎ 160
beat per minute) may indicate, but is not always present in, fetal thyrotoxicosis (55).
If necessary, periumbilical blood sampling will confirm the diagnosis. Although
there are no established normal ranges for fetal thyroid hormone levels, the serum
TSH is greater than 20 mU/L in the published cases of fetal hypothyroidism
(82) or the serum T4 level is markedly elevated in those of fetal hyperthyroidism
(107,108). Periumbilical blood sampling has risks of fetal bleeding, bradycardia,
infection, and death (82) and is usually not indicated as the diagnosis can often
be made clinically. A rational approach for the detection of fetal and neonatal
hyperthyroidism using TRAb measurement and fetal ultrasound is presented in
Table 6.
Treatment of fetal hyperthyroidism is accomplished by giving the mother
ATD therapy, which then crosses the placenta and inhibits fetal thyroid hormone
synthesis. Hypothetically, because MMI is less protein bound, it might more
readily cross the placenta and have a greater effect on fetal thyroid function.
Generally, in the United States, PTU has been used in cases of suspected fetal
hyperthyroidism, and can be initiated at doses of 150 mg/day, with subsequent
normalization of fetal heart rate within two weeks (78). The dose can then be
titrated to maintain a normal fetal heart rate. In a woman who has received prior
radioablation or surgery for Graves’ disease, levothyroxine therapy may have to
be initiated or increased if maternal hypothyroxinemia occurs. Since TSI and TBII
Thyroid Disease and Pregnancy 389

levels usually decline toward term, ATD dosage can be decreased as fetal heart
rate and growth are monitored.
If antibody levels remain elevated at term, the risk of neonatal hyperthy-
roidism is increased (Table 6). Cord blood should be assayed for serum TSH and
FT4 levels. However, if these are normal and the mother has been taking ATD
at term, the baby’s thyroid function tests should be rechecked at four to six days
of life to dectect delayed hyperthyroidism because the clinical manifestations of
hyperthyroidism in the newborn may be masked for the first week of life until
maternal ATD dissipate (103,109). ATD therapy is necessary for treatment of
neonatal Graves’ disease. As maternal antibody levels decrease over the first three
months of life, therapy can usually be discontinued (109).

THYROID NODULES AND THYROID CANCER


Although goiter may occur during pregnancy in areas of borderline iodine defi-
ciency (110), thyroid size does not increase in iodine replete areas (111). The
prevalence of nodules has been reported to be higher in middle-aged women with
a history of three or more prior pregnancies (112), but solitary thyroid nodules
or thyroid cancer do not arise de novo more frequently during pregnancy. The
detection of thyroid nodules in pregnant women usually reflects the careful exam-
ination by an obstetrician of a healthy pregnant woman who had not regularly seen
a physician prior to her conception.
As in the nonpregnant patient (see chapter 5 “Thyroid Nodules and Multin-
odular Goiter”), diagnostic thyroid ultrasound for nodule characterization and fine-
needle aspiration (FNA) as indicated should be performed for diagnosis in nodules
identified during pregnancy (113); the spectrum of cytologic results is the same as
in the nonpregnant patient (114). Radioactive iodine scanning is contraindicated
at this time, but ultrasound can be performed either for characterization and mon-
itoring of the nodule or for guidance of the FNA. Nodules with a benign cytology
may be observed, and those with a cytology suggestive of follicular neoplasm may
require further evaluation and probable surgery after delivery.
If the FNA cytology shows evidence of a thyroid cancer, surgery is rec-
ommended. However, the timing of surgery, either during or after pregnancy, is
debatable with some recommending surgery during the midtrimester (114) and
others advocating waiting until after delivery (115). A recent study has reported no
significant difference in recurrence or survival rates between women with malig-
nant nodules who had surgery either during or after pregnancy (116). Furthermore,
thyroid cancer discovered during pregnancy is not more aggressive than that found
in a similar aged group of nonpregnant women (115). For women with previously
treated thyroid cancer, most reports confirm that subsequent pregnancy does not
increase recurrence rates (117).
For a nodule with malignant cytology, the recently published American
Thyroid Association guidelines recommend that “if discovered early in pregnancy,
[it] should be monitored sonographically and if it grows substantially by 24 weeks’
390 Mandel

gestation, surgery should be performed at that point. However, if it remains stable


by midgestation or if it is diagnosed in the second half of pregnancy, surgery may
be performed after delivery.” (113) In such women, levothyroxine therapy may be
considered to maintain the serum TSH level in the subnormal, but detectable range
(0.1–0.3 mU/L), which would theoretically slow TSH-responsive tumor growth
and should not be associated with pregnancy complications. Radioactive iodine
therapy, if needed, must be delayed until the postpartum period.

POSTPARTUM THYROIDITIS
PPT is a painless, destructive inflammation of the thyroid characterized by transient
hyperthyroidism and/or hypothyroidism that occurs within the first 6 to 12 months
after delivery in women who were euthyroid during pregnancy. It is reported
to occur in 2% to 17% of women. The discrepancies in prevalence may reflect
differences in diagnostic criteria, as well as variable predisposing genetic factors
and iodine intake that differ among screened populations (118).
PPT is thought to be an autoimmune disorder for several reasons. It is
temporally related to the time of immunologic “rebound” that occurs in the first
months after delivery (119). On cytology, lymphocytic thyroiditis is evident and
similar HLA haplotypes are found in patients with PPT and Hashimoto’s thy-
roiditis (118). Several studies have documented a correlation between antithyroid
antibody positivity and the development of PPT. The presence of antithyroid per-
oxidase antibodies in the first trimester is associated with a 33% risk of developing
PPT. Women who develop PPT have consistently higher titers of antithyroid anti-
bodies throughout gestation, compared to women who have detectable antibodies
but do not develop PPT (120). If antithyroid antibodies are detectable two days
postpartum, PPT is reported to affect 67% of women (121). There may also be a
contribution of cellular immunity to the development of PPT. The ratio of helper
to suppresser T-cells declines progressively during pregnancy followed by an
increase postpartum. Although women who develop PPT manifest a decline in the
helper/suppresser ratio during pregnancy, the ratio is higher compared to unaf-
fected women (120). In addition, women with type 1 diabetes mellitus, another
autoimmune disease, have an increased incidence of PPT compared to the gen-
eral population (122). Lastly, euthyroid patients with a history of Graves’ disease
may develop PPT initially after delivery, which may subsequently be followed by
Graves’ hyperthyroidism (60).
The thyrotoxic phase of PPT thyroiditis generally occurs within the first one
to four months after delivery and is transient, lasting for one to two months. There
is release of stored thyroid hormone from the damaged thyroid gland, rather than
increased thyroidal production of thyroid hormone. Hypothyroidism may follow
and last from two to six months until follicular cell synthetic activity recovers and
euthyroidism returns. Not all patients experience both hyper- and hypothyroidism.
In fact, either transient hyper- or hypothyroidism alone is observed in the majority
of reported patients (123). However, this may reflect differences in screening
protocols and the monitoring schedule of postpartum thyroid function.
Thyroid Disease and Pregnancy 391

The majority of patients recover from the hypothyroid phase of PPT. How-
ever, up to 10% of patients may have persistent hypothyroidism and 20% may later
develop permanent hypothyroidism over the next 2 to 10 years (124). The risk
of permanent hypothyroidism is correlated with the severity of the hypothyroid
phase of PPT and the elevation of the serum TSH level as well as the elevation in
antithyroid antibody levels (125). In addition, for those women who recover, there
is a 70% risk of developing recurrent PPT after a subsequent pregnancy (126).
There are no data addressing the future risk to these women of developing “silent”
autoimmune thyroiditis, not associated with pregnancy.

Diagnosis
The clinical manifestations of the thyrotoxic phase of PPT are not usually severe.
Fatigue and emotional lability may be prominent features, both of which are often
attributed to the postpartum state itself. Fatigue is the most consistent finding
during the hypothyroid phase, but again is a nonspecific symptom. When compared
to postpartum euthyroid women, hypothyroid patients with PPT have significantly
increased symptoms of impaired concentration and memory, carelessness, and
depression (121). Among women diagnosed with postpartum depression, PPT
is present twice as often as in women without depression (127). In addition, in
one study depressive symptoms in the postpartum period were associated with
positive antithyroid antibody status even without thyroid dysfunction (128). On
examination, a painless goiter is often detected.
The clinician should consider the diagnosis of PPT in any postpartum woman
who presents with a goiter and/or the nonspecific symptoms of fatigue, emotional
lability, or palpitations. The serum TSH concentration will be suppressed during
the thyrotoxic phase and elevated during the hypothyroid phase. However, in the
transition period from thyrotoxicosis to hypothyroidism, changes in the serum
TSH level may lag behind the decline in thyroid hormone levels, and may there-
fore be in the normal range when serum thyroid hormone concentrations are low.
Consequently, when assessing patients for PPT, it is important to obtain measure-
ments of both serum TSH and FT4 levels. Thyroid receptor antibody levels are
usually elevated in patients with Graves’ disease and not in PPT, unless a patient
has a prior history of Graves’ disease (60).
During the thyrotoxic phase, a radioiodine uptake (contraindicated in preg-
nant women) is low, differentiating hyperthyroidism in PPT caused by leakage of
stored thyroid hormone from the postpartum presentation of Graves’ disease with
increased hormone production. If a patient is nursing and an uptake is considered
critical for diagnosis, radiation safety and nuclear medicine should be consulted.
123
I is excreted in breast milk and has an effective half-life of five to eight hours.
An radioiodine uptake with 123 I (scan is not necessary) can be performed after
counseling and nursing should be stopped for at least 48 hours after 123 I adminis-
tration (129). Afterwards, she may bring in an aliquot of milk for assessment of
any residual radioactivity and at that time, lactation can usually be resumed. If the
patient has entered the recovery phase of PPT at the time of the uptake evaluation,
392 Mandel

the uptake may not be low as the serum TSH has now normalized or become
elevated.
The clinician should also be aware that PPT may occur in two other
settings. PPT may develop in women after a miscarriage or therapeutic abortion,
even if the pregnancy loss occurs as early as 5 weeks gestation (54). In addition,
women previously diagnosed with subclinical hypothyroidism prior to conception
and treated with subreplacement levothyroxine doses (0.025–0.075 mg/day) to
normalize the serum TSH level may not require a dosage increase in pregnancy,
which indicates that the residual thyroid is capable of functioning and synthesizing
thyroid hormone. After delivery, the same residual thyroid is subject to the
immunologic insult that causes PPT. These women will develop a suppressed
serum TSH level during the thyrotoxic phase and levothyroxine should be
stopped. However, during the hypothyroid phase, they may require a larger dose
than their previous subreplacement dose to restore euthyroidism, since they have
transiently lost thyroid hormone secretory capacity from the residual functioning
thyroid (130).

Treatment
Treatment is usually not required during the thyrotoxic phase of PPT. Generally,
by the time a patient recognizes the symptoms and seeks medical attention, thyroid
function normalizes. ATD are not beneficial because the increase in circulating
thyroid hormone is a result of a destructive process affecting the thyroid gland,
not increased thyroidal synthesis. If palpitations are a prominent symptom, beta-
adrenergic blocking drugs may be used transiently. Diagnosis of the hyperthyroid
phase of PPT is important because it identifies a group of women who likely will
develop hypothyroidism, which has a more prolonged duration. Levothyroxine
therapy can be initiated in patients with symptoms or TSH elevation ≥10 mU/L.
Often, women may be treated with a subreplacement dose (0.05–0.075 mg/day)
if the serum TSH level is between 10 and 20 mU/L, but may require a larger dose
if the serum TSH is ⬎20 mU/L. The serum TSH level should be rechecked at
four to six weeks, with dosage adjustment as needed. Since the hypothyroid phase
may last up to six months, levothyroxine therapy may be continued empirically
for several months after attaining a normal serum TSH level. At that point, it may
either be discontinued or reduced by half, with monitoring of a serum TSH level
three to four weeks later. If the serum TSH concentration is normal at that time, it
should be rechecked an additional time four to six weeks later if levothyroxine was
previously stopped or four to six weeks after discontinuation of the halved dose.
If the serum TSH level increases above normal on a lower dose or off therapy,
levothyroxine therapy is still required. Patients who recover from PPT should
have periodic annual monitoring of thyroid function. However, the presence of
autoimmune thyroid disease with positive antithyroid antibodies may confer a
higher risk of miscarriage for future pregnancies and reinitiating levothyroxine
therapy should be strongly considered prior to the next conception.
Thyroid Disease and Pregnancy 393

Consensus recommendations for screening for PPT are controversial and


universal screening cannot be justified currently (131). However, two populations,
women with type 1 diabetes and those with a prior history of PPT, have a suffi-
ciently high incidence of PPT that they might benefit from screening with TSH
measurements at 6-week, 3-month, and 6-month postpartum.

REFERENCES
1. Glinoer D, de Nayer P, Bourdoux P, et al. Regulation of maternal thyroid during
pregnancy. J Clin Endocrinol Metab 1990; 71(2):276–287.
2. Ain KB, Mori Y, Refetoff S. Reduced clearance rate of thyroxine-binding globulin
(TBG) with increased sialylation: A mechanism for estrogen-induced elevation of
serum TBG concentration. J Clin Endocrinol Metab 1987; 65(4):689–696.
3. Burrow GN, Fisher DA, Larsen PR. Maternal and fetal thyroid function. N Engl J
Med 1994; 331(16):1072–1078.
4. Demers LM, Spencer CA. Laboratory medicine practice guidelines: Laboratory
support for the diagnosis and monitoring of thyroid disease. Clin Endocrinol (Oxf)
2003; 58(2):138–140.
5. Abalovich M, Gutierrez S, Alcaraz G, et al. Overt and subclinical hypothyroidism
complicating pregnancy. Thyroid 2002; 12(1):63–68.
6. Mandel SJ, Larsen PR, Seely EW, et al. Increased need for thyroxine during preg-
nancy in women with primary hypothyroidism. N Engl J Med 1990; 323(2):91–
96.
7. Alexander EK, Marqusee E, Lawrence J, et al. Timing and magnitude of increases
in levothyroxine requirements during pregnancy in women with hypothyroidism.
N Engl J Med 2004; 351(3):241–249.
8. Kaplan MM. Assessment of thyroid function during pregnancy. Thyroid 1992;
2(1):57–61.
9. Glinoer D, Delange F, Laboureur I, et al. Maternal and neonatal thyroid function
at birth in an area of marginally low iodine intake. J Clin Endocrinol Metab 1992;
75(3):800–805.
10. Sapin R, D’Herbomez M, Schlienger JL. Free thyroxine measured with equilibrium
dialysis and nine immunoassays decreases in late pregnancy. Clin Lab 2004; 50(9–
10):581–584.
11. Mandel SJ, Spencer CA, Hollowell JG. Are detection and treatment of thyroid
insufficiency in pregnancy feasible? Thyroid 2005; 15(1):44–53.
12. Yoshimura M, Hershman JM. Thyrotropic action of human chorionic gonadotropin.
Thyroid 1995; 5(5):425–434.
13. Panesar NS, Li CY, Rogers MS. Reference intervals for thyroid hormones in pregnant
Chinese women. Ann Clin Biochem 2001; 38(Pt 4):329–332.
14. Casey BM, Dashe JS, Wells CE, et al. Subclinical hypothyroidism and pregnancy
outcomes. Obstet Gynecol 2005; 105(2):239–245.
15. Nelson M, Wickus GG, Caplan RH, et al. Thyroid gland size in pregnancy. An
ultrasound and clinical study. J Reprod Med 1987; 32(12):888–890.
16. Stagnaro-Green A, Roman SH, Cobin RH, et al. Detection of at-risk pregnancy
by means of highly sensitive assays for thyroid autoantibodies. JAMA 1990;
264(11):1422–1425.
394 Mandel

17. Stagnaro-Green A, Chen X, Bogden JD, et al. The thyroid and pregnancy: A novel
risk factor for very preterm delivery. Thyroid 2005; 15(4):351–357.
18. Prummel MF, Wiersinga WM. Thyroid autoimmunity and miscarriage. Eur J
Endocrinol 2004; 150(6):751–755.
19. Pratt D, Novotny M, Kaberlein G, et al. Antithyroid antibodies and the association
with non-organ-specific antibodies in recurrent pregnancy loss. Am J Obstet Gynecol
1993; 168(3 Pt 1):837–841.
20. Negro R, Formoso G, Mangieri T, et al. Levothyroxine treatment in euthyroid preg-
nant women with autoimmune thyroid disease: Effects on obstetrical complications.
J Clin Endocrinol Metab 2006; 91(7):2587–2591.
21. Glinoer D, Riahi M, Grun JP, et al. Risk of subclinical hypothyroidism in pregnant
women with asymptomatic autoimmune thyroid disorders. J Clin Endocrinol Metab
1994; 79(1):197–204.
22. Montoro MN. Management of hypothyroidism during pregnancy. Clin Obstet
Gynecol 1997; 40(1):65–80.
23. Klein RZ, Haddow JE, Faix JD, et al. Prevalence of thyroid deficiency in pregnant
women. Clin Endocrinol (Oxf) 1991; 35(1):41–46.
24. Montoro M, Collea JV, Frasier SD, et al. Successful outcome of pregnancy in women
with hypothyroidism. Ann Intern Med 1981; 94(1):31–34.
25. Abalovich M, Amino N, Barbour LA, et al. Management of thyroid dysfunction
during pregnancy and postpartum: An Endocrine Society Clinical Practice Guideline.
J Clin Endocrinol Metab 2007; 92(8 suppl):S1–S47.
26. Jovanovic-Peterson L, Peterson CM. De novo clinical hypothyroidism in pregnancies
complicated by type I diabetes, subclinical hypothyroidism, and proteinuria: A new
syndrome. Am J Obstet Gynecol 1988; 159(2):442–446.
27. Vaidya B, Anthony S, Bilous M, et al. Detection of thyroid dysfunction in early
pregnancy: Universal screening or targeted high-risk case finding? J Clin Endocrinol
Metab 2007; 92(1):203–207.
28. Allan WC, Haddow JE, Palomaki GE, et al. Maternal thyroid deficiency and preg-
nancy complications: Implications for population screening. J Med Screen 2000;
7(3):127–130.
29. Leung AS, Millar LK, Koonings PP, et al. Perinatal outcome in hypothyroid preg-
nancies. Obstet Gynecol 1993; 81(3):349–353.
30. Davis LE, Leveno KJ, Cunningham FG. Hypothyroidism complicating pregnancy.
Obstet Gynecol 1988; 72(1):108–112.
31. Wasserstrum N, Anania CA. Perinatal consequences of maternal hypothyroidism
in early pregnancy and inadequate replacement. Clin Endocrinol (Oxf) 1995;
42(4):353–358.
32. Glinoer D. The regulation of thyroid function in pregnancy: Pathways of
endocrine adaptation from physiology to pathology. Endocr Rev 1997; 18(3):404–
433.
33. Rotondi M, Caccavale C, Di Serio C, et al. Successful outcome of pregnancy in a
thyroidectomized-parathyroidectomized young woman affected by severe hypothy-
roidism. Thyroid 1999; 9(10):1037–140.
34. Vulsma T, Gons MH, de Vijlder JJ. Maternal-fetal transfer of thyroxine in congenital
hypothyroidism due to a total organification defect or thyroid agenesis. N Engl J Med
1989; 321(1):13–16.
Thyroid Disease and Pregnancy 395

35. Calvo RM, Jauniaux E, Gulbis B, et al. Fetal tissues are exposed to biologically
relevant free thyroxine concentrations during early phases of development. J Clin
Endocrinol Metab 2002; 87(4):1768–1777.
36. Porterfield SP, Hendrich CE. The role of thyroid hormones in prenatal and neona-
tal neurological development–current perspectives. Endocr Rev 1993; 14(1):94–
106.
37. Bonet B, Herrera E. Different response to maternal hypothyroidism during the
first and second half of gestation in the rat. Endocrinology 1988; 122(2):450–
455.
38. Man EB, Brown JF, Serunian SA. Maternal hypothyroxinemia: Psychoneurological
deficits of progeny. Ann Clin Lab Sci 1991; 21(4):227–239.
39. Haddow JE, Palomaki GE, Allan WC, et al. Maternal thyroid deficiency during
pregnancy and subsequent neuropsychological development of the child. N Engl J
Med 1999; 341(8):549–555.
40. Klein RZ, Sargent JD, Larsen PR, et al. Relation of severity of maternal hypothy-
roidism to cognitive development of offspring. J Med Screen 2001; 8(1):18–20.
41. Brown RS, Bellisario RL, Botero D, et al. Incidence of transient congenital hypothy-
roidism due to maternal thyrotropin receptor-blocking antibodies in over one million
babies. J Clin Endocrinol Metab 1996; 81(3):1147–1151.
42. Mandel SJ. Hypothyroidism and chronic autoimmune thyroiditis in the pregnant
state: Maternal aspects. Best Pract Res Clin Endocrinol Metab 2004; 18(2):213–
224.
43. Campbell NR, Hasinoff BB, Stalts H, et al. Ferrous sulfate reduces thyroxine efficacy
in patients with hypothyroidism. Ann Intern Med 1992; 117(12):1010–1013.
44. Kimura M, Amino N, Tamaki H, et al. Gestational thyrotoxicosis and hyperemesis
gravidarum: Possible role of hCG with higher stimulating activity. Clin Endocrinol
(Oxf) 1993; 38(4):345–350.
45. Glinoer D. Thyroid hyperfunction during pregnancy. Thyroid 1998; 8(9):859–864.
46. Glinoer D, De Nayer P, Robyn C, et al. Serum levels of intact human chorionic
gonadotropin (HCG) and its free alpha and beta subunits, in relation to maternal
thyroid stimulation during normal pregnancy. J Endocrinol Invest 1993; 16(11):881–
888.
47. Goodwin TM, Montoro M, Mestman JH. Transient hyperthyroidism and hyper-
emesis gravidarum: Clinical aspects. Am J Obstet Gynecol 1992; 167(3):648–
652.
48. Mori M, Amino N, Tamaki H, et al. Morning sickness and thyroid function in normal
pregnancy. Obstet Gynecol 1988; 72(3 Pt 1):355–359.
49. Goodwin TM, Montoro M, Mestman JH, et al. The role of chorionic gonadotropin
in transient hyperthyroidism of hyperemesis gravidarum. J Clin Endocrinol Metab
1992; 75(5):1333–1337.
50. Rodien P, Bremont C, Sanson ML, et al. Familial gestational hyperthyroidism caused
by a mutant thyrotropin receptor hypersensitive to human chorionic gonadotropin.
N Engl J Med 1998; 339(25):1823–1826.
51. Tan JY, Loh KC, Yeo GS, et al. Transient hyperthyroidism of hyperemesis gravi-
darum. BJOG 2002; 109(6):683–688.
52. Goodwin TM, Hershman JM. Hyperthyroidism due to inappropriate production of
human chorionic gonadotropin. Clin Obstet Gynecol 1997; 40(1):32–44.
396 Mandel

53. Millar LK, Wing DA, Leung AS, et al. Low birth weight and preeclampsia in
pregnancies complicated by hyperthyroidism. Obstet Gynecol 1994; 84(6):946–949.
54. Marqusee E, Hill JA, Mandel SJ. Thyroiditis after pregnancy loss. J Clin Endocrinol
Metab 1997; 82(8):2455–2457.
55. Luton D, Le Gac I, Vuillard E, et al. Management of Graves’ disease during preg-
nancy: The key role of fetal thyroid gland monitoring. J Clin Endocrinol Metab
2005; 90(11):6093–6098.
56. Amino N, Izumi Y, Hidaka Y, et al. No increase of blocking type anti-thyrotropin
receptor antibodies during pregnancy in patients with Graves’ disease. J Clin
Endocrinol Metab 2003; 88(12):5871–5874.
57. Zakarija M, McKenzie JM. Pregnancy-associated changes in the thyroid-stimulating
antibody of Graves’ disease and the relationship to neonatal hyperthyroidism. J Clin
Endocrinol Metab 1983; 57(5):1036–1040.
58. Amino N, Tanizawa O, Mori H, et al. Aggravation of thyrotoxicosis in early
pregnancy and after delivery in Graves’ disease. J Clin Endocrinol Metab 1982;
55(1):108–112.
59. Chan GW, Mandel SJ. Therapy insight: Management of Graves’ disease during
pregnancy. Nat Clin Pract Endocrinol Metab 2007; 3(6):470–478.
60. Momotani N, Noh J, Ishikawa N, et al. Relationship between silent thyroiditis and
recurrent Graves’ disease in the postpartum period. J Clin Endocrinol Metab 1994;
79(1):285–289.
61. Gardiner-Hill H. Pregnancy complicating simple goitre and Graves’ disease. Lancet
1929; 1:120–124.
62. Mestman JH, Manning PR, Hodgman J. Hyperthyroidism and pregnancy. Arch Intern
Med 1974; 134(3):434–439.
63. Davis LE, Lucas MJ, Hankins GD, et al. Thyrotoxicosis complicating pregnancy.
Am J Obstet Gynecol 1989; 160(1):63–70.
64. Mitsuda N, Tamaki H, Amino N, et al. Risk factors for developmental disorders in
infants born to women with Graves disease. Obstet Gynecol 1992; 80(3 Pt 1):359–
364.
65. Sugrue D, Drury MI. Hyperthyroidism complicating pregnancy: Results of treatment
by antithyroid drugs in 77 pregnancies. Br J Obstet Gynaecol 1980; 87(11):970–975.
66. Momotani N, Ito K, Hamada N, et al. Maternal hyperthyroidism and congenital
malformation in the offspring. Clin Endocrinol (Oxf) 1984; 20(6):695–700.
67. Casey BM, Dashe JS, Wells CE, et al. Subclinical hyperthyroidism and pregnancy
outcomes. Obstet Gynecol 2006; 107(2 Pt 1):337–341.
68. Mandel SJ, Brent GA, Larsen PR. Review of antithyroid drug use during pregnancy
and report of a case of aplasia cutis. Thyroid 1994; 4(1):129–133.
69. Marchant B, Brownlie BE, Hart DM, et al. The placental transfer of propylthiouracil,
methimazole and carbimazole. J Clin Endocrinol Metab 1977; 45(6):1187–1193.
70. Gardner DF, Cruikshank DP, Hays PM, et al. Pharmacology of propylthiouracil
(PTU) in pregnant hyperthyroid women: Correlation of maternal PTU concentrations
with cord serum thyroid function tests. J Clin Endocrinol Metab 1986; 62(1):217–
220.
71. Mortimer RH, Cannell GR, Addison RS, et al. Methimazole and propylthiouracil
equally cross the perfused human term placental lobule. J Clin Endocrinol Metab
1997; 82(9):3099–3102.
Thyroid Disease and Pregnancy 397

72. Davies TF, Roti E, Braverman LE, et al. Thyroid controversy–stimulating antibodies.
J Clin Endocrinol Metab 1998; 83(11):3777–3785.
73. Mortimer RH, Tyack SA, Galligan JP, et al. Graves’ disease in pregnancy: TSH
receptor binding inhibiting immunoglobulins and maternal and neonatal thyroid
function. Clin Endocrinol (Oxf) 1990; 32(2):141–1452.
74. Cheron RG, Kaplan MM, Larsen PR, et al. Neonatal thyroid function after propy-
lthiouracil therapy for maternal Graves’ disease. N Engl J Med 1981; 304(9):525–
528.
75. Momotani N, Noh J, Oyanagi H, et al. Antithyroid drug therapy for Graves’ disease
during pregnancy. Optimal regimen for fetal thyroid status. N Engl J Med 1986;
315(1):24–28.
76. Momotani N, Noh JY, Ishikawa N, et al. Effects of propylthiouracil and methimazole
on fetal thyroid status in mothers with Graves’ hyperthyroidism. J Clin Endocrinol
Metab 1997; 82(11):3633–3636.
77. Cooper DS, Saxe VC, Meskell M, et al. Acute effects of propylthiouracil (PTU)
on thyroidal iodide organification and peripheral iodothyronine deiodination: Cor-
relation with serum PTU levels measured by radioimmunoassay. J Clin Endocrinol
Metab 1982; 54(1):101–107.
78. Fisher DA. Fetal thyroid function: Diagnosis and management of fetal thyroid dis-
orders. Clin Obstet Gynecol 1997; 40(1):16–31.
79. Momotani N, Iwama S, Noh J, et al. Anti-thyroid drug therapy for Graves’ dis-
ease during pregnancy: Mildest thyrotoxic maternal free thyroxine concentrations
to avoid fetal hypothyroidism. In: 77th Annual Meeting of the American Thyroid
Association.; 2006; Phoenix, AZ.
80. Chopra IJ. Fetal and neonatal hyperthyroidism. Thyroid 1992; 2(2):161–163.
81. Nachum Z, Rakover Y, Weiner E, et al. Graves’ disease in pregnancy: Prospective
evaluation of a selective invasive treatment protocol. Am J Obstet Gynecol 2003;
189(1):159–165.
82. Davidson KM, Richards DS, Schatz DA, et al. Successful in utero treatment of fetal
goiter and hypothyroidism. N Engl J Med 1991; 324(8):543–546.
83. Van Loon AJ, Derksen JT, Bos AF, et al. In utero diagnosis and treatment of fetal
goitrous hypothyroidism, caused by maternal use of propylthiouracil. Prenat Diagn
1995; 15(7):599–604.
84. Ochoa-Maya MR, Frates MC, Lee-Parritz A, et al. Resolution of fetal goiter after dis-
continuation of propylthiouracil in a pregnant woman with Graves’ hyperthyroidism.
Thyroid 1999; 9(11):1111–1114.
85. McCarroll AM, Hutchinson M, McAuley R, et al. Long-term assessment of children
exposed in utero to carbimazole. Arch Dis Child 1976; 51(7):532–536.
86. Burrow GN, Klatskin EH, Genel M. Intellectual development in children whose
mothers received propylthiouracil during pregnancy. Yale J Biol Med 1978;
51(2):151–156.
87. Messer PM, Hauffa BP, Olbricht T, et al. Antithyroid drug treatment of Graves’
disease in pregnancy: Long-term effects on somatic growth, intellectual development
and thyroid function of the offspring. Acta Endocrinol (Copenh) 1990; 123(3):311–
316.
88. Eisenstein Z, Weiss M, Katz Y, et al. Intellectual capacity of subjects exposed to
methimazole or propylthiouracil in utero. Eur J Pediatr 1992; 151(8):558–559.
398 Mandel

89. Greenberg F. Brief clinical report: Choanal atresia and athelia: Methimazole terato-
genicity or a new syndrome? Am J Med Gen 1987; 28:931–934.
90. Casey BM. Subclinical hypothyroidism and pregnancy. Obstet Gynecol Surv 2006;
61(6):415–420; quiz 23.
91. Wing DA, Millar LK, Koonings PP, et al. A comparison of propylthiouracil versus
methimazole in the treatment of hyperthyroidism in pregnancy. Am J Obstet Gynecol
1994; 170(1 Pt 1):90–95.
92. Mestman JH. Hyperthyroidism in pregnancy. Clin Obstet Gynecol 1997; 40(1):45–
64.
93. Sherif IH, Oyan WT, Bosairi S, et al. Treatment of hyperthyroidism in pregnancy.
Acta Obstet Gynecol Scand 1991; 70(6):461–463.
94. Momotani N, Hisaoka T, Noh J, et al. Effects of iodine on thyroid status of fetus
versus mother in treatment of Graves’ disease complicated by pregnancy. J Clin
Endocrinol Metab 1992; 75(3):738–744.
95. Brodsky JB, Cohen EN, Brown BW Jr, et al. Surgery during pregnancy and fetal
outcome. Am J Obstet Gynecol 1980; 138(8):1165–1167.
96. Masiukiewicz US, Burrow GN. Hyperthyroidism in pregnancy: Diagnosis and treat-
ment. Thyroid 1999; 9(7):647–652.
97. Stoffer SS, Hamburger JI. Inadvertent 131 I therapy for hyperthyroidism in the first
trimester of pregnancy. J Nucl Med 1976; 17(02):146–149.
98. Kampmann JP, Johansen K, Hansen JM, et al. Propylthiouracil in human milk.
Revision of a dogma. Lancet 1980; 1(8171):736–737.
99. Johansen K, Andersen AN, Kampmann JP, et al. Excretion of methimazole in human
milk. Eur J Clin Pharmacol 1982; 23(4):339–341.
100. Azizi F, Khoshniat M, Bahrainian M, et al. Thyroid function and intellectual devel-
opment of infants nursed by mothers taking methimazole. J Clin Endocrinol Metab
2000; 85(9):3233–3238.
101. Mandel SJ, Cooper DS. The use of antithyroid drugs in pregnancy and lactation.
J Clin Endocrinol Metab 2001; 86(6):2354–2359.
102. Burrow GN. Thyroid function and hyperfunction during gestation. Endocr Rev 1993;
14(2):194–202.
103. Laurberg P, Nygaard B, Glinoer D, et al. Guidelines for TSH-receptor anti-
body measurements in pregnancy: Results of an evidence-based symposium orga-
nized by the European Thyroid Association. Eur J Endocrinol 1998; 139(6):584–
586.
104. Matsuura N, Konishi J, Fujieda K, et al. TSH-receptor antibodies in mothers with
Graves’ disease and outcome in their offspring. Lancet 1988; 1(8575–8576):14–
17.
105. Smith C, Thomsett M, Choong C, et al. Congenital thyrotoxicosis in premature
infants. Clin Endocrinol (Oxf) 2001; 54(3):371–376.
106. Peleg D, Cada S, Peleg A, et al. The relationship between maternal serum thyroid-
stimulating immunoglobulin and fetal and neonatal thyrotoxicosis. Obstet Gynecol
2002; 99(6):1040–1043.
107. Porreco RP, Bloch CA. Fetal blood sampling in the management of intrauterine
thyrotoxicosis. Obstet Gynecol 1990; 76(3 Pt 2):509–512.
108. Wenstrom KD, Weiner CP, Williamson RA, et al. Prenatal diagnosis of fetal hyper-
thyroidism using funipuncture. Obstet Gynecol 1990; 76(3 Pt 2):513–517.
Thyroid Disease and Pregnancy 399

109. Skuza KA, Sills IN, Stene M, et al. Prediction of neonatal hyperthyroidism in infants
born to mothers with Graves disease. J Pediatr 1996; 128(2):264–268.
110. Rasmussen NG, Hornnes PJ, Hegedus L. Ultrasonographically determined thyroid
size in pregnancy and post partum: The goitrogenic effect of pregnancy. Am J Obstet
Gynecol 1989; 160(5 Pt 1):1216–1220.
111. Berghout A, Endert E, Ross A, et al. Thyroid function and thyroid size in nor-
mal pregnant women living in an iodine replete area. Clin Endocrinol (Oxf) 1994;
41(3):375–379.
112. Struve CW, Haupt S, Ohlen S. Influence of frequency of previous pregnancies on
the prevalence of thyroid nodules in women without clinical evidence of thyroid
disease. Thyroid 1993; 3(1):7–9.
113. Cooper DS, Doherty GM, Haugen BR, et al. Management guidelines for patients
with thyroid nodules and differentiated thyroid cancer. Thyroid 2006; 16(2):109–
142.
114. Tan GH, Gharib H, Goellner JR, et al. Management of thyroid nodules in pregnancy.
Arch Intern Med 1996; 156(20):2317–2320.
115. Herzon FS, Morris DM, Segal MN, et al. Coexistent thyroid cancer and pregnancy.
Arch Otolaryngol Head Neck Surg 1994; 120(11):1191–1193.
116. Moosa M, Mazzaferri EL. Outcome of differentiated thyroid cancer diagnosed in
pregnant women. J Clin Endocrinol Metab 1997; 82(9):2862–2866.
117. Rosen IB, Korman M, Walfish PG. Thyroid nodular disease in pregnancy: Current
diagnosis and management. Clin Obstet Gynecol 1997; 40(1):81–89.
118. Browne-Martin K, Emerson CH. Postpartum thyroid dysfunction. Clin Obstet
Gynecol 1997; 40(1):90–101.
119. Roti E, Emerson CH. Clinical Review 92 postpartum thyroiditis. J Clin Endocrinol
Metab 1992; 74:3–7.
120. Stagnaro-Green A, Roman SH, Cobin RH, et al. A prospective study of lymphocyte-
initiated immunosuppression in normal pregnancy: Evidence of a T-cell etiology
for postpartum thyroid dysfunction. J Clin Endocrinol Metab 1992; 74(3):645–
653.
121. Hayslip CC, Fein HG, O’Donnell VM, et al. The value of serum antimicrosomal
antibody testing in screening for symptomatic postpartum thyroid dysfunction. Am
J Obstet Gynecol 1988; 159(1):203–209.
122. Gerstein HC. Incidence of postpartum thyroid dysfunction in patients with type I
diabetes mellitus. Ann Intern Med 1993; 118(6):419–423.
123. Solomon BL, Fein HG, Smallridge RC. Usefulness of antimicrosomal antibody
titers in the diagnosis and treatment of postpartum thyroiditis. J Fam Pract 1993;
36(2):177–1782.
124. Tachi J, Amino N, Tamaki H, et al. Long term follow-up and HLA association in
patients with postpartum hypothyroidism. J Clin Endocrinol Metab 1988; 66(3):480–
484.
125. Othman S, Phillips DI, Parkes AB, et al. A long-term follow-up of postpartum
thyroiditis. Clin Endocrinol (Oxf) 1990; 32(5):559–564.
126. Lazarus JH, Ammari F, Oretti R, et al. Clinical aspects of recurrent postpartum
thyroiditis. Br J Gen Pract 1997; 47(418):305–308.
127. Pop VJ, de Rooy HA, Vader HL, et al. Postpartum thyroid dysfunction and depression
in an unselected population. N Engl J Med 1991; 324(25):1815–1816.
400 Mandel

128. Harris B, Othman S, Davies JA, et al. Association between postpartum thyroid
dysfunction and thyroid antibodies and depression. BMJ 1992; 305(6846):152–
156.
129. Gorman CA. Radioiodine and pregnancy. Thyroid 1999; 9(7):721–726.
130. Mandel SJ. Postpartum thyroiditis in women with subclinical hypothyroidism. In:
Annual Meeting of the American Thyroid Association; 1998.
131. Amino N, Tada H, Hidaka Y, et al. Therapeutic controversy: Screening for postpartum
thyroiditis. J Clin Endocrinol Metab 1999; 84(6):1813–1821.
10
Thyroid Disease in
the Elderly

Anne R. Cappola
University of Pennsylvania School of Medicine, Philadelphia,
Pennsylvania, U.S.A.

Myron Miller and Steven R. Gambert


Sinai Hospital of Baltimore and The Johns Hopkins University School of
Medicine, Baltimore, Maryland, U.S.A.

INTRODUCTION
A 30-year-old woman presents with unintentional weight loss, difficulty concen-
trating, insomnia, and palpitations. An 80-year-old woman presents with “not
feeling right.” What is the diagnosis?
Both have overt hyperthyroidism. Yet in the elderly woman, the initial list
of differential diagnoses is far greater, the diagnostic testing likely to be more
extensive, and the time for diagnosis is potentially longer than in her young
counterpart. The 80-year-old woman may attribute some of the same symptoms—
the weight loss, difficulty concentrating, and insomnia—to old age. She is more
likely to be taking a medication, such as amiodarone, that could play a role in
the etiology of hyperthyroidism, and she is more likely to see an exacerbation
of a preexisting chronic disease, such as osteoporosis, in conjunction with the
hyperthyroidism. Even after diagnosis, her other medical problems will potentially
affect the older woman’s treatment course.
This chapter will begin with a summary of normal age-related changes in
the anatomy and function of the thyroid, followed by consideration of the factors
in the management of each type of thyroid disorder that are unique to the care of
older individuals.

401
402 Cappola et al.

NORMAL AGING AND THYROID FUNCTION


Morphology
The normal aging process is accompanied by changes in the gross and microscopic
appearance of the thyroid gland. Autopsy data have indicated that overall thyroid
gland mass declines with age from its normal range of 15 to 25 g; with increasing
age, a progressively larger proportion of individuals will have glands weighing
⬍20 g (1). Other confounding factors were considered in a study of thyroid ultra-
sounds in healthy subjects (2). This study showed that effects of aging on the size
of the thyroid gland were quite small, and that body weight was a greater deter-
minant of thyroid volume than was age, suggesting that thyroid size is maintained
with increasing age. With advancing age, progressive fibrosis, the appearance of
lymphocytes, a decrease in follicle size, and a reduction in the amount of colloid
in the thyroid gland are more likely to be seen (3). Although these changes are
common in the elderly, they are not characteristic of all aged persons. More impor-
tantly, there does not appear to be a decline in thyroid function as a result of the
morphological changes, and neither weight nor histological appearance correlates
with common measures of thyroid function (4).

Thyroid Hormone Physiology and Regulation


Table 1 details the effect of aging on thyroid hormone physiology. The changes in
thyroid function attributable solely to aging are subtle and of questionable clinical
significance. Instead, there is an increase in thyroid disease with increasing age.
Thus, there is little impact of aging on a normal thyroid, but there is an increase
in the incidence of thyroid disease with increasing age.

TRH–TSH Axis
A study of healthy elderly subjects reported a blunted increase in nocturnal thyroid-
stimulating hormone (TSH) secretion (5). This suggested a resetting of the pituitary
threshold of TSH feedback suppression, leading to a decrease in TSH secretion for
a given concentration of circulating thyroid hormone. The ability of the pituitary
gland to synthesize TSH does not appear to be diminished by the aging process,
as reflected by observation that pituitary TSH content undergoes no significant
change over the life span (6). The 24-hour TSH secretion has been reported to be
decreased in healthy elderly men (7), but the secretion rate of TSH has also been
reported to be higher in elderly subjects than in young individuals (8). Some have
postulated, nevertheless, that TSH secretion may decline slightly with age; the
slower clearance of T4 and T3 in older persons implies that less TSH is needed
to maintain thyroid secretion. Studies measuring the serum TSH response to
thyrotropin-releasing hormone (TRH) in elderly people are conflicting, with some
studies showing a decreased response and others showing an increased response
(7,9–14). In men aged 30 to 96 years given a continuous TRH infusion instead
Thyroid Disease in the Elderly 403

Table 1 Influence of Aging on Measures of Thyroid Function


Healthy elderly Healthy centenarians
Measure (65–96 yr)a (⬎100 yr)a

TSH N or Inc Dec


TSH response to TRH N or Dec ND
T4 N ND
T3 N or Dec ND
Free T4 N N
Free T3 N Dec
rT3 N Inc
TBG N ND
T4 synthesis/secretion Dec ND
T3 synthesis/secretion Dec ND
T4 half-life Inc Inc
Antithyroid antibodies Inc N
Thyroglobulin N ND
Radioiodine uptake N or Dec N or Dec
a Compared to values in young, healthy adults.

Abbreviations: N, no change; Inc, increased; Dec, decreased; ND, no data available.

of bolus TRH administration, the serum TSH response was biphasic, with neither
phase being affected by age (11).
TSH exerts its effects by binding to the membrane of thyroid follicular cells.
The serum T4 response to the TSH secreted after bolus TRH administration does
not appear to be altered by age, although the serum T3 response may decrease (15).
Administration of large doses of exogenous TSH has been reported to increase
serum T4 to a lesser extent in older subjects than in younger subjects (16).

TSH
The neuroendocrine mechanisms controlling TSH release may be altered during
the normal aging process. There is some disagreement regarding the effect of aging
on circulating levels of serum TSH. Recent analyses from the U.S. National Health
and Nutrition Examination Survey (NHANES) showed a shift in the distribution
of TSH levels toward higher levels with increasing age, even when individu-
als with positive thyroid peroxidase or thyroglobulin antibodies were excluded
(Fig. 1) (17). However, in an Italian study of healthy centenarians aged 100 to
110 years, serum TSH was lower in the centenarians compared to both healthy
elderly (aged 65–80 years) and healthy younger adults (aged 20–64 years) (median
0.97 mU/L vs. 1.17 mU/L vs. 1.70 mU/L) (18,19). For the older groups as a whole,
there was an inverse relationship between serum TSH and age. These disparate
findings may reflect differences in the iodine sufficiency of these two populations,
with a greater prevalence of Hashimoto’s thyroiditis in iodine sufficient countries
such as the United States and thyroid nodularity with autonomy in Italy (18,20,21).
404 Cappola et al.

(A) 30

25

Age 20–29
20
Age 50–59
Percent

Age 80+
15

10

0
0.1 0.2 0.3 0.4 0.6 0.9 1.4 2.1 3.1 4.7 7.1 10.7 16.1 24.2 36.5 55.1

(B) 30

25

20
Age 20–29
Age 50–59
Percent

15
Age 80+

10

0
0.1 0.2 0.3 0.4 0.6 0.9 1.4 2.1 3.1 4.7 7.1 10.7 16.1 24.2 36.5 55.1
Upper TSH concentration in bin

Figure 1 TSH distribution by age groups in the United States. (A) Disease-free population.
(B) Reference population, NHANES III (1988–1994). Source: From Ref. 17.

The lack of consensus on what a “normal” TSH level is for an older person in
turn leads to disparate recommendations for screening for thyroid disease, the
threshold for initiating therapy for thyroid dysfunction, and the target TSH level
during therapy. For example, if an upper limit of TSH of 2.5 mU/L was used
in the NHANES disease-free population, 39% of those over the age of 70 years
would be considered to have abnormally high TSH levels (17). Using the standard
reference cutoff of 2.5%, the upper limit of TSH would be ⬎9 mU/L in this age
group. The more clinically relevant approach is to link a specific TSH level to an
increase in adverse outcomes, as discussed in the sections on hypothyroidism and
hyperthyroidism later.
Thyroid Disease in the Elderly 405

T4 and T3
In older individuals, there is evidence for a decrease in thyroid hormone clearance
that is paralleled by a decrease in thyroid hormone secretion rate (18). As a result,
the normal range for total and free concentrations of T4 and T3 is unchanged in
older individuals, and deviations from normal must be considered as evidence for
thyroid disease or for other illness or states that may affect hormone measurement
(15,22–25). However, these age-related changes in thyroid hormone metabolism
have important effects on dosing of thyroid hormone replacement, with lower
dose requirements and longer times to achieve steady state in the older adult. In
the normal adult, approximately 80 ␮g of T4 and 30 ␮g of triiodothyronine (T3)
are produced daily (26). In elderly individuals, T4 and T3 production declines
to approximately 60 and 20 ␮g/day, respectively. These decreases in hormone
production may be related to the decrease in thyroidal iodide accumulation that
has been observed with aging (27). In addition, the half-life of thyroxine (T4)
increases with age, with a mean half-life of 6.7 days for adults aged 23 to 36
years and approximately 9 days for those older than 80 years (28). The decrease
in T4 clearance is thought to result from decreases in both the fractional turnover
rate and distribution space of T4. Levels of thyroid-binding globulin (TBG), the
primary thyroid hormone transport protein for T4 and T3, do not appear to differ
between healthy young and old individuals (11). While a greater T4-binding
capacity of TBG has been observed in the elderly, this does not have clinical
significance (29,30). Comorbid conditions or concomitant use of medications that
affect levels of thyroid-binding proteins may be more common in the elderly.
Levels of thyroid-binding prealbumin may be acutely lowered in the presence of
infectious disease, protein-wasting states, surgery, and malnutrition and may be
accompanied by a decline in total T4. TBG levels can be depressed as a result
of severe catabolic illness, chronic hepatic disease, glucocorticoids, and androgen
administration (31). Binding of T4 to TBG can be inhibited by drugs such as high-
dose salicylates and furosemide (32). TBG levels may increase due to therapy with
estrogen or tamoxifen or as an acute-phase reactant during acute hepatocellular
injury. In all of these circumstances, serum free T4 will be normal in those with
normal underlying thyroid reserve, but in those who require thyroid hormone
replacement, dose adjustments may be required to compensate for altered-binding
protein states.
A small decline in serum T3 concentration with age has been found, but
the decline is more likely attributable to mild nonthyroidal illness than age itself
(22,23). While there is little change in mean values for serum T3 as a function of
advancing age, fewer persons have values in the “upper range of normal” as derived
from data on individuals across the life span. In a study of healthy centenarians,
serum FT3 was slightly lower than that of healthy elderly aged 65 to 80 years,
suggesting mild impairment of peripheral 5’-deiodination in advanced age (18,19).
Other data derived from both animal and human models have reported
a small but significant decrease in cellular responsiveness to thyroid hormone
406 Cappola et al.

action (33,34). What clinical significance, if any, this latter phenomenon may have
remains controversial, but could explain the relative lack of adrenergic symptoms
in older patients with thyrotoxicosis (see later).

Autoimmunity
Thyroid antibodies in the serum increase in prevalence progressively with increas-
ing age, reaching a peak prevalence of 20% to 25% in women above the age of
50 years and 5% to 10% in similarly aged men (35–37). These numbers are similar
to those recently reported using data from NHANES (20). In an Italian study with
a highly selected population of healthy elderly ranging in age from 65 to 110 years,
the prevalence of antithyroid antibodies was low and did not differ from the preva-
lence seen in healthy young persons (18). These findings suggest that the high
prevalence of antithyroid antibodies in certain aging populations is a reflection of
disease and not a consequence of normal aging. Thus, the finding of high serum
antithyroglobulin and/or antithyroid peroxidase antibody titers and an elevation of
basal serum TSH concentration is consistent with autoimmune thyroiditis (38).

NONTHYROIDAL ILLNESS
The serum concentration of total T3 is reduced in many nonthyroidal illnesses,
giving rise to the “low T3 syndrome.” This consequence of systemic illness
is the earliest and most common of the alterations in thyroid hormone levels
(16,22,25,39–41). In response to many acute illnesses, there is decreased periph-
eral 5 -monodeiodination of T4 to T3, with consequent reduction in serum T3
concentration and an increase in serum rT3. Lowering of thyroid-binding prealbu-
min and TBG in the presence of acute and chronic illness further contributes to the
marked decrease in serum T3 that is characteristic of the sick elderly. Although
there are limited animal data to suggest that age may alter responses to stressors,
such as cold exposure and starvation, human studies are lacking.
The finding of a low T4 concentration in the presence of nonthyroidal ill-
ness has been termed the “euthyroid sick syndrome” (42). In mild-to-moderate
forms, measurement of free T4 will be normal even though total T4 concentration
is reduced (43). A putative inhibitor of T4 binding to TBG has been detected in
20% to 74% of hospitalized patients with nonthyroidal illness; this may contribute
to the measurement of a low T4 in these patients (44). However, severe illness
can result in a marked reduction of both total and free T4 concentrations. When
this circumstance occurs, the prognosis of the patient is poor, with a mortality
rate of approximately 80% having been reported (45). Since the serum TSH may
also be low in critically ill patients, the ability to differentiate these patients from
those with secondary hypothyroidism is often difficult, although neither T4 nor
T3 therapy improves prognosis in nonthyroidal illness (46–48). Cytokines such as
interleukin-1, interleukin-6, and tumor necrosis factor could be involved as inter-
mediaries in the development of low-T4 and low-T3 states. Many patients with
Thyroid Disease in the Elderly 407

nonthyroidal illnesses and low T4 and/or T3 also have elevated serum concentra-
tions of cytokines (49,50). Experimental increase of tumor necrosis factor-alpha
has been observed to induce low serum concentrations of T4, T3, and TSH (51).
Occasionally mild to moderately elevated concentrations of TSH are found in the
recovery phase of the illness, when TSH values can rise to as high as 20 mIU/L.
This does not indicate true hypothyroidism, since all measures of thyroid function
return to normal in a few weeks (52).
Biologically inactive reverse T3 (rT3) is generated by 5’-monodeiodination
of the inner ring of T4 (26,53). A variety of factors—including starvation, febrile
illness, elevation of glucocorticoid concentration, and drugs such as high-dose
propranolol, amiodarone, and iodinated contrast materials—can result in impaired
T4 conversion to T3, with a decline in serum T3 and a parallel increase in rT3
concentration due to decreased rT3 clearance (26). Serum rT3 may be affected
by normal aging, although an increase may also be a consequence of illness or
drug-induced alteration in rT3 degradation (23,24,54). However, the measurement
of rT3 is not clinically useful.

HYPOTHYROIDISM IN THE ELDERLY


Epidemiology
Several population-based studies report a prevalence of overt hypothyroidism in
the elderly of approximately 1% to 2%, as determined by an elevated TSH and
low free T4 level at a single point in time (55–57).

Pathophysiology
The most common cause of hypothyroidism in the elderly is autoimmune thyroidi-
tis (58). Another major cause of hypothyroidism is the prior treatment of hyper-
thyroidism with radioiodine or subtotal thyroidectomy (58,59). Hypothyroidism
may also be the natural sequel to previous Graves’ disease (60). Medications may
precipitate hypothyroidism, particularly in individuals with autoimmune thyroidi-
tis; these include iodine-containing radiographic contrast agents, amiodarone, and
iodine-containing cough medicines. Long-term lithium therapy can lead to inter-
ference with thyroid hormone synthesis and to inhibition of thyroid hormone
release, with up to a 20% prevalence of subclinical hypothyroidism and a similar
prevalence of up to 20% of overt hypothyroidism in patients taking this medica-
tion (32). Although infrequent, hypothyroidism in the elderly may also result from
pituitary or hypothalamic disease.

Diagnosis
The diagnosis of hypothyroidism in the elderly is often missed, because the pre-
senting complaints are often confused with other age-prevalent disorders. This
408 Cappola et al.

Table 2 Clinical Features of Elderly Patients (Mean Age 84.5 Years) with Overt
Hypothyroidism (n = 100) and with Subclinical Hypothyroidism (n = 20)

Overt Subclinical
Symptom/sign hypothyroidism hypothyroidism

Presenting Feature
Falls/deteriorating mobility/collapse 17 3
Hypothermia 2
General malaise 2
Confusion 4
Angina 1
Congestive cardiac failure 6
Cerebrovascular accident 4
Other conditions 64 17
Total 100 20
Signs
Hypothyroid facies 4
Hoarse voice 1
Slow relaxing reflexes 2 1a
Myopathy 2
Cerebellar signs 2
Effusions 9 4
Total 15 5
a Also had myopathy.

Source: From Ref. 62.

problem is further compounded by the often-insidious onset of illness. Doucet


and colleagues reported on the presenting symptoms and signs of hypothyroidism
in young (mean age 40.8 ± 9 years) and elderly subjects (mean age 79.3 ±
6.7 years). Statistical differences in presenting findings were noted for weight
gain (59% vs. 24%), cold intolerance (65% vs. 35%), paresthesias (61% vs. 18%),
and muscle cramps (55% vs. 20%). Complaints of fatigue and weakness were
present in more than 50% of the elderly patients (61).
Many persons who are later discovered to be hypothyroid will be unable to
identify when the symptoms actually began. Even patients known to have hypothy-
roidism may not have classic physical findings, and it often escapes recognition
until late in its course, after cognitive changes and functional decline interfere with
independent living (Table 2) (62–64). Neurological findings may include demen-
tia, ataxia, and carpal tunnel syndrome. On physical examination, delay in the
relaxation of deep tendon reflexes may be more difficult to elicit with advancing
age, due to other age-prevalent conditions affecting neurological function. Classic
signs and symptoms of dry skin, cold intolerance, paresthesias, constipation, and
hypothermia, among others, are worth looking for despite the patient’s age, but
Thyroid Disease in the Elderly 409

some of these findings can be due to aging itself. For example, with normal aging,
there may be changes in the skin and hair. The hair becomes coarser and there is a
tendency to lose the hair in the lateral aspects of the eyebrows. The skin becomes
dry, and scaling may be noted. Hypercholesterolemia may be more common in
both circumstances as well. For these reasons, the physician should maintain a
high index of suspicion of hypothyroidism when evaluating older persons, espe-
cially women and those with prior histories or family histories of some form of
thyroid disease.
The diagnosis of primary hypothyroidism in the elderly is best made by
measuring the serum TSH concentration (29). Changes in protein binding may
reduce the level of total T4; T3 may be reduced in persons with significant medical
illness or who are malnourished. For these reasons, an increase in serum TSH
remains the best way to detect primary hypothyroidism regardless of age. However,
as noted earlier, during the recovery phase after acute nonthyroidal illnesses, an
elevated serum TSH level may not represent true clinical hypothyroidism; in this
case the serum TSH returns to the normal range within four to six weeks. While
uncommon in the elderly, hypothyroidism can be secondary to pituitary failure or
hypothalamic disease. In this case, both serum TSH and T4 levels are low. While
antithyroid antibodies are detectable in the circulation of many persons with
hypothyroidism, these are not indicated in the workup of every patient suspected
of having hypothyroidism.

Treatment
L-Thyroxine is the preferred medication to treat hypothyroidism. While other
preparations are available, elderly persons tolerate the effects of T3 less well due
to its “burst effect” on the myocardium, causing an acute increase in oxygen
demand as well as the need for more frequent dosing. In general, brand names
are suggested to minimize the potential variability that may occur when generic
preparations are refilled. It is also easier for the elderly person to identify a medi-
cation with a consistent color and shape. Due to decreases in lean body mass and
T4 clearance, elderly patients require a smaller amount of L-thyroxine, on average
110 ± 8 ␮g/day, as compared with younger subjects, who on average require a
weight-related dose of approximately 1.6 ␮g/kg body weight per day (65–67). Due
to the increase in half-life for thyroxine that accompanies the aging process, it will
take longer to reach a steady state; therefore, a longer time between dose increases
is necessary in order to reduce unwanted side effects. The commonly used adage of
“start low and go slow” is usually followed in this age group, although this is based
on widespread acceptance rather than controlled clinical trials in older people. One
randomized clinical trial of initiation of a full replacement dose of levothyroxine
compared to gradual dose titration included a small number of individuals over the
age of 65 years without preexisting cardiovascular disease (68). No adverse seque-
lae were noted in the group of older individuals initiated on a full replacement dose,
but the generalizability of these findings to a larger population of older individuals
410 Cappola et al.

has been questioned (69). Because many elderly patients with hypothyroidism
may have an underlying cardiovascular problem, the conservative approach of ini-
tial therapy with 25 to 50 ␮g/day, with gradually increasing increments of 25 ␮g
every four to six weeks, remains the most consistently recommended approach.
Individuals with significant cardiac disease may require dose changes as low as
12.5 ␮g and could even be started at that low level. Once a dose of 75 ␮g/day is
achieved without side effects, increments of 12.5 ␮g are advised. The final dose
required is the amount of L-thyroxine that reduces the serum TSH into the range
of normal and does not have associated side effects. Caution is advised to avoid
iatrogenic subclinical hyperthyroidism with L-thyroxine. Surveys of the general
population have shown that only 60% of those taking thyroid hormone have TSH
levels in the euthyroid range, with 20% over-replaced (low TSH) and another
20% under-replaced (high TSH) (70). These data may underestimate the scope of
the problem in older individuals taking thyroid hormone, with euthyroid levels in
only 40% of individuals aged 60 years and over in the Framingham Heart Study,
and low TSH levels in an alarming 48% (71). The most likely explanation for the
high degree of over-replacement is a lack of physician monitoring or awareness
of adverse consequences of over-replacement. The limited data about the clin-
ical consequences of mild over-replacement with thyroid hormone (exogenous
subclinical hyperthyroidism) suggest that these consequences parallel those of
endogenous subclinical hyperthyroidism, with increased risk of atrial fibrillation
(72,73) and lower bone mineral density (74,75).

Myxedema Coma
Myxedema coma is a rare but serious consequence of untreated or inadequately
treated hypothyroidism. It occurs almost exclusively in the elderly (76). While
the name implies the presence of coma, the term also encompasses patients with
altered cognition, lethargy, seizures, psychotic symptoms, and/or confusion and
disorientation thought secondary to hypothyroidism. Most affected persons have
experienced a precipitating event such as severe infection, cold exposure, alco-
holism, or use of psychoactive medications. Exposure to sedatives and/or narcotics
has also been associated with this entity. While some patients will have a prior
history of being treated for hypothyroidism, many will be diagnosed for the first
time.

Diagnosis
Recognition of myxedema coma may be difficult, since its clinical presentation can
be attributed to a host of other problems. For those who can provide information,
a history of increased fatigue and somnolence is common. A past history of treat-
ment for thyroid illness—including medication, surgery, or radioactive iodine—
should raise concern. Use of narcotics, sedatives, or antipsychotic medication
may precipitate myxedema coma. Infections such as pneumonia and urosepsis
Thyroid Disease in the Elderly 411

are commonly found at the time of diagnosis; cold exposure, as often occurs in
poorly heated residential environments, has also been associated with myxedema
coma.
Physical examination may demonstrate many of the classic signs and symp-
toms of a hypothyroid state, such as dry scaly skin, edema, and bradycardia.
Profound hypothermia is often present and may be underestimated, because many
commonly used thermometers do not read temperatures less than 35◦ C. Patients
with myxedema coma may also demonstrate hypoventilation and hypotension.
Severe frontal and occipital headaches, psychotic behavior, ataxia, nystagmus,
muscle spasms, and sinus bradycardia may be signs of impending coma and must
not be dismissed. Other findings of severe hypothyroidism include pericardial
effusion, ileus, megacolon, and easy bruising.
Laboratory data may not be available early in the diagnostic process,
although the classic findings are a markedly reduced total and free serum T4 and
elevated serum TSH levels. Hyponatremia and hypoglycemia are not uncommon.
Hyponatremia is thought to result from inappropriately increased levels of antid-
iuretic hormone (ADH) as well as decreased water excretion due to a decrease
in glomerular filtration rate. Hypoglycemia is thought to result from impaired
gluconeogenesis and, in some patients, from cortisol deficiency. Since hypothy-
roidism in the elderly is often due to autoimmune thyroiditis, other autoimmune
deficiency states must be looked for, including diabetes and adrenal insufficiency.
Creatine phosphokinase is often elevated and may suggest the presence of a
myocardial infarction. Fractionation, however, usually demonstrates a muscle eti-
ology for the elevated creatine phosphokinase. Myocardial infarction can occur
in the presence of myxedema coma and may complicate initiation of thyroid hor-
mone therapy. Rarely, associated myoglobinuria and rhabdomyolysis have been
reported, although it is unclear if these findings result from circulatory collapse or
are directly related to the primary hypothyroid state.
Arterial blood gases usually show a decrease in PO2 and an increase in
PCO2 , indicating acute or impending respiratory failure. Anemia is also a common
finding and may range from a normochromic, normocytic variety to a macrocytic
form suggestive of B12 and/or folate deficiency. On chest X-ray, cardiomegaly
is often reported, which may be due to either a pericardial effusion or dilated
myocardium. Evoked potentials may have an abnormal amplitude or latency and
the electroencephalogram may demonstrate the presence of triphasic waves that
disappear after initiation of thyroid hormone replacement therapy.

Treatment
In patients with severe illness and coma, the decision to initiate thyroid replacement
therapy must often be based on clinical evidence prior to obtaining confirming
laboratory data (76–78). No controlled, blinded, or randomized studies exist to
clearly define the optimal therapeutic regimen at any age. The following principles
and guidelines should be followed:
412 Cappola et al.

1. Myxedema coma has a very high mortality rate if left inadequately treated.
2. There is the need to balance the uncertainty of diagnosis prior to receiving
laboratory results with empiric treatment and its complications, especially if
the individual is later found not to be hypothyroid.
3. Supportive therapy must be provided promptly and includes ventilatory sup-
port for respiratory failure, antibiotics for infection as indicated, and manage-
ment of hypothermia by external rewarming. Hypotension can be treated with
fluid replacement or dopamine. Hyponatremia must be treated. Occasionally,
severe hypoglycemia and anemia will require prompt therapy. Careful atten-
tion should be given to avoid aspiration of secretions, fecal impaction, and
urinary retention.
4. Prompt initiation of therapy with L-thyroxine is necessary to treat the hypothy-
roid state. The initial dose is 300 to 500 ␮g given intravenously once. Follow-
ing that, the dose should be reduced to 50 to 100 ␮g intravenously (76). T3
or combinations of T4 and T3 are not recommended in the elderly, since the
acute metabolic impact of T3 can precipitate cardiac arrhythmia or myocardial
infarction. The main principle is that a severely thyroid hormone-deficient
individual may require a high initial dose of thyroxine in order to occupy
hormone-binding sites that have been left free as a result of significant and
prolonged hormone deficiency. In addition, problems such as infection may
increase the turnover of thyroxine and thus further increase the need for a
higher initial replacement dose. The risk, however, is that this high dose can
increase myocardial oxygen consumption and increase the potential for a
myocardial infarction. Therefore, clinical judgment must prevail, especially
in the setting of other comorbidities prevalent in the elderly.
5. Whether or not intravenous glucocorticoids should be administered to all
persons with myxedema coma remains controversial. Adrenal insufficiency
can coexist with myxedema coma, so that suspicion of cortisol deficiency as
based on history, physical examination, or electrolyte abnormalities deserves
prompt treatment. In life-threatening situations, blood for measurement of
plasma cortisol should be drawn and intravenous stress doses of corticosteroids
should be administered and continued until there is laboratory confirmation of
status of adrenal function. At that time, the decision can be made either to treat
for concomitant adrenal insufficiency or to taper and stop the corticosteroid.

In summary, this condition is largely a problem of elderly hypothyroid


persons. It requires aggressive supportive therapy and hormone replacement while
contributing factors are evaluated and appropriately treated. Close monitoring is
required when treatment is initiated to avoid toxicity from large doses of thyroid
hormone.

Subclinical Hypothyroidism
Overt hypothyroidism usually causes at least one symptom, can progress to a life-
threatening condition if untreated (myxedema coma), and has been associated with
Thyroid Disease in the Elderly 413

adverse clinical sequelae. Therefore, management of overt thyroid dysfunction is


not controversial, and treatment is universally recommended. Clinical controversy
surrounds the management of patients with subclinical hypothyroidism, and at
what levels of TSH do the risk/benefit ratios favor therapy.
Subclinical hypothyroidism is characterized by relatively few clinical and
biochemical abnormalities (Table 2) (79–81). In fact, in one study of men and
women aged 60 years and over, there was no difference in the frequency in
any of the symptoms of hypothyroidism between subclinically hypothyroid and
euthyroid individuals (82). By definition, the circulating level of TSH is increased
in subclinical hypothyroidism, with serum T4 and T3 within the range of normal.
When measured in population-based studies at a single point in time, subclinical
hypothyroidism is highly prevalent in women and with increasing age, with a
prevalence of up to 20% of women over the age of 60 years (56,57,70,83). Rates
of persistence vary depending on the duration of follow-up and the degree of
initial TSH elevation, with rates ranging from 42% to 80% (56,84,85). In all of
these studies, an increase in progression to overt hypothyroidism was found in
those with positive antithyroid antibodies. In one study with a mean follow-up of
31.7 months, normalization of serum TSH was common in those with minimally
elevated TSH levels, occurring in 52% of those with TSH levels ⬍10 mU/L,
whereas 66% of those with TSH levels ⬎10 mU/L progressed to overt thyroid
dysfunction (85).
The majority of data supporting adverse sequelae from untreated subclinical
hypothyroidism are derived from cohort studies with baseline thyroid function
testing, with additional data from small case-control studies and clinical trials.
The major findings are outlined later.

Cardiovascular Outcomes
Mechanistic data support the potential for thyroid hormone deficiency to acceler-
ate atherosclerosis, via decreased cholesterol metabolism and increased systemic
vascular resistance, and congestive heart failure, via decreased ventricular filling
and cardiac contractility (86).
Analyses of incident cardiovascular disease have been performed in eight
major cohort studies (57,87–93). Three of these studies showed an increase in car-
diovascular risk in subclinical hypothyroidism (87–89), four showed no difference
(57,90–92), and one showed a decrease in cardiovascular death with increasing
TSH (93). Interestingly, when these cohorts were examined by age distribution,
a pattern emerges of increased risk during middle age and no appreciable risk
or possibly even benefit of subclinical hypothyroidism in old age. One published
study has shown an increase in incident congestive heart failure in those with a
TSH of 7 mU/L or greater (HR 2.33; 95% CI: 1.10–4.96) (92).

Neurocognitive Outcomes
Hypothyroidism has long been considered to be one of the reversible causes of
dementia (94) and of depression (95). In two large studies of older people, there
414 Cappola et al.

0.5
Abnormally low thyrotropin
Normal thyrotropin
0.4
Abnormally high thyrotropin

Cumulative mortality
Normal free thyroxine
Low free thyroxine 0.3

0.2

0.1

0 Cox regression P = .03 for Trend


85 86 87 88 89
Age (y)
Abnormally low thyrotropin 19 18 15 13 11
Normal thyrotropin 472 441 385 335 287
Abnormally high thyrotropin
Normal free thyroxine 30 28 26 25 23
Low free thyroxine 37 36 35 32 28

Figure 2 Cumulative mortality of Leiden 85-Plus study participants by thyroid status.


Thyrotropin levels below 0.3 mIU/L were considered to be abnormally low; levels above
4.8 mIU/L were considered to be abnormally high. Free thyroxine levels below 1.01 ng/dL
were considered to be abnormally low; levels between 1.01 and 1.79 ng/dL were considered
to be normal. Source: From Ref. 93.

were no differences between those with subclinical hypothyroidism and euthy-


roid individuals in cognitive scores or mood (96,97). This is in agreement with the
Leiden 85+ Study, in which there was no association between subclinical hypothy-
roidism and cognitive impairment, depression, or disability (93). Interestingly, in
this study of men and women aged 85+ years, those with higher TSH levels had
lower mortality (Fig. 2), which has led some in the field to postulate that subclinical
hypothyroidism represents an adaptive mechanism in the oldest-old (98).
These studies demonstrate that subclinical hypothyroidism is a prevalent
condition, with high rates of normalization in mild subclinical hypothyroidism and
progression to overt hypothyroidism in more severe subclinical hypothyroidism.
Furthermore, the potential adverse sequelae from untreated subclinical hypothy-
roidism are not consistent across studies, with reports of increased cardiovascular
disease, congestive heart failure, and dementia in some, but not all studies. Limited
data also suggest differential risk by degree of subclinical hypothyroidism and
age strata, supporting the need for risk group refinement. The central, currently
unresolved question is whether older individuals with subclinical hypothyroidism
benefit from replacement therapy with thyroid hormone. There are no substantive
data as yet to indicate that early treatment of subclinical hypothyroidism with
thyroid hormone replacement is effective in reducing the risk of subsequent
Thyroid Disease in the Elderly 415

development of cardiovascular or improvement in neurocognitive outcomes.


Some physicians advocate replacement therapy for all persons with subclinical
hypothyroidism, even those with minimally elevated levels of TSH. However, it
is more widely viewed that treatment should be given to those individuals with
TSH levels that are persistently ≥10 mU/L (80). For individuals with serum TSH
levels between 4.5 and 10 mU/L, follow-up without treatment is a reasonable
course, although many practitioners advocate a therapeutic trial. When the TSH
level is in this range, measurement of antithyroid peroxidase antibodies may
influence the decision about treatment, given the higher rate of progression to
overt disease. The goal of treatment, when initiated, is the same as for overt
hypothyroidism, to normalize serum TSH values. However, the target serum TSH
may be higher in the very old (17,98). This is usually accomplished with lower
doses of L-thyroxine than for those who present with overt hypothyroidism.

HYPERTHYROIDISM IN THE ELDERLY


Overproduction of thyroid hormone leads to the clinical condition of hyperthy-
roidism or thyrotoxicosis. This disorder is accompanied by a broad array of symp-
toms and signs that can differ markedly between young and old patients.

Epidemiology
In the past, hyperthyroidism was regarded as a disorder with preferential expression
in young to middle-aged individuals, especially women. It is now clear that this
disorder is also common in the elderly population. The proportion of patients
with hyperthyroidism who are older than 60 years is estimated to be 15% to 20%
(99,100). Several studies of prevalence indicate the presence of hyperthyroidism
in 1% to 2% of community-residing individuals (70). In a study of the population
of Whickham, England, hyperthyroidism was identified in 19 per 1000 women,
a prevalence rate 10 times greater than that in men. The mean age at diagnosis
was 48 years and the estimate of new cases was 2 to 3 per 1000 women per year
(101). Many other studies confirm that hyperthyroidism is far more common in
women than in men, with estimates of female preponderance ranging from 4:1
to as high as 10:1 (99,101–103). Population-based data from Tayside, Scotland,
obtained using record-linkage technology show both an increase in the incidence
of hyperthyroidism and a decrease in the gender difference with increasing age
(103). In those aged 80 years or older, the annual incidence rate was 1.05 per 1000
women and 0.45 per 1000 men.

Pathophysiology
In young persons, Graves’ disease remains the most common cause of hyperthy-
roidism. With increasing age, there is a change in etiology, so that more cases
are due to multinodular toxic goiter and fewer to Graves’ disease (99,104,105).
Multinodular goiters are common in the elderly and may not be clinically apparent
416 Cappola et al.

(4). Many clinical observations support the concept that long-standing euthyroid
multinodular goiters may evolve to become toxic multinodular thyroid goiters (99).
Toxic adenoma is another less common cause of hyperthyroidism in the
elderly, usually identifiable on thyroid scanning by the demonstration of a solitary
hyperfunctioning nodule with suppression of activity in the remainder of the thy-
roid gland. Hyperthyroidism can also rarely occur in a previously euthyroid elderly
person following ingestion of iodide- or iodine-containing substances. Most com-
monly, this occurs following exposure to iodinated radio-contrast agents and to
amiodarone. Although up to 40% of patients taking amiodarone will have serum
T4 levels above the normal range due to the drug’s effect on T4 metabolism, far
fewer (5%) will develop true thyrotoxicosis (106). This form of hyperthyroidism
can be of rapid onset and severe in magnitude. Because of amiodarone’s fat-
solubility and long half-life, drug-induced thyrotoxicosis can be prolonged and
difficult to treat (107,108).
The possibility of hyperthyroidism must always be considered in the elderly
person who is receiving thyroid hormone, especially if the dose is ⬎0.15 mg of
L-thyroxine daily. Patients who have received such doses for many years without
evidence for hyperthyroidism may insidiously develop hyperthyroid features as
their age passes 60 years owing to the age-associated reduction in rate of thyroid
hormone metabolism (109).
Rare causes of hyperthyroidism in the elderly include TSH-producing pitu-
itary tumors (110). These can be recognized by the finding of nonsuppressed
levels of serum TSH in the presence of increased amounts of circulating thyroid
hormone. An additional uncommon cause of hyperthyroidism is overproduction
of thyroid hormone by widespread metastatic follicular carcinoma.
Transient hyperthyroidism may occur in patients with silent or subacute
thyroiditis as a result of increased discharge of thyroid hormone into the circulation
during the inflammatory phase of the illness. In a similar fashion, radiation injury
to the thyroid can be accompanied by a transient increase in circulating thyroid
hormone levels with associated symptoms.

T3 and T4 Toxicosis
In a small proportion of cases of hyperthyroidism, the expected increase in serum
T3 is noted, but serum T4 is either within the normal range or at the upper end
of normal. This circumstance has been designated as T3 toxicosis and can occur
with any type of hyperthyroidism; however, it commonly occurs in older patients
with toxic multinodular goiter or solitary toxic adenoma (111). The diagnosis
will not be missed if serum T3 is measured in patients with clinically suspected
hyperthyroidism or who have suppressed levels of serum TSH. T4 toxicosis,
or an isolated increase in serum T4 without an elevation in serum T3, is more
commonly noted in the sick elderly patient with hyperthyroidism owing to the
higher prevalence of medical conditions and nutritional inadequacies that interfere
with T4-to-T3 conversion.
Thyroid Disease in the Elderly 417

Table 3 Frequency of Symptoms and Signs of Hyperthyroidism in Elderly Versus


Young

Kawabe et al.
(112) Davis and Davis (99)
Young (n = 48) Elderly (n = 45) Elderly (n = 85)
Symptom/sign (%) (%) (%)

Palpitation 100 60 63
Goiter 98 58 64
Tremor 96 71 55
Excessive perspiration 92 66 38
Weight loss 73 85 69
Eye signs 71 28 57
Arrhythmias (atrial 4.6 16.4 62
fibrillation and ventricular
premature contraction)

Source: From Ref.113.

Diagnosis
As with other disorders occurring in the elderly person, the clinical presentation
of hyperthyroidism often differs from the classical description of the disease in
younger individuals (99,100,104,112–114) (Table 3). The presenting feature may
be a progressive functional decline including muscle weakness, fatigue, changes in
mental status, loss of appetite, weight loss, cardiac arrhythmia, and/or congestive
heart failure. A symptom complex peculiar to the geriatric hyperthyroid patient is
“apathetic hyperthyroidism,” in which the patient lacks the hyperactivity, irritabil-
ity, and restlessness common to the young patient with thyrotoxicosis and presents
instead with weakness, lethargy, listlessness, depression, and the appearance of a
chronic, wasting illness. Often, the initial impression is depression, malignancy,
or cardiovascular disease (115,116).
Clinically detectable thyroid enlargement, present in almost all younger
patients, is absent in as many as 37% of elderly patients (99). Infiltrative oph-
thalmopathy with severe proptosis and exophthalmos occurs infrequently in the
elderly. Thus, none of the elements of the classic triad of Graves’ disease (clinical
hyperthyroidism, diffuse goiter, and infiltrative ophthalmopathy) may be recog-
nizable in the elderly patient, in whom the diagnosis may be suspected only on
the basis of laboratory studies (99,112,113).
Symptoms less commonly present in the elderly include nervousness,
increased sweating, tremor, increased appetite, and increased frequency of bowel
movements. Symptoms more common in the elderly include marked weight loss,
present in ⬎80% of patients, poor appetite, worsening angina, edema, agitation,
and confusion. There may be correction of previously existing constipation. Sim-
ilarly, physical findings differ in elderly patients. In addition to the absence of a
418 Cappola et al.

palpable goiter and eye signs of exophthalmos, the pulse rate tends to be slower,
and hyperreflexia may not be present. Lid lag and lid retraction are frequently seen
(99,112–116).
The spectrum of symptoms and findings due to thyroid hormone excess
is broad and can involve almost all body systems. Cardiac manifestations are
particularly important in the elderly person who may have coexisting heart disease.
An increased heart rate with a related increase in myocardial oxygen demand,
stroke volume, cardiac output, and shortened left ventricular ejection time underlie
the clinical consequences of palpitations, increased risk of atrial fibrillation—often
with slow ventricular response, exacerbation of angina in patients with preexisting
coronary artery disease, and precipitation of congestive heart failure that responds
less readily to digoxin treatment (117).
Gastrointestinal consequences of hyperthyroidism in the elderly include
weight loss, poor appetite, and occasionally abdominal pain, nausea, and vomiting
(99,100). Diarrhea and increased frequency of bowel movements resulting from
thyroid hormone action on intestinal motility can occur, but these symptoms are
often absent in the elderly, in whom constipation is as likely to be present. Hepatic
actions of thyroid hormone can lead to alterations in liver enzymes, including ele-
vation of alkaline phosphatase and gamma-glutamyl transpeptidase levels, which
return to normal following restoration of thyroid function to euthyroid values.
Weakness, especially of the proximal muscles, is a major feature of hyper-
thyroidism in the elderly and is often accompanied by muscle wasting (99,116).
As a consequence, disorders of gait, postural instability, and falls can take place.
Tremor occurs in ⬎70% of elderly thyrotoxic patients, but this sign must be distin-
guished from other causes of tremor that are common in the elderly and are usually
more coarse in nature or primarily resting tremors (112,113). A rapid relaxation
phase of the deep tendon reflexes is common in young patients but is often diffi-
cult to assess in the older patient. Central nervous system manifestations may be
prominent in the elderly patient and include confusion, depression, forgetfulness,
agitation/anxiety, and a shortened concentration span (95,104). These cognitive
impairments may point to a diagnosis of dementia with failure to consider the
presence of hyperthyroidism.
Other clinical manifestations of hyperthyroidism in the elderly may include
glucose intolerance; occasionally, latent diabetes mellitus may be unmasked. Mild
elevations of serum calcium can occur, and hyperthyroidism should be ruled out
as a secondary cause of osteoporosis in persons with decreased bone mass.
Because of the altered clinical presentation of hyperthyroidism in the elderly,
suspicion must always be high and the laboratory should be used for any patient
with possible symptoms. Many have advocated screening tests for thyroid status
in all geriatric patients undergoing initial clinical evaluation (118,119).
Measurement of serum TSH by ultrasensitive methods and serum free T4
or free T4 index is the preferable test for diagnosing thyroid dysfunction. The
findings of a normal or low serum free T4 with suppressed serum TSH raises the
possibility of T3 toxicosis and calls for measurement of serum T3. Demonstration
Thyroid Disease in the Elderly 419

of anti-TSH receptor antibodies can be helpful in making a diagnosis of Graves’


disease, but this is not a routine test (120).
Thyroid scanning and measurement of the 24-hour 131 I uptake can be useful
in distinguishing Graves’ disease from toxic multinodular goiter. Scanning may
also demonstrate the presence of a small, diffusely active goiter that could not be
detected on physical examination. Very low 131 I uptake in a patient with elevated
circulating thyroid hormone levels suggests exogenous thyroid hormone ingestion
(factitious or iatrogenic hyperthyroidism), the hyperthyroid phase of painless or
subacute thyroiditis, or iodine-induced hyperthyroidism.

Treatment
The first step in the management of the elderly patient with hyperthyroidism is
to determine the underlying etiology and exclude the possibility of one of the
transient forms, which may require therapy directed toward the primary process
(hormone ingestion, iodine exposure, and subacute thyroiditis). While the vast
majority of patients with either Graves’ disease or multinodular toxic goiter can
be treated using antithyroid drugs, radioactive iodine, or surgery, radioactive iodine
ablation of the thyroid is the preferred treatment for both etiologies in the elderly
(121).
In the older patient with suspected hyperthyroidism who is still undergo-
ing investigation, a useful initial step in treatment is the administration of beta-
adrenergic blocking agents such as long-acting propranolol, metoprolol, nadolol,
or atenolol. In patients who have palpitations, tachycardia, angina, or agitation,
these symptoms of thyrotoxicosis can be quickly controlled with use of the beta-
blockers, although caution is advised in elderly persons with congestive heart
failure, chronic obstructive pulmonary disease, or diabetes being treated with
insulin.
Once a diagnosis of Graves’ disease or toxic nodular goiter is established,
treatment should be started with the antithyroid drug methimazole (122). The
antithyroid drugs impair biosynthesis of thyroid hormone; this will lead to deple-
tion of intrathyroidal hormone stores and ultimately to decreased hormone secre-
tion. A decline in serum T4 concentration is usually seen within two to four weeks
after initiation of antithyroid drug therapy, and the dose should be tapered once
thyroid hormone levels reach the normal range in order to avoid hypothyroidism.
In 1% to 5% of patients, the antithyroid drugs may cause fever, rash, and arthralgia.
Drug-induced agranulocytosis may be more common in the elderly and is most
likely to occur within the first 3 months of treatment, especially in those patients
who receive more than 30 mg/day of methimazole (123). Some but not all experts
recommend periodic monitoring of the white blood cell count and discontinuation
of the drugs if there is evidence of the development of neutropenia. Long-term
antithyroid drug administration can be an effective therapy in patients above the
age of 60 years with Graves’ disease who, in comparison to younger persons,
appear to respond more quickly and have a greater likelihood of long-lasting
420 Cappola et al.

remission (124). In contrast, this approach is rarely successful in elderly patients


with a toxic multinodular goiter, since it is not an autoimmune disease.
Ablation of the thyroid with 131 I is the recommended treatment in most
elderly persons. Because of the necessary contact precautions immediately fol-
lowing treatment with 131 I, there may be radiation safety issues for caretakers,
especially if patients have urinary incontinence. Once the patient has been ren-
dered euthyroid by antithyroid drugs, these agents can be stopped for three to
five days, following which 131 I is given orally. Therapy with beta-blockers can be
maintained and antithyroid agents can be restarted five days after radiotherapy and
continued for one to three months until the major effect of radioiodine is achieved.
The treatment goal in all elderly patients is to assure ablation of thyroid tissue
and avoid the possibility of recurrence of hyperthyroidism. Using this approach,
patients are monitored following treatment until their serum thyroid hormone
levels reach the hypothyroid range; they are then put on permanent replacement
therapy. Hypothyroidism may be evident as early as four weeks after treatment
but can occur at any time. In some circumstances, when clinical and laboratory
features of hyperthyroidism are mild and there are no significant cardiac problems,
it may be appropriate to treat the elderly hyperthyroid patient with 131 I without
antithyroid drug pretreatment. When this option is chosen, the patient is started on
beta-blocker therapy before administering 131 I and the beta-blocker is continued
until there is evidence that thyroid hormone levels have fallen into the normal
range.
Surgery is not recommended as a primary choice for the treatment of hyper-
thyroidism in the elderly. The frequent accompaniment of hyperthyroidism by
other coexisting illnesses puts the patient at increased operative risk. Surgery
may be of value for the patient with poor response to medical treatment, severe
ophthalmopathy, or tracheal compression secondary to a large goiter.

Atrial Fibrillation
Atrial fibrillation occurs in 10% to 20% of older hyperthyroid patients (103), and
raises several management issues regarding cardioversion and anticoagulation. In a
retrospective study of 163 thyrotoxic patients with atrial fibrillation, approximately
60% had spontaneous reversion to sinus rhythm after return to a euthyroid state.
Most of these reversions took place within three weeks of becoming euthyroid;
none occurred if the patient still had atrial fibrillation after 4 months of euthy-
roidism; and none occurred when atrial fibrillation had been present for more than
13 months before becoming euthyroid. Thus, the patient who still has atrial fibril-
lation beyond 16 weeks of return to euthyroidism is a candidate for cardioversion
(125). Current practice guidelines recommend anticoagulation for patients with
thyrotoxic atrial fibrillation (126). In the absence of contraindications, warfarin
should be given in a dose that will increase the International Normalization Ratio
(INR) to 2.0 to 3.0 and continued until the patient is euthyroid and there has been
restoration of normal sinus rhythm. While in the hyperthyroid state, the patient
Thyroid Disease in the Elderly 421

with atrial fibrillation is more sensitive to the anticoagulant effect of warfarin,


which causes a greater lowering of activity of coagulation factors II and VII and
greater increase in prothrombin ratio and partial thromboplastin time (127).

Thyroid Storm
Acute hyperthyroidism or “thyroid storm” can occur in the patient with either
known or undiagnosed hyperthyroidism who is subjected to acute stress, such as an
operative procedure, trauma, or infection or who is exposed to iodine-containing
drugs (128). It can also occur in the elderly hyperthyroid patient treated with
131
I who did not receive adequate antithyroid medication prior to therapy. The
features of severe hyperthyroidism can develop over several hours and include
fever, tachycardia, arrhythmia, dyspnea, vomiting, diarrhea, dehydration, severe
restlessness, and delirium. Acute heart failure can be precipitated. Elderly patients
with cardiac disease are at especially high risk for the development of acute
myocardial ischemia or congestive heart failure.
Thyroid storm is a life-threatening condition and must be treated vigor-
ously and promptly (128). Immediate treatment focuses on inhibition of thyroid
hormone production and the inhibition of conversion of T4 to T3 in peripheral
tissues. This is accomplished by administration of antithyroid drugs, especially
propylthiouracil, in an initial dose of 200 to 400 mg orally every six to eight hours.
The first dose of sodium iodide or oral Lugol’s solution should follow the first
antithyroid drug dose. High-dose beta-blockers such as propranolol and high-dose
corticosteroids are also given to blunt peripheral action of thyroid hormone and to
inhibit T4-to-T3 conversion. Supportive measures are provided and may include
sedation, correction of fluid and electrolyte imbalances, antipyretics (not aspirin),
and cooling blankets and antibiotics if infection is present.

Subclinical Hyperthyroidism
By definition, these individuals have normal circulating levels of free T4 and
T3 and levels of TSH below the lower limit of normal (usually 0.3–0.5 mU/L),
implying that the circulating level of thyroid hormone is more than is required for
hormonal balance. However, clinical symptoms may not be present. Although the
Wayne score, a clinical index of thyrotoxicosis, was noted to be higher than in a
group of similarly aged euthyroid persons, there were few or no classic features of
hyperthyroidism (129). The prevalence of subclinical hyperthyroidism on a single
serum TSH test is 1% to 6% in an older population (55–57,72). Persistence of this
testing abnormality ranges from 24% to 88%, depending on the degree of TSH
lowering (56,130). In a recent very large study of more than 400,000 persons,
3000 of whom had a low serum TSH level, about 50% of such individuals had
a normal serum TSH on follow-up (131). Progression to overt hyperthyroidism
is uncommon, estimated at approximately 1% per year (72), and may be more
likely in patients with toxic multinodular goiter compared to those with Graves’
disease (132).
422 Cappola et al.

The majority of data supporting adverse sequelae from untreated subclinical


hyperthyroidism are derived from cohort studies in which thyroid function testing
was performed at baseline, with additional data from small case-control studies
and clinical trials. Findings from the major studies are presented by organ system.

Cardiovascular Outcomes
One of the most important findings in the field was the report in 1994 by Sawin
et al., using data from the Framingham Heart Study, of an increased risk of atrial
fibrillation in those with subclinical hyperthyroidism compared to the euthyroid
state (71). They found that individuals with TSH values ≤0.1 mU/L, many of
whom were receiving thyroid hormone therapy, had an adjusted relative risk
of 3.8 (95% CI: 1.7–8.3) for developing atrial fibrillation and those with TSH
values between 0.1 and 0.4 mU/L had an adjusted relative risk of 1.6 (95%
CI: 1.0–2.5). The lack of a statistically significant difference led to lingering
questions about the risks of milder subclinical hyperthyroidism. Furthermore,
individuals with elevated thyroxine levels, indicating overt hyperthyroidism, were
included in their category of TSH values ≤0.1 mU/L, which could have led to an
overestimate of the effect of subclinical hyperthyroidism. Additional analyses in
the Cardiovascular Health Study in 2006 by Cappola et al. confirmed the atrial
fibrillation risk (57). After adjustment for age, sex, clinical cardiovascular disease
at baseline, subsequent thyroid hormone use, and other known risk factors for
atrial fibrillation, participants with subclinical hyperthyroidism had nearly twice
the risk of developing atrial fibrillation (HR 1.98; 95% CI, 1.29–3.03). When these
analyses were repeated, limiting to subjects with mild subclinical hyperthyroidism
(TSH 0.1–0.44 mU/L), the adjusted hazard ratio remained significantly elevated, at
1.85 (95% CI 1.1–3.0). Despite the increase in the frequency of atrial fibrillation
found in this study, overall mortality was not increased. This is at odds with
another study of older persons with subclinical hyperthyroidism, where higher
all cause and cardiovascular mortality rates were observed (91). Increased left
ventricular mass relative to euthyroid individuals has also been noted in small
studies, although the clinical significance of this finding is unclear (133).

Musculoskeletal Outcomes
Using data from the Study of Osteoporotic Fractures, Bauer et al. have shown that
postmenopausal women with serum TSH concentrations of 0.1 mU/L or less had
an increased risk of both hip (HR 3.6; 95% CI: 1.0–12.9) and vertebral fractures
(HR 4.5; 95% CI: 1.3–15.6) compared to women with normal TSH levels (134).
However, T4 levels were not measured, so that the proportion of women with
overt and subclinical hyperthyroidisms is not known, although the intermediate
risk from TSH levels of 0.1–0.5 mU/L (HR 1.9; 95% CI: 0.7–4.8 for hip fractures
and HR 2.8; 95% CI: 1.0–8.5 for vertebral fractures) suggests a valid dose–
response relationship. An additional limitation to this study is that 86% of those
Thyroid Disease in the Elderly 423

with low TSH levels were taking thyroid hormone, so that it was not possible
to distinguish between effects from endogenous and exogenous hyperthyroidism.
Other evidence indicates that treatment of subclinical hyperthyroidism may have
a beneficial effect on bone mineral density. In a nonrandomized prospective study,
postmenopausal women with subclinical hyperthyroidism and nodular goiter who
were treated with 131 I with subsequent restoration of serum TSH into the normal
range demonstrated increases in bone mineral density at the spine and hip after
two years of follow-up, while a similar group of women who were not treated
showed progressive decline in bone mineral density (135).

Neurocognitive Outcomes
An analysis from the Rotterdam Study has shown a more than threefold increased
risk of dementia over a two- to four-year period in those with subclinical hyper-
thyroidism (RR 3.5; 95% CI: 1.2–10.0), although this analysis was only adjusted
for age and sex (136). However, two other large studies have shown no association
of subclinical hyperthyroidism with cognitive impairment or depression (93,97).
These studies demonstrate that subclinical hyperthyroidism is a prevalent
condition, with the available data showing a high rate of persistence and low
rate of progression to overt hyperthyroidism. There are potential adverse sequelae
from untreated subclinical hyperthyroidism, with reports of increased incidence
of atrial fibrillation, fracture, and dementia. Largely based on the atrial fibrillation
and bone data, many practitioners treat older individuals with a suppressed TSH
(⬍0.1 mU/L) (137), and this is recommended in current practice guidelines (80).
The benefit of treatment of those with a TSH of 0.1 to 0.4 mU/L remains unde-
fined, although reasonable if there are any clearly associated symptoms such as
a worsening of cardiovascular function or cardiac arrhythmias, excessive muscle
wasting, anorexia, or depression.

SCREENING FOR THYROID DYSFUNCTION IN THE ELDERLY


The goal of screening is to apply a test to identify disease in an unsuspecting pop-
ulation without any clinical indication of disease, under the assumption that early
detection provides benefit. Because of the high prevalence of thyroid testing abnor-
malities in older people and the nonspecificity of individual symptoms, screening
has particular appeal for the elderly population. Screening for thyroid dysfunction
presents several unique challenges: how to incorporate nonspecific symptoms in
narrowing the tested population, the transience of some thyroid testing abnormali-
ties, and, most importantly, how to manage subclinical thyroid dysfunction. Several
guidelines have been published regarding screening for thyroid dysfunction. The
U.S. Preventive Services Task Force, The Institute of Medicine, and an expert
panel do not recommend generalized screening for thyroid disease (80,138,139).
The American College of Physicians at one time recommended screening women
older than 50 years for unsuspected but symptomatic thyroid disease, although
424 Cappola et al.

this guideline is currently inactive (140), and the American Thyroid Association
recommends screening adults every five years beginning at the age of 35 years
(141). Without data showing the threshold TSH levels for treatment, clinical trials
showing benefits that outweigh risks, and the benefits of early diagnosis of mildly
abnormal serum TSH levels, a universal screening strategy for thyroid dysfunction
seems premature.

NODULAR THYROID DISEASE AND NEOPLASIA


Epidemiology
The development of thyroid nodules is clearly an age-related process that occurs
more commonly in women than in men. Autopsy studies have demonstrated that
thyroid nodules are frequently found in the elderly even when clinical examination
of the neck has failed to reveal abnormality (142). These data show an increase
in frequency of nodules in women and men over the age of 30 years, with a
progressive increase to a frequency of 90% in women and 50% in men over the
age of 70 years. In the Whickham study, clinically detectable nodules were found
in 0.8% of men without a relationship to age (101). However, 5.3% of women had
palpable nodules and the frequency increased from about 4% in those under the age
of 50 years to 9.1% in those aged 75 years or more. By means of ultrasonographic
study, thyroid nodules have been found in approximately 50% of women beyond
the age of 50 years (143).

Pathophysiology
Patients with thyroid nodules should be questioned for a history of childhood
external radiation exposure of the head, neck, and upper thorax, since it is well
established that radiation to this area markedly increases the risk of developing
thyroid malignancy. It was common practice from the 1930s to the early 1960s to
treat facial acne, tonsillar enlargement, cervical adenitis, and thymic enlargement
with external radiation. It is estimated that several million people were irradiated,
many of whom are now older than 60 years. Although a history of irradiation in a
patient with nodular thyroid increases the likelihood of malignancy, it is important
to note that irradiation also increases the development of benign nodules of the
thyroid as well as parathyroid adenomas. Some 16% to 29% of individuals who as
children received low-dose head and neck irradiation will develop palpable thyroid
nodules, approximately one-third of which will be malignant. Nodules become
apparent after a latency of 10 to 20 years, and the incidence of malignant nodules
reaches a peak 20 to 30 years after exposure (144).
Multinodular thyroid glands occur more commonly in individuals who come
from areas of iodine deficiency. Often there is a history of goiter dating back
to childhood or young adult years. Very large multinodular goiters, particularly
those with a sizable substernal component, may compress the trachea and lead
to complaints of dyspnea or wheezing. Disturbances of swallowing may also
Thyroid Disease in the Elderly 425

occur. These compressive symptoms are most common in older women. A large
substernal goiter is often first recognized when the patient has had a chest X-ray
and is noted to be associated with compression or deviation of the trachea or a
superior mediastinal mass (145).

Diagnosis
Thyroid nodules are most commonly asymptomatic. They may be discovered
accidentally by the patient or found by the physician during the course of a physical
examination. Occasionally, a thyroid nodule will be associated with acute onset of
neck pain and tenderness. This circumstance may be the result of acute or subacute
thyroiditis but is more likely due to hemorrhage into a preexisting nodule. The
finding of a solitary palpable nodule raises the possibility of thyroid malignancy. In
many patients, further evaluation will reveal the presence of multiple nodules. The
likelihood of thyroid cancer per patient is independent of the number of thyroid
nodules present in the gland, and approximately 15% of solitary nodules ⬎10 mm
are malignant (146).
Many histological entities can present as a thyroid nodule. The vast major-
ity is benign, including follicular and colloid adenomas, Hashimoto’s thyroidi-
tis, and thyroid cysts. Malignant thyroid neoplasms include papillary, follicular,
medullary, and anaplastic carcinomas as well as thyroid lymphoma and metastases
to the thyroid. Nonthyroid lesions may also appear to be thyroid nodules; these
include lymph nodes, vascular structures, parathyroid adenomas and cysts, and
thyroglossal duct cysts.
The likelihood that a single nodule is malignant is increased if there is a
history of radiation exposure, if it occurs in a man over the age of 60 years,
has been observed to undergo an increase in size, is accompanied by hoarseness
of the voice suggestive of impingement on the recurrent laryngeal nerve, and is
stony-hard on palpation (147).
Data derived from the 53,856 patients with thyroid cancer in the National
Cancer Data Base indicate that approximately 25% of all thyroid cancers are first
diagnosed in individuals older than 60 years (Fig. 3) (148). Age is a factor in
predicting the histological type of malignancy. The overall histological distribu-
tion of thyroid cancer is 79% papillary, 13% follicular, 3% Hürthle cell, 3.5%
medullary, and 1.7% anaplastic. In patients over the age of 60 years, papillary car-
cinoma accounts for approximately 67% of thyroid cancers. Follicular carcinoma
has a peak frequency of diagnosis in both the fourth and sixth decades of life,
with a mean age of 44 years. It, along with related Hürthle cell carcinoma, makes
up approximately 23% of the thyroid malignancies in the over-60 population.
Medullary carcinoma has a peak incidence in the fifth and sixth decades and make
up about 5% of thyroid cancers in the elderly (Table 4).
Anaplastic carcinoma of the thyroid is almost exclusively a disease of older
persons and accounts for approximately 10% of thyroid cancers in this age group.
It is invariably fatal in a short period of time from its first diagnosis, especially
426 Cappola et al.

35

30

25

20

15

10

0
0-29 30-39 40-49 50-59 60-69 70-79 80+
Age (Years)
Papillary Follicular Hurthle Cell Medullary Anaplastic

Figure 3 Histological types of thyroid cancer by age. Source: From Ref. 148.

when it is ⬎5 cm in diameter (149). Clinically, it often arises in a background


of previous multinodular goiter or papillary thyroid cancer and is recognized by
its rapid growth, rock-like consistency, and local invasiveness—with recurrent
laryngeal nerve involvement and tracheal compression common consequences.
Lymphoma and metastatic cancers make up the remaining thyroid malig-
nancies of the older person. Lymphoma is characterized by a rapidly enlarging
painless neck mass, which may cause compressive symptoms. Initially, it may be
difficult to differentiate from anaplastic carcinoma on clinical appearance alone.
Hashimoto’s thyroiditis, identified by histological or antibody testing, is com-
monly present in patients with lymphoma (150).
The major objective of evaluating an elderly patient with a thyroid nod-
ule, especially an apparently single nodule, is to determine whether the nodule
is benign or malignant. A number of diagnostic modalities are available that will
then point the way to appropriate treatment (151). Blood tests of thyroid function
will usually give normal results with the exception of the patient with a hyper-
functioning adenoma or toxic multinodular goiter. In some patients with nodular

Table 4 Occurrence and Survivorship of Thyroid Malignancy in the Older Patient


% of patients with % of patients with
type type 10-yr survival

Cancer type Age ⬎ 40 Age ⬎ 60 Age ⬎ 60 (%)

Papillary/mixed 79 60–67 ⬍65


Follicular 13 20–25 ⬍57
Medullary 3 5 ⬍63
Anaplastic 2 6 0
Lymphoma 3 5 ≤100
Thyroid Disease in the Elderly 427

disease due to Hashimoto’s thyroiditis, serum TSH may be increased. Serum thy-
roglobulin is often elevated in patients with thyroid cancer but cannot reliably
differentiate malignancy from benign adenoma or thyroiditis. Its major usefulness
is in the early recognition of recurrence or metastasis in patients with papillary or
follicular carcinoma who had previously undergone total thyroidectomy. Elevation
of serum calcitonin concentration is highly supportive of a diagnosis of medullary
carcinoma but is not cost-effective as a routine test unless there is a family history
of multiple endocrine neoplasia.
Fine-needle aspiration (FNA) of the thyroid to obtain tissue for cytological
or histological examination is the most reliable and accurate method of separating
benign from malignant disease (151). FNA is indicated in any patient with a
solitary nodule and when there is suspicion of thyroid malignancy based on clinical
assessment, ultrasound findings, or thyroid scan. In skilled hands, the procedure is
safe, inexpensive, and capable of determining presence or absence of malignancy
with 95% accuracy. The reliability of FNA can be further increased if it is done in
conjunction with real-time sonographic guidance. The cytopathological findings
from FNA are assigned to four categories: positive for malignancy, suspicious
for malignancy, negative for malignancy, and nondiagnostic. In the case of a
nondiagnostic aspirate, a repeat FNA is recommended. If FNA reveals malignant
cells, surgery is recommended with little need for further study. The combination
of suspicious cytology by FNA and cold appearance on scanning indicates the need
for surgical excision. Demonstration of benign cytology in either a solid or cystic
nodule indicates that the patient can be managed subsequently by observation.
The patient with clinical or FNA features suggestive of lymphoma should have a
large-needle or surgical biopsy done to establish the diagnosis (152).
Isotopic scanning is no longer considered as an initial diagnostic test because
it is not cost-effective. The major role for isotope imaging is in the evaluation of
the patient with a thyroid nodule who has had a nondiagnostic result from FNA.
Malignant tissue is rarely able to take up iodine, so that identification of a nodule as
“hot” by 123 I or technetium scanning makes malignancy in the nodule less likely.
The finding of a nonfunctioning or “cold” nodule does not establish a diagnosis
of malignancy, since 95% of thyroid nodules are cold, with the frequency of
malignancy in cold nodules being 5%. Nodules identified as being “hot” and with
associated normal thyroid hormone production and no compressive symptoms
warrant only observation, with examination at intervals of 6 to 12 months. If
evidence of hyperthyroidism is subsequently found, appropriate treatment for this
condition should be initiated.
High-resolution ultrasonography can detect lesions as small as 2 mm and
can permit classification of a nodule as solid, cystic, or mixed solid-cystic. The
technique will often demonstrate multinodularity in a gland with a single palpable
nodule. The value of ultrasonography in establishing a diagnosis of malignancy
is limited, since there is considerable overlap in the ultrasound characteristics of
benign and malignant nodules. The procedure may be useful in detecting recurrent
or residual thyroid cancer and in screening individuals who have had a history of
428 Cappola et al.

irradiation exposure. It also may be of value in the patient suspected of having


lymphoma of the thyroid, since this malignancy often produces a characteristic
asymmetrical pseudocystic pattern (152).
Computed tomography and magnetic resonance imaging can provide
detailed information on thyroid anatomy. These procedures are expensive and
appear to add little to the initial clinical assessment of malignancy. However, they
may be useful in evaluating the extent of disease in patients who are found to
have anaplastic carcinoma or lymphoma of the thyroid. They can also provide
information about compression of structures in the neck as well as on the size and
substernal extent of larger nodules and goiters.

Treatment
In the past, thyroid hormone has been used to suppress TSH on the assumption
that benign lesions were more likely to be TSH dependent and therefore likely to
decrease in size. The procedure involves giving L-thyroxine in a dose sufficient to
suppress serum TSH and monitoring the size of the thyroid nodule for a period
of three to six months. Although a meta-analysis of randomized, controlled trials
supports a decrease in thyroid nodule volume with L-thyroxine (153), this ther-
apy is no longer standard practice. In addition, the administration of suppressive
doses of L-thyroxine to elderly patients carries substantial risk for precipitation or
aggravation of ischemic heart disease and for acceleration of bone loss.
The basic principles in management of thyroid cancer do not differ signif-
icantly between elderly and young patients, other than comorbidities that may
affect the ability of the older person to undergo surgery. As the first line therapy
for thyroid cancer, this approach should be encouraged, as the impact on survival
is significant, even in this age group (154). If a diagnosis of papillary or follicular
carcinoma has been confirmed prior to surgery, near total thyroidectomy should
be carried out because older individuals present with more aggressive disease and
a high frequency of multicentricity, and there is the need to remove functional
thyroid tissue in order to monitor the patient with total-body radioiodine scanning
and basal and stimulated serum thyroglobulin levels (155). If radioactive iodine
therapy is required after surgery, an additional factor in the elderly is the dosage
of 131 I employed. Rather than empiric dosing, dosimetry should be considered in
the elderly, due to increased likelihood of exceeding maximum tolerable activ-
ity safety limits to the blood and bone marrow, even in those with normal renal
function (156). Remnant ablation after surgery can be achieved using recombi-
nant human TSH (rhTSH, Thyrogen R
) (157), which is an attractive option in
the elderly, in whom iatrogenic hypothyroidism after thyroid hormone withdrawal
may be associated with significant morbidity. After surgery and following radioio-
dine administration, patients are maintained on suppressive doses of L-thyroxine
with the desired objective of reducing the serum TSH level to below normal.
The older patient will need to be followed carefully and the dose of L-thyroxine
reduced if cardiac symptoms develop. Acceleration of bone loss, particularly in the
Thyroid Disease in the Elderly 429

postmenopausal woman, may also be a limiting factor in maintaining long-term


TSH suppression.
For medullary carcinoma, the operative procedure of choice is total thy-
roidectomy, since the disease is often multicentric. Medullary carcinomas do not
respond to 131 I therapy, so that patients with inoperable residual or recurrent dis-
ease are treated palliatively with external irradiation or could be considered for a
clinical trial of newer chemotherapeutic agents. Patients should undergo testing
for mutations in the RET protooncogene even if there is no family history of
multiple endocrine neoplasia syndrome type 2, since they may be the proband in
a kindred.
The patient with thyroid lymphoma should have clinical staging carried
out by means of CT or MRI imaging. The survival rate can approach 100%
in response to aggressive therapy with external irradiation in combination with
cyclophosphamide, doxorubicin, vincristine, and prednisone chemotherapy (152).

Outcome
Age at time of diagnosis is an important factor in predicting cancer aggressiveness
and mortality from differentiated thyroid cancer. Individuals who are diagnosed
after the age of 50 years have a higher rate of recurrence and death (Table 4)
(158). The 10-year survival for patients with papillary carcinoma is estimated to
be 97% in those under the age of 45 years at the time of diagnosis and ⬍65%
in those older than 60 years at the time of diagnosis. Similarly, 10-year survival
for patients with follicular carcinoma is estimated to be 98% in those under the
age of 45 years and ⬍57% in those older than 60 years (148,158). It appears that
the biologic behavior of thyroid cancer in older individuals is more aggressive,
suggesting that both the features of the thyroid cancer and delayed diagnosis
contribute to the increased thyroid cancer-specific mortality rate in older people
(159–162).
The 10-year survival rate for patients with medullary carcinoma is 84% in
patients younger than 45 years. Rate of survival declines with increasing age at
time of initial diagnosis, with rates substantially lower for persons over the age of
60 years. In patients in their seventh decade of life, approximately two-thirds will
have persistent disease after surgery (163). Efficacy of surgery can be monitored
postoperatively by measurement of blood calcitonin concentration, both in the
basal state and after stimulation (164). Blood levels of carcinoembryonic antigen
may also be elevated in patients with residual or recurrent medullary carcinoma.
The outcome of anaplastic carcinoma of the thyroid remains unsatisfac-
tory, with rare survivorship more than one year after diagnosis (163). Palliative
relief of compression symptoms can sometimes be achieved by surgery followed
by high-dose (40–60 Gy) external irradiation (165). Chemotherapy with doxoru-
bicin and/or cisplatin may be beneficial in combination with surgery and external
irradiation, and there is increasing use of taxane neoadjuvant therapy (165).
430 Cappola et al.

Compressive Goiter
Subtotal thyroidectomy has traditionally been the recommended therapy for com-
pressive goiter. However, in elderly patients, particularly those who may be at
operative risk, thyroid ablation with large doses of 131 I (i.e., 25–125 mCi) can
produce significant shrinkage of the thyroid with accompanying relief of com-
pressive symptoms such as stridor, dyspnea, and dysphagia (145,166). Recent
studies have demonstrated the efficacy of rhTSH to increase the uptake of radioio-
dine in a goiter, and possibly enhance the efficacy of the therapy (167). However,
iatrogenic transient hyperthyroidism is a potential serious adverse event that may
occur after rhTSH administration (168), so caution is necessary if rhTSH is used
in this setting. Replacement of L-thyroxine is required following surgery, and may
also be needed after radioiodine treatment in order to maintain serum TSH within
the normal range and prevent regrowth of thyroid tissue.

REFERENCES
1. Mochizuki Y, Mowafy R, Pasternack B. Weights of human thyroids in New York
city. Health Phys 1963; 9:1299–1301.
2. Hegedus L, Perrild H, Poulsen LR, et al. The determination of thyroid volume by
ultrasound and its relationship to body weight, age, and sex in normal subjects. J
Clin Endocrinol Metab 1983; 56:260–263.
3. Blumenthal HT, Perlstein IB. The aging thyroid. I. A description of lesions and
an analysis of their age and sex distribution. J Am Geriatr Soc 1987; 35:843–
854.
4. Denham MJ, Wills EJ. A clinico-pathological survey of thyroid glands in old age.
Gerontology 1980; 26:160–166.
5. Barreca T, Franceschini R, Messina V, et al. 24-hour thyroid-stimulating hormone
secretory pattern in elderly men. Gerontology 1985; 31:119–123.
6. Ryan N, Kovacs K, Ezrin C. Thyrotrophs in old age. An immunocytologic study of
human pituitary glands. Endokrinologie 1979; 73:191–198.
7. Van CA, Laurent E, Decoster C, et al. Decreased basal and stimulated thyrotropin
secretion in healthy elderly men. J Clin Endocrinol Metab 1989; 69:177–185.
8. Cuttelod S, Lemarchand-Beraud T, Magnenat P, et al. Effect of age and role
of kidneys and liver on thyrotropin turnover in man. Metabolism 1974; 23:101–
113.
9. Snyder PJ, Utiger RD. Response to thyrotropin releasing hormone (TRH) in normal
man. J Clin Endocrinol Metab 1972; 34:380–385.
10. Utiger RD. Thyrotropin-releasing hormone and thyrotropin secretion. J Lab Clin
Med 1987; 109:327–335.
11. Harman SM, Wehmann RE, Blackman MR. Pituitary-thyroid hormone economy in
healthy aging men: Basal indices of thyroid function and thyrotropin responses to
constant infusions of thyrotropin releasing hormone. J Clin Endocrinol Metab 1984;
58:320–326.
12. Targum SD, Marshall LE, Magac-Harris K, et al. TRH tests in a healthy elderly
population. Demonstration of gender differences. J Am Geriatr Soc 1989; 37:533–
536.
Thyroid Disease in the Elderly 431

13. Hay ID, Klee GG. Thyroid dysfunction. Endocrinol Metab Clin North Am 1988;
17:473–509.
14. Snyder PJ, Utiger RD. Thyrotropin response to thyrotropin releasing hormone in
normal females over forty. J Clin Endocrinol Metab 1972; 34:1096–1098.
15. Herrmann J, Heinen E, Kroll HJ, et al. Thyroid function and thyroid hormone
metabolism in elderly people. Low T3-syndrome in old age? Klin Wochenschr
1981; 59:315–323.
16. Einhorn J. Studies on the effect of thyrotropic hormone on the thyroid function in
man. Acta Radiol Suppl 1958; 160:1–107.
17. Surks MI, Hollowell JG. Age-specific distribution of serum thyrotropin and antithy-
roid antibodies in the US population: Implications for the prevalence of subclinical
hypothyroidism. J Clin Endocrinol Metab 2007; 92:4575–4582.
18. Mariotti S, Franceschi C, Cossarizza A, et al. The aging thyroid. Endocr Rev 1995;
16:686–715.
19. Mariotti S, Barbesino G, Caturegli P, et al. Complex alteration of thyroid function
in healthy centenarians. J Clin Endocrinol Metab 1993; 77:1130–1134.
20. Spencer CA, Hollowell JG, Kazarosyan M, et al. National Health and Nutrition
Examination Survey III thyroid-stimulating hormone (TSH)-thyroperoxidase anti-
body relationships demonstrate that TSH upper reference limits may be skewed by
occult thyroid dysfunction. J Clin Endocrinol Metab 2007; 92:4236–4240.
21. Takashima N, Niwa Y, Mannami T, et al. Characterization of subclinical thyroid
dysfunction from cardiovascular and metabolic viewpoints: The Suita study. Circ J
2007; 71:191–195.
22. Olsen T, Laurberg P, Weeke J. Low serum triiodothyronine and high serum reverse
triiodothyronine in old age: An effect of disease not age. J Clin Endocrinol Metab
1978; 47:1111–1115.
23. Kabadi UM, Rosman PM. Thyroid hormone indices in adult healthy subjects: No
influence of aging. J Am Geriatr Soc 1988; 36:312–316.
24. Nishikawa M, Inada M, Naito K, et al. Age-related changes of serum 3,3’-
diiodothyronine, 3’,5’-diiodothyronine, and 3,5-diiodothyronine concentrations in
man. J Clin Endocrinol Metab 1981; 52:517–522.
25. Burrows AW, Shakespear RA, Hesch RD, et al. Thyroid hormones in the elderly
sick: “T4 euthyroidism.” Br Med J 1975; 4:437–439.
26. Chopra IJ, Solomon DH, Chopra U, et al. Pathways of metabolism of thyroid hor-
mones. Recent Prog Horm Res 1978; 34:521–567.
27. Hansen JM, Skovsted L, Siersbaek-Nielsen K. Age dependent changes in iodine
metabolism and thyroid function. Acta Endocrinol (Copenh) 1975; 79:60–65.
28. Gregerman RI, Gaffney GW, Shock NW, et al. Thyroxine turnover in euthyroid man
with special reference to changes with age. J Clin Invest 1962; 41:2065–2074.
29. Klee GG, Hay ID. Assessment of sensitive thyrotropin assays for an expanded role
in thyroid function testing: Proposed criteria for analytic performance and clinical
utility. J Clin Endocrinol Metab 1987; 64:461–471.
30. Hesch RD, Gatz J, Juppner H, et al. TBG-dependency of age related variations of
thyroxine and triiodothyronine. Horm Metab Res 1977; 9:141–146.
31. Kaptein EM. Thyroid hormone metabolism in illness. In: Hennemann G, ed. Thyroid
Hormone Metabolism. New York, NY: Marcel Dekker, 1986:293–333.
32. Surks MI, Sievert R. Drugs and thyroid function. N Engl J Med 1995; 333:1688–
1694.
432 Cappola et al.

33. Gambert SR. Effect of age on basal and 3,5,3’triiodothyronine (T3) stimulated human
mononuclear cell sodium-potassium adenosine-triphosphatase (Na+ -K+ ATP’ase)
activity. Horm Metab Res 1986; 18:649–650.
34. Gambert SR, Ingbar SH, Hagen TC. Interaction of age and thyroid hormone status
on Na+ -K+ ATPase in rat renal cortex and liver. Endocrinology 1981; 108:27–30.
35. Robuschi G, Safran M, Braverman LE, et al. Hypothyroidism in the elderly Endocr
Rev 1987; 8:142–153.
36. Blumenthal HT, Perlstein IB. The aging thyroid. II. An immunocytochemical anal-
ysis of the age-associated lesions. J Am Geriatr Soc 1987; 35:855–863.
37. Pedersen IB, Knudsen N, Jorgensen T, et al. Thyroid peroxidase and thyroglobulin
autoantibodies in a large survey of populations with mild and moderate iodine
deficiency. Clin Endocrinol (Oxf) 2003; 58:36–42.
38. Spaulding SW. Age and the thyroid. Endocrinol Metab Clin North Am 1987;
16:1013–1025.
39. Chopra IJ. Clinical review 86: Euthyroid sick syndrome: Is it a misnomer? J Clin
Endocrinol Metab 1997; 82:329–334.
40. Simons RJ, Simon JM, Demers LM, et al. Thyroid dysfunction in elderly hospitalized
patients. Effect of age and severity of illness. Arch Intern Med 1990; 150:1249–1253.
41. Adler SM, Wartofsky L. The nonthyroidal illness syndrome. Endocrinol Metab Clin
North Am 2007; 36:657–672.
42. De Groot LJ. Dangerous dogmas in medicine: The nonthyroidal illness syndrome. J
Clin Endocrinol Metab 1999; 84:151–164.
43. Surks MI, Hupart KH, Pan C, et al. Normal free thyroxine in critical nonthyroidal
illnesses measured by ultrafiltration of undiluted serum and equilibrium dialysis. J
Clin Endocrinol Metab 1988; 67:1031–1039.
44. Chopra IJ, Huang TS, Beredo A, et al. Serum thyroid hormone binding inhibitor in
nonthyroidal illnesses. Metabolism 1986; 35:152–159.
45. Slag MF, Morley JE, Elson MK, et al. Hypothyroxinemia in critically ill patients as
a predictor of high mortality. JAMA 1981; 245:43–45.
46. Klemperer JD, Klein I, Gomez M, et al. Thyroid hormone treatment after coronary-
artery bypass surgery. N Engl J Med 1995; 333:1522–1527.
47. Bennett-Guerrero E, Jimenez JL, White WD, et al. Cardiovascular effects of intra-
venous triiodothyronine in patients undergoing coronary artery bypass graft surgery.
A randomized, double-blind, placebo- controlled trial. Duke T3 study group. JAMA
1996; 275:687–692.
48. Brent GA, Hershman JM. Thyroxine therapy in patients with severe nonthyroidal
illnesses and low serum thyroxine concentration. J Clin Endocrinol Metab 1986;
63:1–8.
49. Bartalena L, Brogioni S, Grasso L, et al. Relationship of the increased serum
interleukin-6 concentration to changes of thyroid function in nonthyroidal illness. J
Endocrinol Invest 1994; 17:269–274.
50. Rogy MA, Coyle SM, Oldenburg HS, et al. Persistently elevated soluble tumor
necrosis factor receptor and interleukin-1 receptor antagonist levels in critically ill
patients. J Am Coll Surg 1994; 178:132–138.
51. Pang XP, Hershman JM, Mirell CJ, et al. Impairment of hypothalamic-pituitary-
thyroid function in rats treated with human recombinant tumor necrosis factor-alpha
(cachectin). Endocrinology 1989; 125:76–84.
Thyroid Disease in the Elderly 433

52. Hamblin PS, Dyer SA, Mohr VS, et al. Relationship between thyrotropin and thy-
roxine changes during recovery from severe hypothyroxinemia of critical illness. J
Clin Endocrinol Metab 1986; 62:717–722.
53. Engler D, Burger AG. The deiodination of the iodothyronines and of their derivatives
in man. Endocr Rev 1984; 5:151–184.
54. Van den Beld AW, Visser TJ, Feelders RA, et al. Thyroid hormone concentrations,
disease, physical function, and mortality in elderly men. J Clin Endocrinol Metab
2005; 90:6403–6409.
55. Hollowell JG, Staehling NW, Flanders WD, et al. Serum TSH, T(4), and thyroid anti-
bodies in the United States population (1988 to 1994): National Health and Nutrition
Examination Survey (NHANES III). J Clin Endocrinol Metab 2002; 87:489–499.
56. Parle JV, Franklyn JA, Cross KW, et al. Prevalence and follow-up of abnormal
thyrotrophin (TSH) concentrations in the elderly in the United Kingdom. Clin
Endocrinol (Oxf) 1991; 34:77–83.
57. Cappola AR, Fried LP, Arnold AM, et al. Thyroid status, cardiovascular risk, and
mortality in older adults. JAMA 2006; 295:1033–1041.
58. Diez JJ. Hypothyroidism in patients older than 55 years: An analysis of the etiology
and assessment of the effectiveness of therapy. J Gerontol A Biol Sci Med Sci 2002;
57:M315–M320.
59. Sridama V, McCormick M, Kaplan EL, et al. Long-term follow-up study of compen-
sated low-dose 131I therapy for Graves’ disease. N Engl J Med 1984; 311:426–432.
60. Hirota Y, Tamai H, Hayashi Y, et al. Thyroid function and histology in forty-five
patients with hyperthyroid Graves’ disease in clinical remission more than ten years
after thionamide drug treatment. J Clin Endocrinol Metab 1986; 62:165–169.
61. Doucet J, Trivalle C, Chassagne P, et al. Does age play a role in clinical presentation
of hypothyroidism? J Am Geriatr Soc 1994; 42:984–986.
62. Rai GS, Gluck T, Luttrell S. Clinical presentation of hypothyroidism in older persons.
J Am Geriatr Soc 1995; 43:592–593.
63. Tachman ML, Guthrie GP Jr. Hypothyroidism: Diversity of presentation. Endocr
Rev 1984; 5:456–465.
64. Bastenie PA, Bonnyns M, Vanhaelst L. Natural history of primary myxedema. Am
J Med 1985; 79:91–100.
65. Rosenbaum RL, Barzel US. Levothyroxine replacement dose for primary hypothy-
roidism decreases with age. Ann Intern Med 1982; 96:53–55.
66. Davis FB, LaMantia RS, Spaulding SW, et al. Estimation of a physiologic replace-
ment dose of levothyroxine in elderly patients with hypothyroidism. Arch Intern
Med 1984; 144:1752–1754.
67. Kabadi UM. Optimal daily levothyroxine dose in primary hypothyroidism. Its rela-
tion to pretreatment thyroid hormone indexes. Arch Intern Med 1989; 149:2209–
2212.
68. Roos A, Linn-Rasker SP, van Domburg RT, et al. The starting dose of levothyroxine
in primary hypothyroidism treatment: A prospective, randomized, double-blind trial.
Arch Intern Med 2005; 165:1714–1720.
69. Wartofsky L. Levothyroxine therapy for hypothyroidism: Should we abandon con-
servative dosage titration? Arch Intern Med 2005; 165:1683–1684.
70. Canaris GJ, Manowitz NR, Mayor G, et al. The Colorado thyroid disease prevalence
study. Arch Intern Med 2000; 160:526–534.
434 Cappola et al.

71. Sawin CT, Geller A, Wolf PA, et al. Low serum thyrotropin concentrations as a
risk factor for atrial fibrillation in older persons. N Engl J Med 1994; 331:1249–
1252.
72. Sawin CT, Geller A, Kaplan MM, et al. Low serum thyrotropin (thyroid-stimulating
hormone) in older persons without hyperthyroidism. Arch Intern Med 1991;
151:165–168.
73. Gammage MD, Parle JV, Holder RL, et al. Association between serum free thyroxine
concentration and atrial fibrillation. Arch Intern Med 2007; 167:928–934.
74. Schneider DL, Barrett-Connor EL, Morton DJ. Thyroid hormone use and bone
mineral density in elderly women. Effects of estrogen. JAMA 1994; 271:1245–
1249.
75. Uzzan B, Campos J, Cucherat M, et al. Effects on bone mass of long term treatment
with thyroid hormones: A meta-analysis. J Clin Endocrinol Metab 1996; 81:4278–
4289.
76. Kwaku MP, Burman KD. Myxedema coma. J Intensive Care Med 2007; 22:224–231.
77. Gavin LA. Thyroid crises. Med Clin North Am 1991; 75:179–193.
78. Jordan RM. Myxedema coma. Pathophysiology, therapy, and factors affecting prog-
nosis. Med Clin North Am 1995; 79:185–194.
79. Helfand M. Screening for subclinical thyroid dysfunction in nonpregnant adults: A
summary of the evidence for the U.S. Preventive Services Task Force. Ann Intern
Med 2004; 140:128–141.
80. Surks MI, Ortiz E, Daniels GH, et al. Subclinical thyroid disease: Scientific review
and guidelines for diagnosis and management. JAMA 2004; 291:228–238.
81. Biondi B, Cooper DS. The clinical significance of subclinical thyroid dysfunction.
Endocr Rev 2008; 29:76–131.
82. Bemben DA, Hamm RM, Morgan L, et al. Thyroid disease in the elderly. Part 2.
Predictability of subclinical hypothyroidism. J Fam Pract 1994; 38:583–588.
83. Sawin CT, Castelli WP, Hershman JM, et al. The aging thyroid. Thyroid deficiency
in the Framingham Study. Arch Intern Med 1985; 145:1386–1388.
84. Huber G, Staub JJ, Meier C, et al. Prospective study of the spontaneous course of
subclinical hypothyroidism: Prognostic value of thyrotropin, thyroid reserve, and
thyroid antibodies. J Clin Endocrinol Metab 2002; 87:3221–3226.
85. Diez JJ, Iglesias P. Spontaneous subclinical hypothyroidism in patients older than
55 years: An analysis of natural course and risk factors for the development of overt
thyroid failure. J Clin Endocrinol Metab 2004; 89:4890–4897.
86. Klein I, Ojamaa K. Thyroid hormone and the cardiovascular system. N Engl J Med
2001; 344:501–509.
87. Hak AE, Pols HA, Visser TJ, et al. Subclinical hypothyroidism is an independent
risk factor for atherosclerosis and myocardial infarction in elderly women: The
Rotterdam Study. Ann Intern Med 2000; 132:270–278.
88. Imaizumi M, Akahoshi M, Ichimaru S, et al. Risk for ischemic heart disease and
all-cause mortality in subclinical hypothyroidism. J Clin Endocrinol Metab 2004;
89:3365–3370.
89. Walsh JP, Bremner AP, Bulsara MK, et al. Subclinical thyroid dysfunction as a risk
factor for cardiovascular disease. Arch Intern Med 2005; 165:2467–2472.
90. Vanderpump MP, Tunbridge WM, French JM, et al. The development of ischemic
heart disease in relation to autoimmune thyroid disease in a 20-year follow-up study
of an English community. Thyroid 1996; 6:155–160.
Thyroid Disease in the Elderly 435

91. Parle JV, Maisonneuve P, Sheppard MC, et al. Prediction of all-cause and cardiovas-
cular mortality in elderly people from one low serum thyrotropin result: A 10-year
cohort study. Lancet 2001; 358:861–865.
92. Rodondi N, Newman AB, Vittinghoff E, et al. Subclinical hypothyroidism and the
risk of heart failure, other cardiovascular events, and death. Arch Intern Med 2005;
165:2460–2466.
93. Gussekloo J, van EE, de Craen AJ, et al. Thyroid status, disability and cognitive
function, and survival in old age. JAMA 2004; 292:2591–2599.
94. Dugbartey AT. Neurocognitive aspects of hypothyroidism. Arch Intern Med 1998;
158:1413–1418.
95. Jackson IM. The thyroid axis and depression. Thyroid 1998; 8:951–956.
96. Lindeman RD, Schade DS, LaRue A, et al. Subclinical hypothyroidism in a biethnic,
urban community. J Am Geriatr Soc 1999; 47:703–709.
97. Roberts LM, Pattison H, Roalfe A, et al. Is subclinical thyroid dysfunction in the
elderly associated with depression or cognitive dysfunction? Ann Intern Med 2006;
145:573–581.
98. Cooper DS. Thyroid disease in the oldest old: The exception to the rule. JAMA 2004;
292:2651–2654.
99. Davis PJ, Davis FB. Hyperthyroidism in patients over the age of 60 years. Clinical
features in 85 patients. Medicine (Baltimore) 1974; 53:161–181.
100. Ronnov-Jessen V, Kirkegaard C. Hyperthyroidism—a disease of old age? Br Med J
1973; 1:41–43.
101. Tunbridge WM, Evered DC, Hall R, et al. The spectrum of thyroid disease in a
community: The Whickham survey. Clin Endocrinol (Oxf) 1977; 7:481–493.
102. Furszyfer J, Kurland LT, McConahey WM, et al. Epidemiologic aspects of
Hashimoto’s thyroiditis and Graves’ disease in Rochester, Minnesota (1935–1967),
with special reference to temporal trends. Metabolism 1972; 21:197–204.
103. Frost L, Vestergaard P, Mosekilde L. Hyperthyroidism and risk of atrial fibrillation
or flutter: A population-based study. Arch Intern Med 2004; 164:1675–1678.
104. Martin FI, Deam DR. Hyperthyroidism in elderly hospitalised patients. Clinical
features and treatment outcomes. Med J Aust 1996; 164:200–203.
105. Diez JJ. Hyperthyroidism in patients older than 55 years: An analysis of the etiology
and management. Gerontology 2003; 49:316–323.
106. Harjai KJ, Licata AA. Effects of amiodarone on thyroid function. Ann Intern Med
1997; 126:63–73.
107. Bartalena L, Brogioni S, Grasso L, et al. Treatment of amiodarone-induced thyrotox-
icosis, a difficult challenge: Results of a prospective study. J Clin Endocrinol Metab
1996; 81:2930–2933.
108. Dickstein G, Shechner C, Adawi F, et al. Lithium treatment in amiodarone-induced
thyrotoxicosis. Am J Med 1997; 102:454–458.
109. Banovac K, Papic M, Bilsker MS, et al. Evidence of hyperthyroidism in apparently
euthyroid patients treated with levothyroxine. Arch Intern Med 1989; 149:809–812.
110. Smallridge RC. Thyrotropin-secreting pituitary tumors. Endocrinol Metab Clin
North Am 1987; 16:765–792.
111. Marsden P, Facer P, Acosta M, et al. Serum triiodothyronine in solitary autonomous
nodules of the thyroid. Clin Endocrinol (Oxf) 1975; 4:327–330.
112. Kawabe T, Komiya I, Endo T, et al. Hyperthyroidism in the elderly. J Am Geriatr
Soc 1979; 27:152–155.
436 Cappola et al.

113. Griffin MA, Solomon DH. Hyperthyroidism in the elderly. J Am Geriatr Soc 1986;
34:887–892.
114. Trivalle C, Doucet J, Chassagne P, et al. Differences in the signs and symptoms of
hyperthyroidism in older and younger patients. J Am Geriatr Soc 1996; 44:50–53.
115. Thomas FB, Mazzaferri EL, Skillman TG. Apathetic thyrotoxicosis: A distinctive
clinical and laboratory entity. Ann Intern Med 1970; 72:679–685.
116. Tibaldi JM, Barzel US, Albin J, et al. Thyrotoxicosis in the very old. Am J Med
1986; 81:619–622.
117. Cappola AR, Ladenson PW. Hypothyroidism and atherosclerosis. J Clin Endocrinol
Metab 2003; 88:2438–2444.
118. Helfand M, Redfern CC. Clinical guideline, part 2. Screening for thyroid disease:
An update. American College of Physicians. Ann Intern Med 1998; 129:144–158.
119. Livingston EH, Hershman JM, Sawin CT, et al. Prevalence of thyroid disease and
abnormal thyroid tests in older hospitalized and ambulatory persons. J Am Geriatr
Soc 1987; 35:109–114.
120. Cooper DS. Hyperthyroidism. Lancet 2003; 362:459–468.
121. Solomon B, Glinoer D, Lagasse R, et al. Current trends in the management of Graves’
disease. J Clin Endocrinol Metab 1990; 70:1518–1524.
122. Cooper DS. Antithyroid drugs. N Engl J Med 2005; 352:905–917.
123. Cooper DS, Goldminz D, Levin AA, et al. Agranulocytosis associated with antithy-
roid drugs. Effects of patient age and drug dose. Ann Intern Med 1983; 98:26–29.
124. Yamada T, Aizawa T, Koizumi Y, et al. Age-related therapeutic response to antithy-
roid drug in patients with hyperthyroid Graves’ disease. J Am Geriatr Soc 1994;
42:513–516.
125. Nakazawa HK, Sakurai K, Hamada N, et al. Management of atrial fibrillation in the
post-thyrotoxic state. Am J Med 1982; 72:903–906.
126. Fuster V, Ryden LE, Cannom DS, et al. ACC/AHA/ESC 2006 guidelines for the
management of patients with atrial fibrillation–executive summary: A report of
the American College of Cardiology/American Heart Association Task Force on
Practice Guidelines and the European Society of Cardiology Committee for Practice
Guidelines (Writing Committee to Revise the 2001 Guidelines for the Management
of Patients With Atrial Fibrillation). J Am Coll Cardiol 2006; 48:854–906.
127. Kellett HA, Sawers JS, Boulton FE, et al. Problems of anticoagulation with warfarin
in hyperthyroidism. Q J Med 1986; 58:43–51.
128. Nayak B, Burman K. Thyrotoxicosis and thyroid storm. Endocrinol Metab Clin
North Am 2006; 35:663–686.
129. Stott DJ, McLellan AR, Finlayson J, et al. Elderly patients with suppressed serum
TSH but normal free thyroid hormone levels usually have mild thyroid overactivity
and are at increased risk of developing overt hyperthyroidism. Q J Med 1991; 78:77–
84.
130. Eggertsen R, Petersen K, Lundberg PA, et al. Screening for thyroid disease in a
primary care unit with a thyroid stimulating hormone assay with a low detection
limit. BMJ 1988; 297:1586–1592.
131. Meyerovitch J, Rotman-Pikielny P, Sherf M, et al. Serum thyrotropin measurements
in the community: Five-year follow-up in a large network of primary care physicians.
Arch Intern Med 2007; 167:1533–1538.
132. Woeber KA. Observations concerning the natural history of subclinical hyperthy-
roidism. Thyroid 2005; 15:687–691.
Thyroid Disease in the Elderly 437

133. Cappola AR. Subclinical Thyroid Dysfunction and the Heart. J Clin Endocrinol
Metab 2007; 92:3404–3405.
134. Bauer DC, Ettinger B, Nevitt MC, et al. Risk for fracture in women with low serum
levels of thyroid-stimulating hormone. Ann Intern Med 2001; 134:561–568.
135. Faber J, Jensen IW, Petersen L, et al. Normalization of serum thyrotrophin by
means of radioiodine treatment in subclinical hyperthyroidism: Effect on bone loss
in postmenopausal women. Clin Endocrinol (Oxf) 1998; 48:285–290.
136. Kalmijn S, Mehta KM, Pols HA, et al. Subclinical hyperthyroidism and the risk of
dementia. The Rotterdam study. Clin Endocrinol (Oxf) 2000; 53:733–737.
137. Cooper DS. Approach to the patient with subclinical hyperthyroidism. J Clin
Endocrinol Metab 2007; 92:3–9.
138. U.S. Preventive Services Task Force. Screening for thyroid disease: Recommenda-
tion statement. Ann Intern Med 2004; 140:125–127.
139. Committee on Medicare Coverage of Routine Thyroid Screening. Medicare Cov-
erage of Routine Screening for Thyroid Dysfunction. Washington, DC: National
Academies Press, 2003.
140. Clinical guideline, part 1. Screening for thyroid disease. American College of Physi-
cians. Ann Intern Med 1998; 129:141–143.
141. Ladenson PW, Singer PA, Ain KB, et al. American Thyroid Association guidelines
for detection of thyroid dysfunction. Arch Intern Med 2000; 160:1573–1575.
142. Mortensen JD, Woolner LB, Bennett WA. Gross and microscopic findings in
clinically normal thyroid glands. J Clin Endocrinol Metab 1955; 15:1270–
1280.
143. Tan GH, Gharib H. Thyroid incidentalomas: Management approaches to nonpalpable
nodules discovered incidentally on thyroid imaging. Ann Intern Med 1997; 126:226–
231.
144. DeGroot LJ. Clinical review 2: Diagnostic approach and management of patients
exposed to irradiation to the thyroid. J Clin Endocrinol Metab 1989; 69:925–928.
145. Huysmans DA, Hermus AR, Corstens FH, et al. Large, compressive goiters treated
with radioiodine. Ann Intern Med 1994; 121:757–762.
146. Frates MC, Benson CB, Doubilet PM, et al. Prevalence and distribution of carci-
noma in patients with solitary and multiple thyroid nodules on sonography. J Clin
Endocrinol Metab 2006; 91:3411–3417.
147. Belfiore A, La Rosa GL, La Porta GA, et al. Cancer risk in patients with cold thyroid
nodules: Relevance of iodine intake, sex, age, and multinodularity. Am J Med 1992;
93:363–369.
148. Hundahl SA, Fleming ID, Fremgen AM, et al. A National Cancer Data Base report on
53,856 cases of thyroid carcinoma treated in the U.S., 1985–1995 [see comments].
Cancer 1998; 83:2638–2648.
149. Kobayashi T, Asakawa H, Umeshita K, et al. Treatment of 37 patients with anaplastic
carcinoma of the thyroid. Head Neck 1996; 18:36–41.
150. Holm LE, Blomgren H, Lowhagen T. Cancer risks in patients with chronic lympho-
cytic thyroiditis. N Engl J Med 1985; 312:601–604.
151. Cooper DS, Doherty GM, Haugen BR, et al. Management guidelines for patients
with thyroid nodules and differentiated thyroid cancer. Thyroid 2006; 16:109–142.
152. Matsuzuka F, Miyauchi A, Katayama S, et al. Clinical aspects of primary thyroid
lymphoma: Diagnosis and treatment based on our experience of 119 cases. Thyroid
1993; 3:93–99.
438 Cappola et al.

153. Castro MR, Caraballo PJ, Morris JC. Effectiveness of thyroid hormone suppressive
therapy in benign solitary thyroid nodules: a meta-analysis. J Clin Endocrinol Metab
2002; 87:4154–4159.
154. Uruno T, Miyauchi A, Shimizu K, et al. Favorable surgical results in 433 elderly
patients with papillary thyroid cancer. World J Surg 2005; 29:1497–1501.
155. Mazzaferri EL. An overview of the management of papillary and follicular thyroid
carcinoma. Thyroid 1999; 9:421–427.
156. Tuttle RM, Leboeuf R, Robbins RJ, et al. Empiric radioactive iodine dosing regimens
frequently exceed maximum tolerated activity levels in elderly patients with thyroid
cancer. J Nucl Med 2006; 47:1587–1591.
157. Pacini F, Ladenson PW, Schlumberger M, et al. Radioiodine ablation of thyroid
remnants after preparation with recombinant human thyrotropin in differentiated
thyroid carcinoma: Results of an international, randomized, controlled study. J Clin
Endocrinol Metab 2006; 91:926–932.
158. Schlumberger MJ. Papillary and follicular thyroid carcinoma. N Engl J Med 1998;
338:297–306.
159. Coburn MC, Wanebo HJ. Age correlates with increased frequency of high risk factors
in elderly patients with thyroid cancer. Am J Surg 1995; 170:471–475.
160. Lin JD, Chao TC, Chen ST, et al. Characteristics of thyroid carcinomas in aging
patients. Eur J Clin Invest 2000; 30:147–153.
161. Vini L, Hyer SL, Marshall J, et al. Long-term results in elderly patients with differ-
entiated thyroid carcinoma. Cancer 2003; 97:2736–2742.
162. Falvo L, Catania A, Sorrenti S, et al. Prognostic significance of the age factor in the
thyroid cancer: Statistical analysis. J Surg Oncol 2004; 88:217–222.
163. Thoresen SO, Akslen LA, Glattre E, et al. Survival and prognostic factors in differ-
entiated thyroid cancer—a multivariate analysis of 1,055 cases. Br J Cancer 1989;
59:231–235.
164. Ball DW. Medullary thyroid cancer: Monitoring and therapy. Endocrinol Metab Clin
North Am 2007; 36:823–837.
165. Pudney D, Lau H, Ruether JD, et al. Clinical experience of the multimodality man-
agement of anaplastic thyroid cancer and literature review. Thyroid 2007; 17:1243–
1250.
166. Bonnema SJ, Nielsen VE, Hegedus L. Long-term effects of radioiodine on thyroid
function, size and patient satisfaction in non-toxic diffuse goitre. Eur J Endocrinol
2004; 150:439–445.
167. Bonnema SJ, Nielsen VE, Boel-Jorgensen H, et al. Improvement of goiter volume
reduction after 0.3 mg recombinant human thyrotropin-stimulated radioiodine ther-
apy in patients with a very large goiter: A double-blinded, randomized trial. J Clin
Endocrinol Metab 2007; 92:3424–3428.
168. Albino CC, Mesa CO Jr, Olandoski M, et al. Recombinant human thyrotropin as
adjuvant in the treatment of multinodular goiters with radioiodine. J Clin Endocrinol
Metab 2005; 90:2775–2780.
Index

Acute infectious thyroiditis and family planning, 53


diagnosis, 110 as primary therapy, 52–55
epidemiology, 110 side effects, 55–59
pathophysiology, 110 Anti-TSH receptor antibody measurements, 48
treatment, 111 APECED (Autoimmune Polyendocrinopathy
Acute suppurative thyroiditis, 350–351 Candidiasis Ectodermal Dystrophy)
Addison’s disease, 50, 346 syndrome, 346
Adrenal medullary disease, 299 Atrial arrhythmias, 41
Agranulocytosis, 57, 60 Atrial diastolic dysfunction, 42
Aluminum hydroxide, 173 Atrial fibrillation, 60, 412–413
Amennorhea, 44 Atrophic gastritis, 174
Amiodarone, 11, 13, 47, 115, 170, 173 Atypical thyroiditis. See Silent or “painless”
Amiodarone-induced hyperthyroidism, 11 thyroiditis
diagnosis, 117 Autoimmune gastritis, 174
epidemiology, 115–116
pathophysiology, 116–117 Benign nodules, 209
treatment, 117–118 Beta-adrenergic blocking agents, 106, 109, 115,
Amiodarone-induced thyrotoxicosis (AIT), 129, 356, 377
106 Bexarotene, 171
Anaplastic thyroid carcinoma (ATC), 240 Bleomycin, 324
diagnosis, 310–312 “Block-replacement’’ regimen, 54
incidence and demographics, 308–309 Blood dosimetry, 267–268
natural history and mortality rates, 312 BRAF mutation, 248
origin, 309 BRAF V600E mutation, 258
pathology, 309–310 Bromocriptine, 126
prognostic factors, 313
treatment Café-au-lait spots, 81
chemotherapy, 314–315 Calcitonin, 301
combination doxorubicin and in the evaluation of thyroid nodules, 303
hyperfractionated radiation therapy, pentagastrin-stimulated screening of, 304–305
315–316 secretion, physiological stimuli to, 28
external radiotherapy, 314 and tumor mass, 303
surgery, 313–314 Calcium, impact on thyroid hormones, 42
Anticytoplasmic neutrophil antibodies (ANCA), Calcium carbonate, 173
58 Calcium stimulation test, 303–304
Antithyroglobulin antibodies, 19 Cancer and radioactive iodine therapy, 66
Antithyroid arthritis syndrome, 58 Carbamazepine, 173
Antithyroid drugs, 115 Cat-scratch disease, 102
clinical pharmacology, 51–52 C-cell hyperplasia (CCH), 29, 299
in clinical practice, 52 C cells, thyroid, 28
drug-related leukopenia, 57 Celiac disease, 174

439
440 Index

Cervical lymphadenopathy, 110 hypothyroidism


Cervical lymph node metastases, 302 diagnosis, 399–401
Cervical ultrasonography, 277–278 epidemiology, 399
Chemotherapy, 273–274, 324 myxedema coma, 402–404
Chernobyl children, 258 pathophysiology, 399
Children and thyroid functions, 335 subclinical, 404–407
Cholestyramine (or colestipol), 74, 173 treatment, 401–402
Choriocarcinoma, 128 nodular thyroid disease and neoplasia
Chronic lymphocytic thyroiditis, 346 diagnosis, 417–420
Coccidioides immitis, 110 epidemiology, 416
Columnar cell variant, 248 pathophysiology, 416–417
Congenital hypothyroidism, 335–339 treatment, 420–421
Corticosteroid therapy, 65, 104, 106, 115 occurrence and survivorship of thyroid
Cowden syndrome, 357 malignancy, 418
Coxsackievirus, 102 outcome of anaplastic carcinoma of the
Cranial synostosis, 345 thyroid, 421
Creeping thyroiditis, 102 screening, 415–416
Cushing’s syndrome, 302 Embryogenesis, 332
Cyclophosphamide, 58 Encapsulated papillary carcinoma, 247
Cytokine-induced hyperthyroidism Euthyroid hyperthyroxinemia, 9–10
diagnosis, 120 Euthyroid hypothyroxinemia, 11
epidemiology, 119 Euthyroidism and thyroid autoimmunity,
pathophysiology, 119 363–364
treatment, 120 Euthyroid sick syndrome, 50, 160, 398
Cytomel, 164 External beam radiation therapy, 273

De Morsier syndrome, 337 Familial dysalbuminemic hyperthyroxinemia


Diffuse follicular variant papillary carcinoma, (FDH), 8
247–248 Familial medullary thyroid carcinoma, 304–305
Diffuse sclerosing variant, 248 Ferrous sulfate, 173
Diltiazem, 60 Fetal hyperthyroidism, 344, 379–380
Diphenylhydantoin, 13 Fetal hypothyroidism, 333
Diplopia, 68 Fetal thyroid function, 332–333
Distant metastases, 245 Fetal T4 metabolism, 333
Dopamine agonists, 124 Fiber supplements, 173
Down syndrome, 346 Fibroblasts, 68
Doxorubicin, 314 Fine-needle aspiration biopsy, 28, 75, 102, 203,
and hyperfractionated radiation therapy, 209, 251–253, 311, 320, 358, 419
315–316 indications, 213
reasons for repetition, 219
Elderly results, 214–217
compressive goiter, 422 stains, 214
hyperthyroidism technique, 214
atrial fibrillation, 412–413 ultrasonographically guided, 217–218
diagnosis, 409–411 Follicular cell neoplasms, 217
epidemiology, 407 Follicular thyroid carcinoma
pathophysiology, 407–408 diagnosis
subclinical, 413–415 clinical presentation, 250–251
T3 and T4 toxicosis, 408 evaluating patients with a history of
thyroid storm, 413 head-and-neck irradiation, 253
treatment, 411–412 FNA biopsy, 251–253
Index 441

incidence, 242 orbitopathy, 168


pathology, 249–250 pathophysiology, 40
prognosis during pregnancy, 370–371
patient variables influencing, 254–255 presentation, 39
treatment variables influencing, 259 signs and symptoms, 41–46
tumor variables influencing, 255–259 treatment
treatment antithyroid drugs, 51–59
chemotherapy, 273–274 beta-adrenergic antagonist drugs, 60–61
completion thyroidectomy, 261 choice of, 68
external beam radiation therapy, 273 131 I therapy, 61–66

follow up, 274–278 lithium, 61


gamma knife therapy, 273 potassium perchlorate, 61
ipsilateral lobectomy, 260 thyroidectomy, 67
isthmusectomy, 260
radioiodine (131 I) therapy, 262–272 Hamburger thyrotoxicosis, 121
subtotal lobectomy, 260 Hashimoto’s disease, 19, 28, 161, 212, 244, 317,
surgical complications, 262 319
thyroidectomy during pregnancy, 262 Hashimoto’s thyroiditis, 48–49, 105, 319, 321,
thyroid hormone therapy, 272–273 364, 368
total or near-total thyroidectomy, 261 Hashitoxicosis, 346
Follicular variant of PTC (FVPC), 243, 247 Helicobacter pylori-related gastritis, 174
Free T3 index, 8 Hepatitis C and thyroid abnormalities, 119
Free T3/T4 measurements, 4–8 Hodgkin’s disease, 253
Furosemide, 13 Hook effect, 18
Hormone-binding protein concentrations, 7
Galactorrhea, 349 Hormone specificity “spillover” syndrome, 363
Gallium-67 (67 Ga), 312 24-hour radioiodine uptake test, 47–48, 62, 103,
Gamma knife therapy, 273 112, 335
Gestation and thyroid hormone functions, Human chorionic gonadotrophin (hCG),
361–363 128–129
Glomerulonephritis, 58 Hürthle cells, 106
Glucocorticoid therapy, 13, 58–59, 173 carcinomas, 240
Gluten sensitive enteropathy, 174 neoplasm, 214, 217
Goiter, 54, 318, 346. See also Thyroid gland tumors, 252
enlargement (goiter) Hyperdefecation, 42
elderly, 422 Hyperparathyroidism, 299
Gonadal damage, 271 Hyperthyroidism, 22
G proteins, 75, 81, 206 caused by thyroid cancer, 130–131
Granulocyte-colony stimulating factor (G-CSF), causes, 351–352
57 clinical effects, 41
Granulocyte counts, 57 clinical manifestations, 352
Graves’ disease, 8, 20–21, 26–27, 339, 346 diagnosis, 26–27
association with ophthalmopathy, 39, 49, 121 drug-induced, 112–122
treatment, 68–70 due to exogenous thyroid hormone, 121–122
in elderly, 410–411 in elderly
epidemiology, 40 atrial fibrillation, 412–413
impact of cigarette smoking and stressful life diagnosis, 409–411
events, 40 epidemiology, 407
laboratory diagnosis, 47–51 pathophysiology, 407–408
pitfalls, 50–51 subclinical, 413–415
‘NO SPECS’ classification of eye changes, 69 T3 and T4 toxicosis, 408
442 Index

Hyperthyroidism (cont.) and Hodgkin’s disease, 347


thyroid storm, 413 infections, 152
treatment, 411–412 laboratory evaluation, 349
of extrathyroid origin, 127–131 musculoskeletal disorders, 152
Graves’ disease, 351 neurological disorders, 151
laboratory evaluation, 352–353 pitfalls in management of
pediatric, 351–352 changing thyroid hormone requirements,
during pregnancy 172–175
fetal or neonatal, 379–381 consequences of excess thyroid hormone
Graves’ disease, 370–371 replacement, 178–179
hCG-induced, 369–370 depressed patients, 178
hyperemesis gravidarum, 369–370 desiccated thyroid, 175–176
pregnancy outcome, 372 obesity patients, 177–178
therapy, 372–378 patients allergic to thyroid hormone tablets,
therapy, 353–357 176–177
thyrotrophin-induced, 122–127 patients unable to take oral medications,
treatment, 27 177
Hyperthyrotropinemia patients who miss LT4 dose, 175
causes, 16–17 patients with premenstrual syndrome, 178
isolated, 339–343 presenting symptoms post LT4
Hyperthyroxinemia, 8 replacement, 172
Hypokalemia, 43 surgery, 177
Hypokalemic periodic paralysis, 43 during pregnancy
Hypophosphatemia, 44 diagnosis, 365
Hypothalamic-pituitary-thyroid axis, physiology overt, 364
of, 2–3 pregnancy outcome, 365–367
Hypothalamic tripeptide thyrotropin-releasing treatment, 367–369
hormone (TRH), 3 psychiatric disturbances, 151
Hypothyroidism, 22 pulmonary abnormalities with, 149–150
cardiovascular abnormalities with, 150 screening for, 162
clinical manifestations secondary and tertiary, 347
myxedema coma, 158 subclinical (SCH), 152–158
overt, 147–152 treatment, 25–26
subclinical, 152–158 patient-oriented approach, 165–171
conditions associated with an increased risk thyroid hormone replacement therapy,
of, 163 162–165
diagnosis, 24–25 Hypothyrotropinemia
assay analysis, 158–161 causes, 15–16
measurement of antithyroid antibodies, 161 Hypothyroxinemia, 12–13, 364
drug-induced
treatment, 170–171 Iatrogenic thyrotoxicosis, 8, 177
in elderly 123 Iimaging, 211
diagnosis, 399–401 Immunoassay methods, of free hormone
epidemiology, 399 concentration, 6
myxedema coma, 402–404 Immunoglobulins, 320
pathophysiology, 399 Immunometric (IMA) methods, 13
subclinical, 404–407 Incidentalomas, 225–227
treatment, 401–402 Infants and thyroid functions, 335
endocrine abnormalities with, 150–151 Infectious thyroiditis. See Acute infectious
epidemiology, 146–147 thyroiditis
etiology, 147 Infiltrative diseases, 102
Index 443

Inflammatory thyroiditis, 8 Lanreotide, 124


Insular carcinoma, 249 Large-cell carcinomas, 309
Interferon-alpha, 119 Leukopenia, 57
Interferon therapy–induced hyperthyroidism, Levothyroxine (LT4). See LT4 replacement
119–120 therapy
Interleukin-6 (IL-6), 13, 116–117 Levothyroxine sodium, 25–26
Iodide-induced hyperthyroidism, 8 Liquid chromatography-tandem mass
Iodides, during pregnancy, 377 spectrometry, 5
Iodine-induced hyperthyroidism Lithium, 61, 268
diagnosis, 114–115 Lithium-associated thyrotoxicosis, 120
epidemiology, 112–113 Liver function tests, 42
neonatal, 344 L-thyroxine therapy, 54, 67, 79, 121, 408
pathophysiology, 113–114 LT4/LT3 combination therapy, 172
therapy, 115 LT4 replacement therapy, 8, 154, 156, 163–166,
Iodine supplements, dietary, 113 365–366, 368
Iodothyronine concentration, measurement of absorption of, 165
total, 3–4 amiodarone-induced hypothyroidism, 170
Iopanoate (Telepaque), 74 brand-name preparations, 164–165
Iopanoic acid, 124 complications, 273
Ipodate, 74, 173 dosages, 166
Ipsilateral lobectomy, 260 drugs that decrease absorption, 173
Ischemic optic neuropathy, 68 drugs that increase absorption, 173
Isthmusectomy, 260 myxedema coma, 171
131 I therapy, 61–66, 72, 77–78, 81–83, 224–225, overt primary hypothyroidism, 166
355–356, 378, 412. See also pitfalls, 176–179
Radioiodine (131 I) therapy postradioiodine hypothyroidism, 168
subclinical hypothyroidism (SCH), 166–168
Jod-Basedow phenomenon, 113 suppression of TSH, 272
Lugol’s solution, 74, 413
Klinefelter syndrome, 346 Lung metastases, 248
Lymphadenopathy, 45
Laboratory evaluation Lymph node metastases, 245
of thyroid disease Lymphocytic thyroiditis. See Silent or “painless”
hyperthyroidism, 26–27 thyroiditis
hypothyroidism, 24–26 Lymphoma, 317–318
screening and case finding, 23–24
of thyroid function Macrofollicular PTC, 247
assays of thyroid-stimulating hormone, Malignant pseudothyroiditis, 102, 130, 310
13–15 Maternal-fetal T4 transfer, 333, 366–367
causes of hyperthyrotropinemia, 16–17 McCune-Albright syndrome, 81
causes of hypothyrotropinemia, 15–16 MCT8 thyroid hormone transporter, 338
reverse T3 (rT3), 21 Means-Lerman “scratch” murmur, 41
T4 and T3 concentrations, 4–8 Medullary thyroid carcinoma (MTC), 29–31,
causes of decrease, 11–13 207
causes of increase, 8–11 association of somatic mutations and
thyroglobulin (Tg), 17–19 prognosis, 300–301
thyroid autoantibodies, 19–21 classification, 298
thyroid hormone effects in extrapituitary demographics, 297
tissues, 21–22 diagnosis, 302–304
total serum iodothyronine concentrations, elderly, 421
3–4 EUROMEN study, 308
444 Index

Medullary thyroid carcinoma (MTC) (cont.) Noonan syndrome, 346


family screening, 307–308 Normal aging and thyroid function
Gubbio Conference recommendations, autoimmunity, 398
307–308 morphological changes, 394
multiple endocrine neoplasia type 2 (MEN-2) T4 and T3, 397–398
syndrome TRH–TSH axis, 394–395
familial, 298 TSH, 395–396
genetic alterations, 299 ‘NO SPECS’ classification of eye changes, of
type 2A, 298 Graves’ disease, 69
type 2B, 298 NTRK1 mutation, 248
parathyroid hyperplasia, 299
pathology, 301–302 Occult subacute thyroiditis. See Silent or
pediatric, 360 “painless” thyroiditis
pheochromocytomas, 299–300 Octreotide, 124, 126
prevalence, 297 Omeprazole-induced hypergastrinemia, 29
prognosis, 304–305 Omeprazole stimulation test, 304
sporadic, 300 Oragraffin. See Ipodate
therapy, 305–306 Orbital radiotherapy, 69
follow-up, 306–307 Ovarian teratoma, 128
tumor location, 299 Overt primary hypothyroidism
MEN2A gene, 30 LT4 replacement therapy, 166
Metaiodobenzyguanidine (MIBG), 306 Oxyphilic PTC, 249
Metastases, 245, 257, 301–302, 311
Metastatic lesions, 251 Painless thyroiditis. See Silent or “painless”
Metastatic lymph nodes, 256–257 thyroiditis
Metformin, 16, 174 Papillary cancer within a thyroglossal duct, 247
Methimazole (MMI), 51–53, 55–56, 59, 65, Papillary carcinomas, 127
334–335, 353–355, 372–373, 378–380 Papillary microcarcinoma, 245–246
Metoprolol, 60 Papillary thyroid carcinoma (PTC)
Mitral valve prolapse, 41 diagnosis
Mucosa-associated lymphoid tissue (MALT), clinical presentation, 250–251
317 evaluating patients with a history of
Multikinase inhibitors, 306 head-and-neck irradiation, 253
Multinodular goiter (MNG), 110, 205 FNA biopsy, 251–253
management, 223–225 familial, 242
Myxedema coma, 171 FNA biopsy, 243
imaging studies, 278–280
Nadolol, 60 incidence, 242
Neonatal hyperthyroidism, 343–346 microscopic appearance, 243
Neonatal thyroid function, 334 multiple tumor foci, 244
in premature infants, 334–335 pathology, 243–249
Neonates and TSH level, 16 prognosis, 244
Nephrotic syndrome, 174 patient variables influencing, 254–255
Nodular thyroid disease, 203 treatment variables influencing, 259
incidentalomas, 225–227 tumor variables influencing, 255–259
thyroid gland enlargement (goiter), 204–225 sporadic, 242
Non-Hodgkin’s lymphomas, 319 treatment
Nonsteroidal anti-inflammatory drugs chemotherapy, 273–274
(NSAIDs), 104 completion thyroidectomy, 261
Nonthyroidal illness, 15, 398–399 external beam radiation therapy, 273
Index 445

follow up, 274–278 epidemiology, 107


gamma knife therapy, 273 pathophysiology, 107
ipsilateral lobectomy, 260 during pregnancy, 382–385
isthmusectomy, 260 treatment, 109
radioiodine (131 I) therapy, 262–272 Postradioiodine hypothyroidism, LT4
subtotal lobectomy, 260 replacement therapy, 168
surgical complications, 262 Potassium iodate, 113
thyroidectomy during pregnancy, 262 Potassium iodide, 114
thyroid hormone therapy, 272–273 Potassium perchlorate, 61
total or near-total thyroidectomy, 261 Prednisone, 65, 104, 117, 345
Papillary thyroid microcarcinoma (PTMC), 239 Pregnancy
Parathyroid hyperplasia, 299 euthyroidism and thyroid autoimmunity,
PAX8, 332 363–364
Pediatric thyroid disorders glomerular filtration rate, 368
in childhood and adolescence guidelines for clinical management of
chronic lymphocytic thyroiditis, 346 maternal hyperthyroidism, 376
clinical manifestations, 348–349, 352 guidelines for clinical management of
drug-induced disorder, 347 maternal hypothyroidism, 369
hyperthyroidism, 351–352 hyperthyroidism
iodine deficiency, 347 complications reported, 372
laboratory evaluation, 349, 352–353 fetal or neonatal, 379–381
other causes, 347 Graves’ disease, 370–371
painful thyroid enlargement, 350–351 hCG-induced, 369–370
secondary or tertiary, 347 hyperemesis gravidarum, 369–370
therapy, 349–350, 353–357 outcome, 372
thyroid dysgenesis, 347 therapy, 372–378
thyroid hormone resistance, 347–348 hypothyroidism
infants complications reported, 366
congenital hypothyroidism, 335–339 diagnosis, 365
isolated hyperthyrotropinemia, 339–343 outcome, 365–367
neonatal hyperthyroidism, 343–346 overt, 364
thyroid nodules treatment, 367–369
clinical evaluation, 358 renal iodine clearance, 363
laboratory evaluation, 358 thyroid hormone physiology, 361–363
other screening tools, 358 Premature infants, thyroid function in,
prognosis, 360 334–335
therapy, 359–360 Pretibial myxedema, 45–46, 70
Perchlorate, 74 Primary biliary cirrhosis, 42
Percutaneous ethanol injection (PEI), 79, Primary thyroid lymphoma
222–223 clinical features, 318–319
Percutaneous laser thermal ablation, 223 demographics, 316–317
Phenobarbital, 173 diagnosis, 319–320
Phenytoin, 173 and Hashimoto’s thyroiditis, 317
Pheochromocytomas, 299–300 imaging studies, 320–321
PIT-1, 337 incidence, 316
Pituitary or hypothalamic disease, 25 pathology, 317–318
Pituitary responsiveness, fetal, 332 prognostic features, 322–325
Placenta and thyroid functions, 333–334 staging of, 322
Postpartum thyroiditis, 104 PROP-1, 337
diagnosis, 107–109 Propranolol, 13, 44, 60, 74, 173, 413
446 Index

Propylthiouracil (PTU), 13, 74, 173, 334–335, RhTSH-stimulated Tg measurements,


353–355, 372–373, 378–380 278
Proton pump inhibitors, 173 Rifampin, 173
Psammoma bodies, 243
P53 suppressor gene, 310 Salicylates, 104
PTU-induced hepatitis, 59 Salvage therapy, 325
Pulmonary hypertension, 42 Saturated solution of potassium iodide (SSKI),
67
Quantitative tumor dosimetry, 268 Scintigraphy
indications, 210–211
Radiation-induced thyroid carcinoma results, 211–212
epidemiology, 242–243 technique, 211
Radiation–induced thyrotoxic thyroiditis, Screening, for thyroid disease, 23–24
external beam, 111 Selenium, 109
Radiation thyroiditis Septo-optic dysplasia, 337
diagnosis, 111–112 Sertraline, 13, 174
epidemiology, 111 Sevelamer, 173
pathophysiology, 111 Silent or “painless” thyroiditis, 161, 169
trauma-induced thyroiditis, 112 diagnosis, 105–106
treatment, 112 epidemiology, 105
Radioiodine (131 I) therapy pathophysiology, 105
and ablation of thyroid bed, 262–263 treatment, 106
amount estimation for remnant ablation, 266 Small-cell carcinomas, 309
diagnostic (“pretreatment”) whole-body scan, Sodium ipodate, 10
266–266 Solid or trabecular variant PTC, 248
immediate complications, 268–269 Solitary autonomous thyroid nodules
infertility and gonadal failure, 271–272 clinical considerations, 76
lacrimal duct obstruction, 271 diagnosis, 76–77
leukemia and second primary malignancies pathogenesis, 75
after, 270–271 pathology, 75
with lithium, 268 treatment, 77–79
parotid dysfunction, 269 Somatic RET mutation, 300–301
posttreatment scans and false-positive scans, Somatostatin analogues, 123–124
266–267 Soy supplements, 173
preparation, 263–265 Spermatogenesis, impaired, 45
radiation pneumonitis, 269 Splenomegaly, 45, 58
sodium-iodide symporter activity in PTC, 262 Spontaneously resolving lymphocytic
stunning effect, 265 thyroiditis. See Silent or “painless”
treatment of residual or recurrent carcinoma thyroiditis
with, 267–268 Staphylococcus, 110
Radionuclide scanning, 321 Streptococcus, 110
Raloxifene, 10, 16, 173 Struma ovarii tumor
Ras oncogenes, 206 diagnosis, 127–128
Raynaud’s phenomenon, 60 epidemiology, 127
Recombinant human TSH (rhTSH), 225 pathophysiology, 127
Renal iodine clearance, during pregnancy, 363 treatment, 128
Resin T3 uptake test, 47 Subacute thyroiditis, 351
RET protooncogene, 421 diagnosis, 102–103
RET/PTC gene rearrangement, 248 epidemiology, 101–102
Reverse T3 (rT3), 2, 21 pathophysiology, 102
RhTSH, 276–278 treatment, 104
Index 447

Subclinical hypothyroidism (SCH) follow-up, 274–278


clinical symptoms, 153 gamma knife therapy, 273
definition, 152 imaging studies, 278–279
diagnosis, 71–72 ipsilateral lobectomy, 260
and hypertension, 154 isthmusectomy, 260
LT4 therapy, 166–168 of patients with high serum Tg levels and
neurobehavioral and neuromuscular negative imaging studies, 279–280
symptoms in, 158 during pregnancy, 262
risk factor for myocardial infarctions, 156 radioiodine (131 I) therapy, 262–272
role in development of atherosclerosis, 155 subtotal lobectomy, 260
skeletal muscle abnormalities, 158 surgical complications, 262
treatment, 72 thyroid hormone therapy, 272–273
treatment and cardiovascular outcomes, 157 total or near-total thyroidectomy, 261
Subtotal lobectomy, 260 tumor staging systems and prognostic scoring
Subtotal thyroidectomy, 377, 422 systems, 259–260
Sucralfate, 173 Thyroid dysgenesis, 336, 347
Supraphysiologic doses, of thyroid hormone, Thyroidectomy, 67, 115, 169
126 completion, 261
Surgical excision treatment, 224 during pregnancy, 262
Surveillance, Epidemiology, and End Results prophylactic total, 307
(SEER) program, 238 Thyroid function tests, 50, 55
Thyroid gland
Tachycardia, 41 development, 331–332
Tall cell variant, 248 functions of, 2
Tamoxifen, 10, 16 role of, 1
T4 and fetal hypothalamic-pituitary-thyroid Thyroid gland enlargement (goiter)
system, 333–334 classification and prevalence, 204–205
T3 and T4 toxicosis, 408 history and examination, 206
Technetium (99m Tc) scanning, 76, 211 laboratory and radiologic diagnosis
Thrombocytopenia, 60 FNA biopsy, 213–218
Thymic enlargement, 45 other imaging techniques, 212–213
Thyroglobulin (Tg), assessment, 17–19 scintigraphy, 210–212
Thyroid autoantibodies, 19–21, 105 serum calcitonin (CT) measurement,
Thyroid autoimmunity, during pregnancy, 207–208
363–364 ultrasonography, 208–210
Thyroid cancer, 169 management
during pregnancy, 381–382 multinodular goiter, 223–225
Thyroid carcinoma solitary nodules, 218–223
classification, 239–242 pathogenesis, 205–206
diagnosis, 250–253 Thyroid hormone binding ratio (THBR), 7–8
epidemiology, 238–239 Thyroid hormone receptor-beta gene (TR-␤),
factors influencing prognosis and affecting 125
outcome, 253–259 Thyroid hormone replacement therapy, 25
follicular, 242, 249–250 Thyroid hormone resistance syndromes, 160,
incidence, 239 337
papillary, 242–249 diagnosis, 126
radiation-induced, 242–243 epidemiology, 125
treatment pathophysiology, 125–126
chemotherapy, 273–274 pediatric, 347–348
completion thyroidectomy, 261 treatment, 126–127
external beam radiation therapy, 273 Thyroid hormone therapy, 272–273
448 Index

Thyroid hormonogenesis, inborn error of, 337 Treponema pallidum, 110


Thyroid neoplasia, 27–28 Triiodothyronine-containing thyroid hormone
Thyroid nodules, during pregnancy, 381–382 preparations, 8
Thyroid nodules, pediatrics Triiodothyronine (T3), 2, 121
clinical evaluation, 358 Triostat, 164
laboratory evaluation, 358 Trophoblastic tumors
other screening tools, 358 diagnosis, 129
prognosis, 360 epidemiology, 128
therapy, 359–360 pathophysiology, 129
Thyroid peroxidase (TPO), 2 treatment, 129
Thyroid replacement therapy, 350 TSH-binding inhibitors, 20
Thyroid storm, 73–74, 413 TSH/free T4 relationship, 2
Thyroid transcription factor (TTF), 331–332 TSH levels, implications of, 16–17
Thyrotoxicosis, 16, 41 TSH-producing pituitary tumors (TSHomas)
due to thyroid cancer, 130–131 diagnosis, 122–124
symptoms of, 102–103 epidemiology, 122
Thyrotoxicosis factitia, 121 pathophysiology, 122
diagnosis, 121 treatment, 124–125
epidemiology, 121 TSH receptor gene, 206
pathophysiology, 121 TSH receptor–stimulating antibodies,
treatment, 122 measurement of, 48–49
Thyrotoxic periodic paralysis, 43 TSH receptor–stimulating immunoglobulins, 40
Thyrotropin-binding inhibitory immunoglobulin TSH-secreting pituitary adenomas, 16
assay (TBII assay), 48–49 TSI titers, 52–53
Thyroxine-binding globulin (TBG), 2, 333 T4 suppressive therapy, 221
Thyroxine suppressive therapy, 221 TTF2, 332
Thyroxine (T4), 2 T3 thyrotoxicosis, 26, 55
Tirosint (Institute Biochimique), 164 T3-toxicosis, 8, 47
Tissue responses, to thyroid hormone action, T4 toxicosis, 47
21–22 Tumor calcitonin, 302
T3 measurement methods, 4 Tumor histology, 255
decreased, 11–13 Tumor necrosis factor, 12
free hormone, 4–8, 47 Turner syndrome, 346
increased, 8–11 “Two-step” assay method, 7
T4 measurement methods, 4 Type 1 deiodinase (D1), 333
decreased, 11–13 Tyrol, 112
free hormone, 4–8 Tyrosine kinases (TKs), 258
increased, 8–11
TNM classification, 259–260 Ultrasonography, 253, 321
Total or near-total thyroidectomy, 261 Unithroid (Jerome Stevens), 164
Toxic multinodular goiter US-FNA, 217–218
diagnosis, 81
pathogenesis of, 81 Vitiligo, 42
treatment, 81–83
Transient hypothyroidism, 169, 338–339 Wechlser Intelligence Scale for Children, 367
Transient painless thyroiditis. See Silent or Wolff-Chaikoff effect, 74
“painless” thyroiditis
Transient thyrotoxicosis with lymphocytic X-linked inherited TBG excess, 10
thyroiditis. See Silent or “painless”
thyroiditis Yttrium-90 (99 Y)-labeled humanized anti-CEA
Transthyretin (thyroxine-binding prealbumin), 2 monoclonal antibodies, 306
Endocrinology

Second Edition
Medical Management of Thyroid Disease
Medical
about the book…
For general practitioners and endocrinologists, the new Second Edition of this
bestselling book offers the most up-to-date and practical guidance to diagnose
and manage common and uncommon thyroid diseases.
New to the Second Edition:
• information on thyroid neoplasia, leading to new effective treatments of
advanced thyroid cancer
• important new research on subclinical thyroid disease in the elderly and
Management
of Thyroid
thyroid disorders in pregnancy
• new research on thyroid physiology, pathophysiology, and therapeutics
The new edition is fully evidence-based and updated to include the most current

Disease
treatment and latest findings:
• the screening and case finding for thyroid disease
• the use of calcitonin in the diagnosis of medullary thyroid cancer
• the diagnosis and management of subclinical hyperthyroidism
(mild hyperthyroidism)

Second Edition
• thyroid disease related to interferon therapy and amiodarone therapy
about the editor...
DAVID S. COOPER is Professor of Medicine, The Johns Hopkins University School
of Medicine; Professor of International Health, Johns Hopkins Bloomberg School
of Public Health; and Physician and Director, the Thyroid Clinic, Johns Hopkins
Hospital, Baltimore, Maryland, USA. Dr. Cooper received his M.D. from Tufts
University, Boston, Massachusetts, USA. He is a member of the Endocrine Society
and is a past president of the American Thyroid Association. Dr. Cooper is currently
Deputy Editor of the Journal of Clinical Endocrinology and Metabolism, Editor–in-
Chief of Endocrinology, Up-to-Date, and Contributing Editor of the Journal of the
American Medical Association (JAMA). Dr. Cooper was also the editor of the first
edition of Informa Healthcare’s Medical Management of Thyroid Disease.
Printed in the United States of America

Cooper
H7064

Edited by
David S. Cooper

You might also like