You are on page 1of 335

JOURNAL OF CHROMATOGRAPHY LIBRARY - Volume 36

selective gas chromatographic


detectors
This Page Intentionally Left Blank
JOURNAL OF CHROMATOGRAPHY LIBRARY - volume 36

selective gas
chromatographic
detectors
M. Dressler
Institute of Analytical Chemistry, Czechoslovak Academy of Sciences, Leninova 82,
61142 Bmo, Czechoslovakia

ELSEVIER
Amsterdam - Oxford - New York - Tokyo 1986
ELSEVIER SCIENCE PUBLISHERS B.V.
Sara Burgerhartstraat 25
P.O. Box 211, 1000 AE Amsterdam, The Netherlands

Distributors for the United States and Canada:

ELSEVIER SCIENCE PUBLISHING COMPANY INC.


52, Vanderbilt Avenue
New York, NY 10017, U.S.A.

Library of CO'lgress Cataloging-in-PubIiClltion Data

Dressler, M., 1940-


Selective gas chromatographic detectors.
(Journal of chromatography library ; v. 36)
Includes bibliographies and index.
1. Gas chromatography. I. Title. II. Series.
QD79.c45D74 1986 543'.0896 86-13366
ISBN 0-444-42488-1
ISBN 0-444-42488-1 (Vol. 36)
ISBN 0-444-41616-1 (Series)

© Elsevier Science Publishers B.V., 1986

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system or
transmitted in any form or by any means, electronic, mechanical, photocopying, recording or other-
wise, without the prior written permission of the publisher, Elsevier Science Publishers B.V./Science &
Technology Division, P.O. Box 330, 1000 AH Amsterdam, The Netherlands.

Special regulations for readers in the USA - This publication has been registered with the Copyright
Clearance Center Inc. (CCC), Salem, Massachusetts. Information can be obtained from the CCC
about conditions under which photocopies of parts of this publication may be made in the USA. All
other copyright questions, including photocopying outside of the USA, should be referred to the
publisher.

Printed in The Netherlands


CONTENTS

Journal of Chromatography Library (other volumes in the series) IX


Preface XIII

1. Introduction
References . 3

2. Basic terms relating to detectors 5


2.1. Detector sensitivity 5
2.2. Minimum detectability 6
2.3. Detection limit. . . 7
2.4. Detector noise 8
2.5. Dependence of detector response on amount of compound 10
2.6. Selectivity of response 12
References . . . . . . . . • • . 13

3. Alkali flame-ionization detector 15


3.1. Introduction . . . • . . . 15
3.2. Detector design. • . • . . 16
3.3. Detector life, reproducibility of response 23
3.4. Background current (hydrogen flow-rate) 24
3.5. Negative response. . • . . . . • . . . . . 29
3.6. Response to individual heteroatoms 33
3.7. Influence of compound structure on detector response 41
3.8. Influence of main operational parameters on detector response 43
3.9. Detection mechanism 54
References . . . '. 59

4. Flameless alkali sensitized detectors 63


4.1. Introduction . . . • • • • . 63
4.2. The Perkin-Elmer detector .• 64
4.3. The Hewlett-Packard detector 72
4.4. The Tracor detector . . . . . 73
4.5. The Varian detector . . . • . 74
4.6. The Detector En~ineering Technology detector 78
4.7. The chemi-ionization detector • . • • . . • . 83
4.8. Detector life and reproducibility of response 84
4.9. Detectors for halogen compounds 87
References • . . . • . . . • • • • • • • . . . . . 90
VI

5. Flame-ionization detector . • • . . • • • . • . • . 91
5.1. Introduction . . . . . . • . . . • . . . . . . 91
5.2. Hydrogen atmosphere flame-ionization detector 92
5.3. Hydrogen atmosphere flame-ionization detector for silicon
compounds 102
5.4. Flame-ionization detector with hydrocarbon background 105
5.5. Selective detection of halogen compounds 106
References 106

6. Photoionization detector 109


6.1. Introduction. 109
6.2. Response model 111
6.3. Sensitivity of response and minimum detectability 112
6.4. Selectivity of response 118
6.5. Carrier gas 126
References 131

7. Flame photometric detector 133


7.1. Introduction . . . . . 133
7.2. Response model . . . . 136
7.3. Detector sensitivity and minimum detectability 137
7.4. Selectivity of response . . 144
7.5. Tin and germanium compounds 145
7.6. Halogen compounds • . . . . 147
7.7. Other detection possibilities 149
7.8. Linearity of response 150
7.9. Sulphur background. 152
7.10. Response quenching 152
7.11. Flame stability . . 157
7.12. Other identification possibilities 157
References ••. . . . . . 158

8. Chemiluminescence detectors 161


8.1. Introduction . . . . . 161
8.2. Detector for N-nitroso compounds 161
8.3. Detector for nitroaromatic compounds 169
8.4. Detector for nitrogen-containing compounds 170
8.5. Ozone chemiluminescence detector for compounds not containing
nitrogen . • . . . . . . • • • . . . . . . . . 174
8.5A. Redox chemiluminescence detector . . . . . . 174/312
8.6. Chemiluminescence detector with sodium metal 174
8.7. Fluorine-induced detector 177
References 179
VII

9. Electrolytic conductivity detector 181


9.1. Detector construction. 181
9.2. Selectivity of response 186
9.3. Response 189
9.4. Solvent . . 196
9.5. Gases . . . 201
9.6. Temperature 203
References . . . 206

10. Coulometric detector 209


10.1. Introduction. 209
10.2. Response . . . 210
10.3. Quantitative results 212
References . . • • • . • • 215

11. Electron-capture detector 217


11.1. Introduction . . . 217
11.2. Design . . . . . • 218
11.3. Sources of primary electrons 219
11.4. ~lethods of measuring detector current 224
11.5. Response theory . • . 230
11.6. Response . . . . • . • . • . • . . . 235
11.7. Linearity of response . . . . . . . . 249
11.8. Selective electron-capture sensitization 251
11.9. Coulometric and hypercoulometric response 263
11.10. Use of the electron-capture detector with capillary columns 266
References . . . . . . 269

12. Ion mobility detector 275


12.1. Introduction .. 275
12.2. Principle of the technique 275
12.3. Detection principles 279
12.4. Effect of background 286
References • • . • • • . 288

13. Miscellaneous detectors 291


13.1. Introduction • . • • . . . • 291
13.2. Plasma-emission spectrometry 291
13.3. Atomic-absorption spectrometry 294
13.4. Ion-selecti~e electrodes •.• 294
13.5. Piezoelectric sorption detector 295
13.6. Mass and infrared spectrometry 296
References . . . . • . . • . . . • . . 305
VIII

14. Conclusion 311


References 311

List of abbreviations 313

Subject index . . . . 315


IX

JOURNAL OF CHROMATOGRAPHY LIBRARY


A Series of Books Devoted to Chromatographic and Electrophoretic
Techniques and their Applications
Although complementary to the Journal of Chromatography, each volume in the Library
Series is an important and independent contribution in the field of chromatography and
electrophoresis. The Library contains no material reprinted from the journal itself.

Other volumes in this series

Volume 1 Chromatography of Antibiotics (see also Volume 26)


by G.H. Wagman and M.J. Weinstein
Volume 2 Extraction Chromatography
edited by T. Braun and G. Ghersini
Volume 3 Liquid Column Chromatography. A Survey of Modern Techniques and
Applications
edited by Z. Deyl, K. Macek and J. Janak
Volume 4 DetectoR in Gas Chromatography
by J. ~ev<!ik
Volume 5 Instrumental Liquid Chromatography. A Practical Manual on High-Per-
formance Liquid Chromatographic Methods (see also Volume 27)
by N.A. Parris
Volume 6 Isotachophoresis. Theory, Instrumentation and Applications
by F.M. Everaerts, J.L. Beckers and Th.P.E.M. Verheggen
Volume 7 Chemical Derivatization in Liquid Chromatography
by J.F. Lawrence and R.W. Frei
Volume 8 Chromatography of Steroids
by E. Heftmann
Volume 9 HPTLC - High Performance Thin·Layer Chromatography
edited by A. Zlatkis and R.E. Kaiser
Volume 10 Gas Chromatography of Polymers
by V.G. Berezkin, V.R. Alishoyev and LB. Nemirovskaya
Volume 11 Liquid Chromatography DetectOR
by R.P.W. Scott
Volume 12 Affinity Chromatography
by J. Turkova
Volume 13 Instrumentation for High·Performance Liquid Chromatography
edited by J.F.K. Huber
Volume 14 Radiochromatography. The Chromatography and Electrophoresis of
Radiolabelled Compounds
by T.R. Roberts
Volume 15 Arltibiotics. Isolation, Separation and Purification
edited by M.J. Weinstein and G.H. Wagman
Volume 16 Porous Silica. Its Properties and Use as Support in Column Liquid Chro-
matography
by K.K. Unger
Volume 17 75 Yean of Chromatography - A Historical Dialogue
edited by L.S. Ettre and A. Zlatkis
x

Volume 18A Electrophoresis. A Survey of Techniques and Applications.


Part A: Techniques
edited by Z. Deyl
Volume 18B Electrophoresis. A Survey of Techniques and ApPlications.
Part B: Applications
edited by Z. Deyl
Volume 19 Chemical Derivatization in Gas Chromatography
by J. Drozd
Volume 20 Electron Capture. Theory and Practice in Chromatography
edited by A. Ziatkis and C.F. Poole
Volume 21 Environmental Problem Solving using Gas and Liquid Chromatography
by R.L. Grob and M.A. Kaiser
Volume 22A Chromatography. Fundamentals and Applications of Chtomatographic
and Electrophoretic Methods. Part A: Fundamentals
edited by E. Heftmann
Volume 22B Chromatography. Fundamentals and Applications of Chromatographic
and Electrophoretic Methods. Part B: Applications
edited by E. Heftmann
Volume 23A Chromatography of Alkaloids. Part A: Thin-Layer Chromatography
by A. Baerheim Svendsen and R. Verpoorte
Volume 23B Chromatography of Alkaloids. Part B: Gas-Liquid Chromatography and
High-Performance Liquid Chromatography
by R. Verpoorte and A. Baerheim Svendsen
Volume 24 Chemical Methods in Gas Chromatography
by V.G. Berezkin
Volume 25 Modern Liquid Chromatography of Macromolecules
by B.G. Belenkii and L.Z. Vilenchik
Volume 26 Chromatography of Antibiotics
Second, Completely Revised Edition
by G.H. Wagman and M.J. Weinstein
Volume 27 Instrumental Liquid Chromatography. A Practical Manual on High·Per·
formance Liquid Chromatographic Methods
Second, Completely Revised Edition
by N.A. Parris
Volume 28 Microcolumn High-Performance Liquid Chromatography
by P. Kucera
Volume 29 Quantitative Column Liquid Chromatography. A Survey of Chemometric
Methods
by S.T. Balke
Volume 30 Microcolumn Separations. Columns, Instrumentation and Ancillary
Techniques
edited by M.V. Novotny and D. Ishii
Volume 31 Gradient Elution in Column Liquid Chromatography. Theory and Practice
by P. Jandera and J. Chunicek
Volume 32 The Science of Chromatography. Lectures Presented at the A.J.P. Martin
Honorary Symposium, Urbino, May 27-31, 1985
edited by F. Bruner
XI

Volume 33 Liquid Chromatography Detectors. Second, Completely Revised Edition


by R.P.W. Scott

Volume 34 Polymer Characterization by Liquid Chromatography


by G. Glockner

Volume 35 Optimization of Chromatographic Selectivity. A Guide to Method


Development
by P.J. Schoenmakers

Volume 36 Selective Gas Chromatographic Detectors


by M. Dressler
This Page Intentionally Left Blank
XIII

PREFACE

In the last decade, specialized chromatographic literature and chromatographic


practice have placed emphasis on the identification of individual compounds from
complicated gas chromatograms. Of course, a gas chromatograph-mass spectrometer-
data system has been the most efficient combination. However, much simpler tech-
niques are provided by use of the so-called selective detectors. Selective
detectors give a response only to certain heteroatoms, resulting in a simplified
chromatogram. Many selective systems exist and some of them are manufactured
commercially and employed in routine chromatographic practice.
Recently, new selective detectors have been developed and known detector
designs have been innovated. The aim of this book is to collect and to collate
up-to-date information on this topic to give the reader a detailed understanding
of selective detectors in general, their principles, designs and analytical
possibilities.
Throughout the preparation of the manuscript, I have appreciated the assis-
tance of many people from the Institute of Analytical Chemistry of the Czecho-
slovak Academy of Sciences. Special acknowledgements are due to my colleague
Dr. Josef Novak who read the manuscript and made valuable suggestions, and to
Mrs. Melita Radevova for translation into English.

Brno, Mareh 1986 M. DRESSLER


This Page Intentionally Left Blank
1

Chapter 1

INTRODUCTION

Qualitative analysis by gas chromatography (GC) is based on the concept of


the retention characteristics of sample compounds. The absolute values of the
retention characteristics such as retention times, retention volumes or specific
retention volumes, and also the relative values of these quantities such as rela-
tive retention times or volumes or Kov6ts retention indices are used. The reten-
tion volume, VR' is determined by the quantities that characterize the chromato-
graphic system, i.e., the dead volume of the column, VO' the volume of the sta-
tionary phase, VS' or the surface area of the adsorbent and the partition coef-
ficient, K. The value of K depends on the substance being analysed and the sta-
tionary phase employed:

( 1.1)

As the partition coefficient is a function of the thermodynamic properties


of the system, the retention volume of a given solute in a given chromatographic
system is constant at constant temperature and pressure, but it is neither
selective nor specific. In addition, the separation power of any chromatographic
column, even the best, is limited. This means that the column has a limited ca-
pacity for peaks and, therefore, the separation number (Trennzahl, TZ)l, given
by the number of the separated peaks that can be placed between two successive
peaks of neighbouring n-alkanes:

(1. 2)

(where d2 - d1 is the distance between the peak maxima of the n-alkanes and Y
i
are the peak widths at half-heights) is finite. For complex mixtures (of natural
or biological origin, for instance) not all of the peaks can be separated by the
column even if the partition coefficients of the compounds differ. Consequently,
from a theoretical point of view, only a negative result can be considered as
conclusive in chromatography, i.e., that a compound a is not identical with b
when VRa F VRb .
I~hen performing qualitative analyses on the basis of retention data only, one

should be aware of the risk that the chromatographic peak may not pertain to the
2

substance selected for calibration, although the latter has the same retention
value, and even that it may not be due to a single substance but to two or more
substances that have the same partition coefficients in the given system. Hence
it has been generally accepted that the identification power of GC (and of
chromatography in general) is far less than its excellent separation power.
Therefore, a number of auxiliary techniques have been used for identification
purposes, such as methods utilizing regularities in the partition coefficients
within a homologous series which are known or can be predicted from experimental
data. The retention values can be correlated with values characterizing the
homologous series. Such values are either those that cannot be determined in any
way from the chromatogram (e.g., molecular weight, number of carbon atoms or
boiling point) or those found by chromatographic experiments (e.g., the ratio
of the retention values on two stationary phases differing in polarity, or re-
tention values measured at different temperatures)2,3. By interpolating these
relationships, the retention value for a particular member of a homologous
series can be obtained and compared with the retention value obtained experi-
mentally. The agreement between the two sets of data increases the probability
that the predicted identity wfll agree with that of the substance being analysed.
However, these identification approaches are labourious and time consuming.
Reaction gas chromatography3,4 is another approach used to facilitate the
identification of individual components on a chromatogram. With this approach,
the sample is subjected to selective reactions intended to remove selected types
of substances from the chromatographic spectrum or to convert them into different
substances. Subsequently, the chromatogram of the original sample is compared
with those obtained after reaction. By introducing chemical reactions into the
system, additional information on the identity of the sample compounds is ob-
tained from the chromatogram, and the possibility of confusing the identities of
the substances is again reduced.
The utilization of the detector itself for the identification of substances
is an efficient approach to the application of auxiliary techniques for qualita-
tive purposes. From the viewpoint of quantitative analysis, chromatography re-
quires a detector that responds as far as possible to all types of sample com-
pounds. If it is sensitive enough, the detector provides a record of all the
solutes. thus making possible their subsequent determination. If, in addition,
the detector response per unit solute mass (weight or number of moles) is si-
milar for different types of compound, which is very advantageous for quanti-
tative analysis. the detector itself provides no data for qualitative purposes.
The availability of a detector that gives a response that differs in some way
for a certain type of compound from that for other compounds is obviously ad-
vantageous in qualitative analysis. Therefore, let us consider the ways in which
the response of a certain detector can differ for different types of compounds.
3

It can differ, first of all, in the level of the response per unit solute mass.
The ideal case would be represented by a detector responding to a certain type
of compound only (e.g. to a certain kind of heteroatom in a molecule of these
compounds). As will be seen later, no gas chromatographic detector meets this
requirement. However, selective detectors 5- 8 are available the response of which
per unit mass to compounds containing a certain heteroatom differs 'considerably
from the response to other compounds. In addition to the level of the response,
the detector response to various compounds can also differ in polarity. Spectral
detectors such as the mass or infrared spectrometer supply, in addition to the
common chromatographic record, data for each peak that allow one to characterize
the compound. It is therefore evident that the detector itself can contribute to
the identification of chromatographed compounds, if a suitable selective detector
is properly selected for a particular case and if optimum operating conditions
for the chosen detector are maintained. In order to function properly, selective
detectors must be operated under optimum conditions. In other words, there are
a number of operating variables that can either adversely affect or even nullify
the function of the selective detector.
Descriptions of the individual selective detectors and of the principles
providing the basis for their operation, an,analysis of the effects of the
various operating conditions on the basic parameters of the selective detectors
and consideration of the potential use of these detectors in qualitative analysis
are the subjects of this book.

REFERENCES

1 R.E. Kaiser, Z. Anal. Chem., 189 (1962) 1.


2 J.H. Purnell, Gas ChFomatogFaphy, Wiley, New York, 1962.
3 R.C. Crippen, Identification of OFganic Compounds with the Aid of Gas ChFoma-
togFaphy, McGraw-Hill, New York, 1973.
4 V.G. Berezkin, AnaZytical Reaction Gas ChFomatogFaphy, Plenum Press, New York,
1968.
5 M. Krejc1 and M. Dressler, ChFomatogF. Rev., 13 (1970) 1.
6 M.L. Seluck9, ChFomatogFaphia, 4 (1971) 425.
7 D.F.S. Natusch and T.M. Thorpe, Anal. Chem., 45 (1973) 1185A.
8 L.S. Ettre, J. ChFomatogF. Sci., 16 (1978) 396.
This Page Intentionally Left Blank
5

Chapter 2

BASIC TERMS RELATING TO DETECTORS

CONTENTS

2.1. Detector sensitivity 5


2.2. Minimum detectability 6
2.3. Detection limit . . . 7
2.4. Detector noise . . . . . . . . . . . . . . . . . . . 8
2.5. Dependence of detector response on amount of compound 10
2.5.1. Detector linearity 10
2.5.2. Dynamic detector range 11
2.5.3. Linear dynamic detector range 11
2.6. Selectivity of response 12
References . . . . • . . . • . . . 13

2.1. DETECTOR SENSITIVITY

Detector sensitivity is the basic term used in describing any detector. In


spite of the frequent use of this term, there is great non-uniformity in
expressing detector sensitivity in the literature l - 9 , which has resulted in
misleading interpretations in many instances. Frequently, terms such as molar
response, detection limit and minimum detectable amount are used to express
detector sensitivity. In addition, different terms have been used for the same
thing. For instance, the terms minimum detectable amount, minimum detectable
quantity, minimum detectable limit, detection limit, minimum detectability and
limit of detectability have been used to express the minimum detectable rate
of introduction of solute mass into the detector.
The detector responds to an eluted sample compound, to the carrier gas and
to other compounds that may be present and the response (H) can be expressed
by

R=R.+H (2.1)
'1.- m +Rx

Under stabilized chromatographic conditions, the response to the carrier gas


(H ) and to the impurities (H ) is constant and can be compensated to zero by
m x
the applied counter voltage. Thus, the net response given by the detector during
the passage of an eluted sample compound through the detector is equal to the
response to that compound, Hi'
6

The detector sensitivity (s) to a compound in the effluent from the chro-
matographic column can generally be defined as the change in the net detector
response, i.e., in response to a substance i, with the change in the concentra-
tion (and/or mass flow) of this compound in the fluid supplied:

S=dR./da.
1.- 1.-
(2.2)

or

S = dR./(dm./dt)
1.- 1.-
(2.3)

depending on whether a concentration-sensitive or a mass-rate-sensitive detector


one is used 10 .
Detector sensitivity is generally expressed as the detector response per
unit mass, i.e., as the molar response, R~Ol (response per mol), or the weight
response, R~ (response per gram), of a given compound. With ionization detec-
tors, for example, it is coulombs per mol or coulombs per gram of the compound.
The detector sensitivity itself does not enable one to determine the smallest
amount of the compound that could be detected by means of a given detector. For
these purposes the detector noise must also be considered.

2.2. MINIMUM DETECTABILITY

A peak observed over the detector noise should be large enough relative to
the noise level to allow the reliable determination of the compound concerned.
It follows from Young's statistical studies 11 that a peak that is at least
twice as high as the width of the noise zone almost certainly belongs to a
solute. The larger the multiple of this minimum value, the greater is the
precision of determination. Karger et al. 12, for instance, considered a quintuple
value, but generally a multiple of two is used. Thus, if the peak height in the
chromatogram is at least twice the detector noise band width, Rn , this peak can
safely be considered to be the peak of an eluted compound. This value then
corresponds to the minimum detectable solute concentration in the column
effluent, a~in
1.-
(g/ml), or the minimum detectable solute mass rate, m~in
1.-
(g/sec),
for a concentration-sensitive detector and/or a mass rate-sensitive detector,
respectively:

(2.4)
7

With a selective detector, the detector response is conditioned by the


presence in the solute molecule of the heteroatom to which the detector is
selective, i.e. the detector responds predominantly to d certain type of com-
pound. For this reason, the minimum detectable mass rate should be expressed
in grams of the heteroatom per second rather than in grams of the whole compound
per second (provided that a mass-rate-sensitive detector is used). For instance,
with an alkali flame-ionization detector in the P mode (see Chapter 3), the
minimum detectable mass rate of 6.1 • 10- 13 g/sec for ethion (16.3% P) cor-
responds to 1.0 • 10- 13 g/sec of P. With most selective detectors, the minimum
detectable mass rate, in grams of the heteroatom per second, is approximately
the same for different types of compound containing the same heteroatom.
However, this finding does not hold for detectors that give a substance-
selective response. For instance, the response of an electron-capture detector
to halogen-containing compounds varies over several orders of magnitude (see
Chapter 11), depending on the structure of the solute molecule. Hence the
minimum detectability of a heteroatom also varies over a range of several orders
of magnitude in this instance.
It should be borne in mind that the minimum detectable solute concentration
and/or mass rate refers to the concentration or mass rate of the compound in
the detector and not in the sample injected. This fact is frequently neglected,
which results in discrepancies when comparing different detectors.

2.3. DETECTION LIMIT

The term detection limit, expressed in grams of solute compound, refers to


sample injection into the chromatograph. It denotes the smallest solute mass,
w~in, that still yields a discernible peak on injection into the chromatograph.
This value is always larger than the minimum detectable mass rate because of
dilution of the sample in the chromatographic column during the chromatographic
process. Owing to zone dispersion, the peak height of the solute compound is
smaller at the end of the chromatographic column than that at the inlet. Pro-
vided the solute peak is Gaussian, the concentration of the solute compound
at the peak maximum, a~ax, is related to the mass of the solute injected, ws '
or to the original concentration of the solute in the sample, as, by

(2.5)
8

where N is the number of theoretical plates of the column, VR is the retention


volume of the solute and vs is the volume of sample injected. Similarly, the
mass rate of the solute at the peak maximum, m~ax,
'Z-
is given by

(2.6)

where F is the carrier gas flow-rate as measured in the detector.


Hence, in contrast to the minimum detectability, the detection limit is not
constant but depends on the column conditions, i.e., the length and kind of
the column packing, the column temperature and the retention time of the sample
compound. The relationships between the minimum detectable concentration,
minimum detectable mass rate and detection limit can be written as

= C.'min
Z-
(2.7)

and

VR 1271
= m.'min
Z- • IN F (2.8)

As far as the solute concentration in the original sample is concerned (in


ppm, for instance), the detection limit for the solute concentration, c~in, is
the lowest solute concentration in the material being analysed that gives a
discernible peak on injecting the maximum permissible volume, v~ax, of this
sample material into the chromatograph:

(2.9)

2.4. DETECTOR NOISE

The detector response fluctuates owing to lack of constancy of experimental


parameters such as temperature, gas flow-rates, power line voltage, thermal
stability of the electrometer circuitry and stability of the detector components.
All perturbations of the detector output that are not due to an eluted solute
are referred to as noise. These perturbations can be divided into three types
(Fig. 2.1). The first type is the short-term noise, consisting of baseline
perturbations the frequency of which is substantially higher than that of the
9

Long term noise

Short term noise

Drift:
I
I
I
I
-----------j

Fig. 2.1. Detector perturbations.

eluted peaks. Frequently, this type of noise can be eliminated by means of a


filter. The long-term noise consists of perturbations that have a frequency
similar to that of the eluted peaks. This type of noise can neither be
distinguished from the eluted peaks nor be filtered, because solute peaks having
widths commensurate with the noise period would also be filtered out. Baseline
perturbations with frequencies substantially lower than those of the eluted
peaks are called drift.
The expression of detector noise is also ambiguous. Some authors base it on
the fact that the noise oscillates around the average signal, and they express
the noise value by the baseline fluctuation about this average value, i.e.,
by half the value of the average noise band width. In order to evaluate the
level of the detector noise correctly, the total noise band width should be
considered in any case, however. Sometimes the short-term noise only is supposed
to be the detector noise, but this concept is not correct either. The value of
the detector noise should be represented by the maximum amplitude of the com-
bined short- and leng-term noise measured over a period of about 10 min.
The noise of a certain detector may not be constant, as it usually depends
on the experimental conditions such as the type of stationary phase, column
temperature and detector temperature. For instance, the higher the background
current in the flame-ionization detector, the higher is the noise. With selec-
tive detectors, such as the alkali flame-ionization detector (AFID)13 and the
flame photometric detector (FPD)14, the noise level also increases with
increasing background response.
This means, of course, that the minimum detectability with a given detector
is not constant, but varies as a function of the experimental conditions; with
a given detector type, these variations may also depend on the kind of hetero-
10

atom present in the sample molecule. For instance, the noise of the FPD in-
creases with increasing detector temperature. The response of this detector to
phosphorus compounds also increases with temperature, but the response to sul-
phur compounds decreases. Therefore, for phosphorus compounds, the minimum
detectable mass rate remains about constant, whereas for sulphur compounds it
increases from 1.6 • 10- 12 g/sec of S at a detector temperature of BOoC to
3.B • 10- 12 g/sec of S at 1520 C14 .

2.5. DEPENDENCE OF DETECTOR RESPONSE ON AMOUNT OF COMPOUND

2.5.1. Deteetor linearity

The relationships between the amount of solute (concentration, mass, mass


rate) in the sensor of the detector and the detector response for different
types of detectors, and the linearity of response, have been discussed in detail
by Novcik 15 .
The dependence of the detector response on the amount of solute compound
can generally be expressed by

R. = ken (2.10)
'/..

where k is a constant. A detector for which n = is then a perfectly linear


detector.
This dependence can be expressed graphically in two ways: on a logarithmic
scale or as a semi-logarithmic dependence in which the response to unit amount
of solute (e.g., the molar response) is represented as a function of the amount
of solute (Fig. 2.2). If the detector response is linear, the slope of the
dependence is unity in the former instance, whereas in the latter instance the
response per unit amount of solute is constant. In both instances, tolerances
up to ±5% are allowed.
However, sometimes a response is considered to be linear when n is not unity
but is constant, i.e., when represented graphically on a logarithmic scale

log R.'/.. = log k + n log e (2.11)

a straight line is obtained for the dependence of the response on the amount
of solute. In this book, the term linear response will be applied only for the
case when n = 1 ± 0.05.
11

:!: 5 % Envelope
r~-----.-]l----------- [iL---------
i--------f------~---- -----~
CI I
I
C I
GI
.... I
C I
~
:,- Linear range
~ I
I
I
I
I Upper limit I
: Minimum detectability of linearity :
V- ~

Fig. 2.2. Dependence of the response per gram on amount of compound.

2.5.2. Dynamic detector range

The dynamic detector range is the range of the solute concentrations and/or
mass rates over which the detector yields a concentration-dependent or a mass-
rate-dependent output. The lower limit of this range is given by the amount
of solute for which the response is twice the noise (i.e., the minimum de-
tectability) and the upper limit is given by the pOint from which the detector
no longer responds to increasing amounts of solute.

2.5.3. Linear dynamic detector range

This is the range of solute concentrations and/or mass rates over which the
detector response is linear. With most detectors, the response ceases to be
linear with large amounts of solute, but continues to increase with increasing
amount of solute within a certain range. Hence the linear dynamic range differs
from the dynamic range. The lower limit of the linear dynamic range equals that
of the dynamic range, i.e., that given by the minimum detectable concentration
or mass rate. The upper limit of the linear range is represented by the point
at which the plot crosses the -5% envelope (see Fig. 2.2).
It is the dependence of the response on the amount of substance injected
that is usually used in plots illustrating the linearity of the detector re-
sponse. This approach is not correct in principle, as the figure expressed in
this way does not relate to the amount of the compound in the detector (see
12

section 2.3) and, consequently, it indicates nothing about the operation of


the detector. In any event, the true concentration and/or mass rate of the
solute in the detector should be taken into account. These data can be calculated
from the known amount of the substance injected, ws' by means of the following
equations:

eomax (2.12)
'!..

and

(2.13)

where s is the recorder chart speed (cm/min) and Y is the peak width (cm) at
0.607 of the peak height.
The linear range can be expressed numerically as the ratio of the amount of
solute corresponding to the upper limit of the linear range to the minimum
detectable mass rate or concentration.

2.6. SELECTIVITY OF RESPONSE

A detector is considered to be selective if its response to a certain type


of compound differs markedly from that to another type of compound. By the
term "markedly" a factor higher than 10 is usually meant 4 ,16. Thus, the se-
lectivity 8 is expressed by the ratio of the responses per unit amount of the
compound under consideration and a standard compound. These responses are
expressed as the relative molar response, RMRir :

R~ol
8 = RMRo =_'!..- (2.14 )
H goo 1
r

or as the relative weight response:

R~
s = RWR 0 = -1:... (2.15)
H RW
r

Hydrocarbons, i.e., compounds without heteroatoms, are commonly used as reference


compounds. When comparing chromatograms obtained with a selective and a non-
selective detector, in the former chromatogram we can observe substantially
reduced and/or entirely lacking peaks of the compounds (compared with the latter
13

chromatogram) to which the selective detector is not sensitive and, conversely,


in most instances an increased response to compounds to which the selective
detector is sensitive.
Occasionally, the term specific detector is used, but we should distinguish
between the terms selective detector and specific detector 17 . Specific detec-
tion is the identification of a single substance and/or an element. In fact,
the responses of GC detectors giving a common chromatographic record cannot be
classified as specific. It is not true that a selective detector responds only
to a certain heteroatom, e.g., the FPO to phosphorus or sulphur or the AFIO to
phosphorus or nitrogen, and, consequently, that the detector is specific for
this element. Although the selectivity of some detectors is high, e.g., 10 7 :1
for the AFIO, a response to hydrocarbons is also always obtained. Thus, it is
only the concentrations of the components in the original sample that will
decide the nature of the chromatographic record. Provided that the concentra-
tions of the heteroatom-containing compounds to which the detector is sensitive
and those of the hydrocarbons are commensurate, the ratios of the peak areas
of the former to those of the latter will equal the ratios of the corresponding
response selectivities. However, if the concentrations were inversely propor-
tional to the selectivity of response, the peak areas would be commensurate.
The only exception may be the specificity of response in the sense of a negative
response (e.g., with the AFIO). In this instance, the response of some detectors
to compounds that contain a certain heteroatom show opposite polarity to the
response to other compounds.
Additional, non-chromatographic data can be considered as specific, e.g.,
spectral data obtained with a mass or infrared spectrometer. In this instance,
the spectrum is characteristic of the compound and can be used to identify the
latter.
Certain selective detectors, e.g., the electron-capture detector, give a
response that vari es over a range of several orders of magnitude, dependi ng
on the structure of the compound, with the same heteroatom in the molecule. In
this instance we can speak of substance-s~lective detection.

REFERENCES

1 S. Oal Nogare and R.S. Juvet, Jr., Gas-Liquid Chromatogl'aphy, Wiley, New York,
1962, p. 183.
2 O. Jentzsch and E. Otte, Detektol'en in del' Gas-Chl'omatogl'aphie, Akademische
Verlagsgesellschaft, Frankfurt am Main, 1970, p. 18.
3 O.J. David, Gas Chromatogl'aphic Detectol's, Wiley, New York, 1974.
4 J. 5ev~fk, Detectol's in Gas Chromatography (Journal of Chromatogl'aphy Libl'al'Y,
Vol. 4), Elsevier, Amsterdam, 1976, p. 31.
14

5 R.P.W. Scott, Liquid Chromatography Detectors (Journal of Chromatography


Library, Vol. 11), Elsevier, Amsterdam, 1977, p. 5.
6 L.S. Ettre, J. Chromatogr. Sci., 16 (1978) 396.
7 American National Standard, ANS1/ASTM E 685-79, American Society for Testing
and Materials, Philadelphia, PA, 1979.
8 L.S. Ettre, J. Chromatogr., 165 (1979) 235.
9 J. Novak, Kvantitativnt Analyza Kolonovou Chromatografit, Pokroky Chemie
(Quantitative Analysis by Column Chromatography, Advances in Chemistry),
Academia, Prague, 1981.
10 I. Halasz, Anal. Chem., 36 (1964) 1428.
11 I.G. Young, in H.J. Noebels, R.F. Walland and N. Brenner (Editors), Gas
Chromatography, Academic Press, New York, 1961, p. 75.
12 B.L. Karger, M. Martin and G. Guiochon, Anal. Chern., 46 (1974) 1640.
13 M. Dressler and J. Janak, Collect. Czech. Chem. Commun., 33 (1968) 3970.
14 M. Dressler, J. Chromatogr., 262 (1983) 77.
15 J. Novak, Quantitative Analysis by Gas Chromatography, Marcel Dekker, New
York, 1975, pp. 25-45.
16 M. Krej~f and M. Dressler, Chromatogr. Rev., 13 (1970) 1.
17 H. Egan, ~oc. Soc. Anal. Chem., 9 (1972) 283.
15

Chaptel' :3

ALKALI FLAME-IONIZATION DETECTOR

CONTENTS

3.1. Introduction. . . . . . . . . . . . . . • 15
3.2. Detector design . . . . . . . . . . . . . 16
3.3. Detector life, reproducibility of response 23
3.4. Background current (hydrogen flow-rate) 24
3.5. Negative response . . . . . . . . 29
3.6. Response to individual heteroatoms 33
3.6.1. Phosphorus compounds 33
3.6.2. Nitrogen compounds 34
3.6.3. Halogen compounds 36
3.6.4. Hydrocarbons 37
3.6.5. Sulphur compounds. 38
3.6.6. Arsenic compounds. 38
3.6.7. Boron compounds. . . 39
3.6.8. Tin and lead compounds 39
3.6.9. Silicon compounds. . • . . . 40
3.7. Influence of compound structure on detector response 41
3.8. Influence of main operational parameters on detector response 43
3.8.1. Voltage and polarity of the electrodes 43
3.8.2. Height and shape of the collector electrode 45
3.8.3. Diameter of the tip bore . . . . . . • . . 48
3.8.4. Cations and anions of the alkali metal salt 49
3.8.5. Detector temperature 52
3.8.6. Carrier gas, air . . . 53
3.9. Detection mechanism . . . . . 54
3.9.1. Solid-phase reactions. 54
3.9.2. Gaseous-phase reactions 56
3.9.3. Photoeffects 59
3.9.4. Negative response 59
References . . . . . . 59

3.1. INTRODUCTION

The alkali flame-ionization detector (AFID) is a modification of the flame-


ionization detector (FID). An alkali metal (Na, K, Rb, Cs) salt is introduced
into the detection system and heated to volatilize it. Thus, an alkali metal
enters the flame, where it is ionized. If a compound containing a certain
heteroatcm (P, N, halogens, S, As, B, Si, Pb, Sn) is introduced into the system,
variations in the background current of the detector occur. The response ob-
tained in this way differs in magnitude and, with certain heteroatoms, in
polarity from that of the FlO. The detector response to compounds that do not
16

contain the particular heteroatom, and/or that contain a heteroatom under


operating conditions that are not optimum for that heteroatom, is lO\ler than
that of the FID. Hence the selectivity of the AFID is given both by the level
and the polarity of response.
The response of the AFID depends strongly on all the operating conditions.
In addition, these dependences differ significantly for various AFID designs.
For this reason, data obtained with the individual types of AFID differ con-
siderably in both magnitude and quality. It is difficult or even impossible to
assess the common features of these interrelations for the different designs.
Therefore, a generalization will be made in the following sections wherever
possible. The conditions under which the results were obtained are specified
for the design variants where the interrelations can be affected by additional
parameters.
The flame in the AFID serves a dual purpose: (1) it volatilizes the alkali
metal salt and (2) it produces the ions that are collected. The amount of
hydrogen determines the flame temperature and, therefore, the extent of the
above processes. Therefore, slight variations in the hydrogen flow-rate strongly
affect the detector response. Difficulties with maintaining very accurately a
stable hydrogen flow-rate and some problems with the service life of the source
of the alkali metal salt have resulted in recent years in a reduced interest in
the AFID in favour of the so-called flameless alkali sensitised detectors (see
Chapter 4) that were developed later. This trend is not general, however,
evidence for which is provided by the fact that some manufacturers continue to
equip their gas chromatographs with AFIDs (e.g., pye Unicam and Carlo Erba).
In addition, the low cost and the ease of converting an FID into an AFlD make
this detector still attractive.
There is no unanimity in the designation of the alkali flame-ionization
detector, as it is often also called the thermionic detector (TID) or nitrogen-
phosphorus detector (NPD). However, the term alkali flame-ionization best
describes the character of this type of detector.

3.2. DETECTOR DESIGN

The individual AFID designs mainly differ in the means of inserting the
alkali metal salt into the detection system and in the method of signal sensing.
A schematic representation of the original AFID deSign 1,2 is shown in Fig.
3.1a. Sodium sulphate is moulded on one of the electrodes and heated by the
flame. This simple design was modified in minor detail by moulding the salt to
a carrier (platinum wire, coil, etc.)3-5. Compared with the FID, the sensitivity
of the detector to halogen compounds is increased by a factor of about 10 and
17

a b

Ion -Probe!-)
collector!+)
-i- -Ignitor

Sodium
sulpha te
Ceramic ------.
bead
Quartz
flame tip t Carrier gas
_ + hydrogen

c d

e f

E
E
t'l

Alr--==*~::JJ~~~='Column
L H2

Fig. 3.1. Schematic diagrams of some AFIOs. a, From ref. 21; b, from ref. 6;
c, from ref. 7; d, from ref. 17; e, from ref. 13; f, from ref. 21. Reproduced
with permission.

to phosphorus compounds by a factor of about 300. The main disadvantages of


this AFID design are a reduced useful life and stability (see section 3.3).
This poor life and stability are the fundamental drawbacks of the AFIO that are
responsible for the difficulties in the quantitative interpretation of the
chromatograms. The elimination of these disadvantages has been the principal
objective of all other designs of the AFID.
TABLE 3.1
-
ex>

PRINCIPAL PARAMETERS OF DIFFERENT TYPES OF AFID

Compound Refe- Atom Salt Type of Sensitivity Back- Noise Minimum Recommended flow-rate
rence detec- used AFID (C/mol) ground (pA) detectable (ml/min)
ted (Fig.3.1) current mass rate
(nA) (g/sec) H2 Air Carrier
gas
Methyl pa ra th i on 8 P CsBr c 3 2 3.10- 13 16 170
Hydrocarbon 8 C 1.10-8
Diisopropyl 12 P Na S04 c 220 6 5 8.3.10- 12 70 660 60
methanephosphonate S
K2 04 350 6 5.8 40 660 60
Pyridine 12 N K2S03 3.7 2.5.10- 10
Triethoxymethyl 12 Si Na2S 4 1.9
silane K2S03 6.7
Bromobenzene 12 Br Na2S 4 1.9
Chlorobenzene 12 Cl Na2S04 1.1
Iodobenzene 12 I Na2S04 0.64
K2 S04 1.6
Thiophene 12 S· K2 S04 -6.0 1.10- 10
Tetraethyllead 12 Pb Na2S04 -3.7
K2 S04 -4.2
Tetraethyl tin 12 Sn Na2S04 -8.9
K2 S04 -44
Phosphorus-con- 11 P Rb2S04 Three 0.03 0.02 1.5-10- 14 38 210 19
taining compounds elec-
(1 P atom per trode
molecule)
Azobenzene 11 N 1.0.10- 12
Eicosane 11 C 1.0.10- 7
Methyl parathi on 13 P CsBr e 0.01 1.0.10- 13 21 200 28
Methyl parathi on 14 P CsBr In ga- 8.8* 2.0.10- 13 25 250 20-40
seous
phase
Propasin 14 N 0.33*
Hexane 14 C 5.4.10- 5*
Di isopropyl 15 P Na2S04 f 130 4.5 75 2.1.10- 10 34 1000 33
methanephosphonate
Tetraethyl- 15 P 160 2.6.10- 10 lower flame (FlO)
pyrophospha te
Triethylarsine 15 As 0.23 7.0.10-8 81 102
Arsine 15 As 0.35 2.2.10- 8 upper flame (AFID)
Hydrogen phosphide 15 P 60 9.0.10- 11
p-Dichlorbenzene 15 Cl 2.9
Tetrapropyl ti n 15 Sn -7.0
Benzene 15 C 4.3.10- 2 7.7.10- 7

*
C/g of X.
20

Fig. 3.1b shows the design of Coahrane6 , who placed a ceramic cup filled
with sodium bromide crystals on the detector jet. The hydrogen-carrier gas
mixture flows through the salt bed to be burned at the surface of the latter.
In the next step, the salt was compacted to fit the shape of the tip of the
detector jet (Fig. 3.1c). Here, the gases flow through a channel in the centre
of the tip and again they burn on the surface of the salt 7- 9 . The useful life
of the types of AFIO that utilize a compacted salt as the source of alkali metals
1
substantially exceeds that of the original design by Giuffrida and Karmen ,2; for
this reason, some commercial detectors employ this type of alkali metal source.
Many studies concerned with the AFrO were made with this particular type of
detector (for the principal parameters, see Table 3.1). This design also differs
10
in detail, e.g., the tip may be moulded on to the detector jet •
The sensitivity of the AFIO to compounds that contain a heteroatom is higher
than that of the FlO (1-3 orders of magnitude, depending on the heteroatom),
12 14
but the noise of the AFIO is also higher than that of the FlO (10- _10- A);
thus, the minimum detectable mass rate is comparable to that of the FlO or 1-2
orders of magnitude lower.

TABLE 3.2

COMPARISON OF FlO ANO AFIO SENSITIVITIES

Responses are given in C/mol (signal measured as peak area) and in A/ng (signal
measured as peak height and without any correction for the different retention
times of the test compounds); the latter are given in parentheses. (From ref. 18.)

Parameter AFIO* FIO


Phosphorus and Sulphur
nitrogen
Air flow-rate (ml/min) 550 450 400
Nitrogen flow-rate (ml/min) 68.0 68.0 25
Hydrogen flow-rate (ml/min) 34.0 34.5 25
Probe gap O' o
Background current (A) 1.2.10- 9 9.0.10- 10 8.0.10- 12
Background noise level (A) 1.0,10-12 2.0,10- 13 2.0,10- 14
Triethyl phosphite response 3150 400 0.8
(3.1,10- 9) (4.0'10- 10 ) ( 1.0,10- 12 )
Oibutyl sulphide response 3.0 15.0 1. 50
(2.5,10- 12 ) (1.2.10- 11 ) ( 1.3,10- 12 )
o-Toluidine response 13.5 2.2 1.0
(9.2,10- 12 ) (1.5,10- 12 ) (7.5,10- 13 )
Tetradecane response 0.5 -0.2 2.4
(7.5,10- 14 ) (-3,10- 14 ) (7.0-10- 13 )
* Oetector set to maximum sensitivity, but this does not necessarily correspond
to a setting recommended for optimum detector performance.
21

Attempts to reduce the high AFID noise led to the design of three-electrode
detectors in which the signal of the compound is monitored by means of a circuit
separated from that of the noise. In the Pye three-electrode detector (shown
schematically in Fig. 3.1d), the compacted salt (CsBr, RbCl) is placed in a
metallic cylinder to which a negative voltage is applied. The ions produced by
the burning hydrogen flame and by the ionization of organic compounds that-do
not contain a heteroatom occur only a few millimetres above the flame; these
ions are collected in circuit 2. The ions of the compounds that contain a
heteroatom can be monitored by the collector electrode up to a distance of 30 mm
above the flame (circuit 1). In this manner, the detector noise is separated
from the signal (only a minor part of the background ionization current is mea-
sured by circuit 1)17. Table 3.2 lists the principal parameters of this detector
under operating conditions that result in the maximum response for a given
heteroatom. Data obtained with an FlO are also given for comparison. For routine
work, Hoodless et al. 18 often recommend operating conditions that make allowances
for useful life, noise, selectivity and critical positions of the electrodes,
and in many instances these conditions differ from those given in Table 3.2.
A similar detector was described by Brazhnikov and Shmidel 13 (Fig. 3.1e).
In contrast to the above-mentioned design, the compacted-salt cylinder is
placed in the upper collector electrode while the central electrode with the
applied voltage is made in the shape of a ring that is covered with a platinum

Height of C peak HeightofN


peak

Fig. 3.2. Effect of position of salt crystal on the response to carbon and
nitrogen compounds. 1,2 and 3 are positions of the alkali source with respect
to the flame. (From r~f. 19.)
22

screen. The detector jet is earthed. The principal parameters are listed in
Table 3.1.
The Hewlett-Packard N-FID type B15161 detector 19 was also designed as a
three-electrode detector. The upper collector electrode, containing a compacted-
salt cylinder with a central bore, is movable. At the top of the jet (+350 V)
there is a gate electrode to which the corresponding negative voltage (0 to -350
V) is applied so as to compensate the background ionization current induced by
the rubidium ions. In this manner, the noise is reduced by a factor of 10. The
response to nitrogen compounds between positions 1 and 3 of the collector elec-
trode is approximately constant, whereas there is a decrease in the response
to compounds that do not contain a heteroatom to which the detector is sensitive
(Fig. 3.2). Thus, the detector displays the highest selectivity in position 3
with the following ratios: N:C, Cl, Br = 5000:1, N:I = 200:1 and N:P = 1:10.
The detector noise is the lowest in position 1, and the minimum detectabi1ity
of the detector is also the lowest in this position 19 .
Hartmann 11 described an AFID (Varian) in which the alkali metal salt is
pressed into an earthed stainless-steel cup that forms part of the flame sup-
port. A polarizing voltage (-300 V) is applied to the ignitor coil placed below
the upper surface of the salt source.

o
Air
NP
NP MODE
r-ITl
I
:
I
I
I
I
I
I
L_
~
I

I
I
®

t: * '
0-i

Air
P MODE

'1u
:
I
p:-orl:
I
,0
I

I
t- erw- ~.
H2
C
N

Air
N
N MODE
:-EW-0
I
I

L_
I

__J 1-\
H2
Column Column Column

Fig. 3.3. Operational modes of the NPD-40 (Carlo Erba) • (From ref. 20. )

The NPD-40 detettor (Carlo Erba) employs20 an AFID (Fig. 3.3) in which three
different configurations inside the detector allow its operation in the NP. P
and N modes. The NP mode is used for detection of either nitrogen- or phospho-
rus-containing compounds. The P and N modes show an enhanced response to
phosphorus and nitrogen, respectively, in comparison with other heteroatoms
(see a1so Fig. 3.21).
23

All the above types of design relate to the so-called single-flame AFID in
which the detector operates as an AFID only. Double-flame detectors 21 consist
of two detection systems placed one above the other (Fig. 3.1f). The lower
system functions as a FlO and the upper system as an AFID. An alkali metal salt,
mainly heated by the lower flame (or electrically22), is positioned on a carrier
between the two systems. Two simultaneous chromatograms, an FlO and an AFID
chromatogram, are obtained from the double-flame detector. The response to
phosphorus, halogen and tin compounds is similar to that obtained with the
single-flame AFID (cf., Table 3.1); the response to hydrocarbons, alcohols and
sulphur compounds is approximately two orders of magnitude lower than that of
the FlO. The response of the upper system to nitro compounds is low 23 •
In all of the AFID designs described, the alkali metal salt is located
within the detector space proper. However, the atoms of the alkali metal can
also be brought into the detector compartment in the gaseous phase. The alkali
metal salt is heated (to above 500 oC) in the temperature-controlled compartment
and transported into the detector proper with the flow of inert gas. A stable
flow of alkali metal results in this way24. The advantage of this type of
detector is a low dependence of the response on the variations in gas flow-rate
and the drawback is its bulky construction, because the whole space from the
salt source to the flame has to be thermostated.

3.3. DETECTOR LIFE, REPRODUCIBILITY OF RESPONSE

As the alkali metal salt is heated in the AFID it is volatilized, which


leads to a loss of the salt in the source. By the service life of the AFID is
meant the time span during which the detector behaves as an AFID, i.e., the
time span during which enough alkali metal can be introduced into the detection
system to make the detector respond to compounds that contain heteroatoms.
Hence the detector life is determined by the exhaustion of the alkali metal
source. The reproducibility of the response is qualified by the constancy of
the detector response during the service life of the detector.
With the first AFID designs that had the alkali metal salt moulded on a
carrier (Fig. 3.1a), the source life was usually only a few days and the
reproducibility of the response became poor soon after the operation had
started 2 ,25,26. According to Coahrane 6 (Fig. 3.1b), the detector life was a few
weeks and the decrease in response with time was slower than that with the
preceding type. Mounting the compacted salt on the jet in the form of a tip
(Fig. 3.1c) extends the service life to 1000 h9 ,20 and the response is repro-
ducible over approximately 8 h9.
24

The background ionization current gradually decreases during detector opera-


tion, which results in a decrease in sensitivity27,28. If the background
ionization current is kept constant by increasing the hydrogen flow-rate, a
. d even after 18 h 0 f operatlon
reproducible response can be 0 btalne . 27 . G'
rln d'lng
of the salt surface also results in the background ionization current being
restored to its original value and, thereby, to the original response. With a
compacted salt source placed into the flame above the jet (Fig. 3.1d), the
compensation of the decrease in response to the original value could be achieved
by increasing the hydrogen flow-rate even after operation for 2 months 17 .
The decrease in response with time is often caused by deposits on the detec-
tor electrodes 29 ,30 (cleaning restores the response to its original value), by
deposits of silica from the polysilicone stationary phases or silylating agents 31
on the alkali metal source or by the deposition of products generated by the
combustion of lead and tin compounds on the surface of the salt source 12 .
The life of an AFID in which the alkali metal is introduced into the flame
by the flow of an inert gas is several thousand hours 14 • No changes in the
structure of the salt source occur due to the contact of the flame with the
source, and the reproducibility of the response should be very high.
The stability of the flow-rates of the gases, mainly that of hydrogen, is of
great importance with regard to the reproducibility of the response. Temperature
variations in the flame and, as a result, changes in the concentration of the
alkali metal in the flame occur with slight variations in the flow-rate of
hydrogen (cf., section 3.4). Precise regulation, mainly that of the hydrogen
flow-rate, is essential for this reason (at least to 0.1 ml/min).

3.4. BACKGROUND CURRENT (HYDROGEN FLOW-RATE)

As the function of the AFID is conditioned by the presence of an alkali


metal in the flame, the temperature of the flame and that of the source of the
alkali metal salt housed in the detector are the most important factors affecting
the detector response. The temperature of the salt determines the amount of
alkali metal emitted from the source into the flame, and the temperature of the
flame determines the degree of ionization of the alkali metal. The higher the
hydrogen flow-rate, the greater is the heat released by the combustion of
hydrogen. With increasing hydrogen flow-rate, the flame temperature increases
in all zones of the flame (cf., temperature distribution in the flame and in
the source, Fig. 3.4) and, as a result, the temperature of the alkali metal salt
source also increases. For instance, the temperature at the surface of a jet
tip made of CsBr increases from 400 0 C at a hydrogen flow-rate of 14.1 ml/min
to 750 0 C at 63.8 ml/min, causing the CsBr saturation vapour pressure to increase
25

(mm)
13

11
41
E

-
'0
c
9

7 t =Temperature
~
CI of salt surface
41 5
I

41
U
c
't: SOO 900 1000 ( °C I
" A
0- UI

UI
1-
-----MlO"'C'...--.....!.l..s. Temperatures of flame
and salt tip
E 2

-a.e
.J::. )
t.c.)

t.c.1.
41
0 (mm)

Fig. 3.4. Temperature distribution in flame and in the salt tip depending on
the hydrogen flow-rate. Carrier gas flow-rate, 30 ml/min; air flow-rate, 250
ml/min. Hydrogen flow-rate: 1 = 63.8; 2 = 46.8; 3 = 33.9; 4 = 14.1 ml/min.
A. CsBr melting point. t.c.1-t.c.4 are positions of four thermocouples pressed
into the salt tip. (From ref. 32.)

from 5'10- 5 to 1 mmHg 32 Hence the amount of the salt evaporated from the source
and. at the same time, the number of ionized atoms of the alkali metal increase
with increasing hydrogen flow-rate and, as a result, the background ionization
current also increases. Therefore, the hydrogen flow-rate is considered to be
the principal parameter determining the AFID response, and all response inter-
relations for the individual compounds are often related to the hydrogen flow-
rat/ ,33.
As mentioned in the section on the detector life and reproducibility of
response, the decrease in response with the operating time of the detector can
be compensated for by increasing the hydrogen flow-rate so as to maintain a
constant background ionization current 27 . In this instance, the molar response
is independent of the hydrogen flow-rate within a certain range. With the three-
electrode detector the response remains constant 13 even if the hydrogen flow-
rate is changed from 20 to 28 ml/min.
The temperature differs at various axial distances from the base of the flame
(see Fig. 3.4); the distance of the collector electrode or its dimensions should
affect the response (see section 3.8). It has been found for an AFID fitted with
a jet tip of compacted alkali metal salt that the background ionization current
also changes with variation in the distance of the electrode from the source
TABLE 3.3

DEPENDENCE OF AFID RESPONSE ON THE DISTANCE AND SHAPE OF THE COLLECTOR ELECTRODE, Na 2S0 4

From ref. 34.

Electrode shape Distance from H~ flow-rate Background Relative response


the jet ti p (mm) ( l/min) (nA) Cl P
Ring, diameter 3 mm 0.3 81.5 5.8 0.80
1 72.0 5.8 1.01
3 65.0 5.8 0.97
3 72.0 8.4 1.24
6 66.5 5.8 1.02
6 72.0 9.3 1.30
10 67.5 5.8 1.00
Ring, diameter 8 mm 1 68.5 5.8 1.03
3 68.0 5.8 0.97 0.97
3 72.0 7.1 1.15
6 67.0 5.8 0.98 1.01
10 67.5 5.8 0.99
10 72.0 6.9 1. 11
13 69.0 5.8 1.01
Cylinder, diameter 15 mm 65.0 5.8 1.00 1.00
27

C/moi A
2

C/mol B C/mol c

-1rP

A 10-10 A

Fig. 3.5. Dependence of the response on the background current for halogen
compounds. A, Sodium salt; B, potassium salt; C, caesium salt. 1 = Chlorobenzene;
2 = bromobenzene; 3 = iodobenzene; xl. x2 and x3 = FlO responses at optimum
hydrogen and nitrogen flow-rates. (From ref. 37.)
28

and with the diameter of the ring-shaped electrode. If this current is kept
constant by varying the hydrogen flow-rate, the response is constant again for
phosphorus, halogen and sulphur compounds (Table 3.3)34.
It seems to be obvious from the above results that, in some instances, it
is the background ionization current (as a measure of the concentration of the
alkali metal and the temperature in a given location) rather than the hydrogen
flow-rate that constitutes the principal factor determining the AFID response.
For this reason, the individual relationships between the response and the
operating conditions are often reported with reference to the background
ionization current. If, for a given compound. the AFIO response displays the
same polarity over the whole range of the investigated background ionization
currents, the response follows the variations in the background ionization

CARRIER NITROGEN
101 "",
~.
b
1rP'>~ Bromine
'\
\, .>", f'~~ine

T''.
rJ '"
'\,
'.~
I .11,
\ , ''0 %, '\
II! l'o I d' ~\\.\ \ '~
5a. \\orne I 0

III
~
Q)
'0, v ;'i\
p c
~ ,
~ . ,,
:g r:I' cf'
CI
Q)
C '102
Hycjrogen flow 30ml/min 33ml/min 35ml/min 38 ml/min
1~ 25ml/min

-
Q)
,~ CI.,
'iii
o 101 "'"\
a. q.~ Bromine j1,-o."lodine
loDo.~ \'\ , a..,
"
I
, .
0'
\,Iodine
'.~.
'"'" , , ,
~'.
\ C\,.~
II!
c
o
a.
III
Q)
~

CARRIER HELIUM

Fig. 3.6. Halogen response profiles for varying carrier gas flow-rates (nitrogen
and helium) and five selected hydrogen flow-rates. Electrode height, 2 mm. Bead
bore, '.0 mm. Duplicate injections of 1 ~l of 0.01% chlorobenzene. bromobenzene
and iodobenzene. (From ref. 39.)
29

current 1,27,35. For a compound whose response changes its polarity, the response
level and polarity are functions of the background ionization current and these
functions are characteristic of the given heteroatom and the alkali metal ap-
plied 12 ,34,36,37

3.5. NEGATIVE RESPONSE

The AFID yields a response that is selective also with regard to the polarity
of the response. Under certain operating conditions of the detector (flow-rates
of the gases, alkali metal employed, detector design), halogen 8 ,9,36-41, sul-
phur 12 ,36,42-45, tin 12 , lead 12 and nitrogen 27 ,39 compounds, and hydrocarbons 8 ,9,
17,19,20,36,44 give negative responses, i.e., the background current is de-

tJ2 CARRIER NITROGEN ,......... ;,; ......


-
"~
"iii
o
c.
10"''''
,-........
\~
.. ..
" ,,11'
nitrogen \,
yo
'" I

:
t
T oi,
,

~
i
1cfrrr~T"TTTTT'~'-r\n~" ~
0 \

: e \
8~

jcg~oos.

~....Hydrogen flow 25ml/min l)ml/min~ 33~l/min


--......,
,- ......
.t
~

'....
\9.. \
\\
o i .
I
I

.
Carrier gas ml/\mln
I
I

o I


I

I
I

CARRIER HELIUM~'i/
1cY r
Fig. 3.7. Nitrogen and carbon response profiles for varying carrier gas flow-
rates (nitrogen and helium) and three selected hydrogen flows-rates. Electrode
height, 2 mm. Bead bore, 1.0 mm. Single injections of 1 ~l of 1% aniline (~---~)
and 1% each of p-xylene (6), n-decane (a), p-cymene (0), l-octanol (~) and
anisole (0). (From ref. 39.)
30

creased. At low background currents (low hydrogen flow-rates), all the mentioned
compounds yield a positive response that gradually changes to negative on
increasing the background ionization current (increasing hydrogen flow-rate or
decreasing carrier gas flow_rate B,9,12,34,36,37,39,43; Figs. 3.5-3.7). As the
response changes from a positive to a negative value, changes in the peak shape
occur, as can be seen from the example illustrated by thiophene (Fig. 3.B).
The background ionization current (hydrogen flow-rate) at which the response
of an AFID with a jet tip of compacted salt is negative is different for each
heteroatomB,9,12,34,36,37,39,43,49 and usually increases in the order Sn<Pb<S<
Cl<C<Br<I. This means that, at a certain hydrogen flow-rate, a negative response
is obtained only with certain compounds in the mixture being analysed (e.g.,
sulphur compounds, Fig. 3.9). Then, by changing the hydrogen flow-rate, we can

A
1.8.10 -loA
All
2.2.10 -loA 3.10 -lOA

-9
1.2.10 A

)( -8
1.3.10 A

Fig. 3.B. Dependence of the response to thiophene on the background current.


x indicates that the scale range is reduced by a factor of 5. (Reproduced from
ref. 42 with permission.)
31

change stepwise the polarity of the response to compounds containing other


heteroatoms.
The type of alkali metal used affects the negative response in such a way
that the lower the atomic number of the metal, the higher is the background
ionization current that induces a negative response 12 ,36,37. An example of this
relationship for sulphur compounds is shown in Fig. 3.10. Halogen compounds
(cf., Fig. 3.5) yield a positive response when using a sodium salt; when a
potassium salt is employed, the AFID response is positive only with iodine
compounds36 ,37
The diameter of the jet tip bore also affects the conditions under which a
negative response occurs. With halogen compounds, an increased diameter shifts
the occurrence of a negative response to the region of smaller hydrogen flow-
rates 37 (Fig. 3.11).
At a constant hydrogen flow-rate, a decrease in the carrier gas flow-rate
leads to a negative response (cf., Figs. 3.6 and 3.7). When a constant back-
ground ionization current is maintained while increasing the flow-rate of the
carrier gas (by increasing the hydrogen flow-rate), a negative response occurs
27
at higher flow-rates of the carrier gas with lower ionization currents .

C Cl

5
5 10 min

Fig. 3.9. Chromatogram of a model mixture. S thiophene; C toluene; Cl


chlorobenzene. (Reproduced from ref. 42 with permission.)
32

Negative responses can also be induced by decreasing the potential of the


gate electrode in three-electrode detectors (see section 3.B.1, Fig. 3.2)46 or
by modifying the electrical configuration inside the detector 20 (P mode in
Figs. 3.3 and 3.21; see also section 3.B.1).
Negative AFIO responses have been described only for detector designs in
which the flame either burns on the surface of the jet tipB,9,12,27,34,36-40,
42-45 or where the salt cylinder is situated above the jet, but very deeply
in the flame 19 ,20,41. Only tin and lead compounds also display the negative
response with a double-flame detector 22 .

C~

~8 A

Fig. 3.10. Plot of AFIO response for thiophene against ba~kground current for
various alkali metals. 1 = NaCl; 2 = KC1; 3 = RbCl; 4 = CsCl. x = FlO response
at optimum hydrogen and nitrogen flow-rates. (From ref. 36.)
33

IV Bead: 05 075 1.0 9 1.5 2.0 3.0mm 1.0.


.~ 103 ,·0
;
iii fbdine p p
J
0
0- i
1 I
I
I
, p
./'" \

'1
i t) \
.rr.j ; i . I
, ; i
; ! i
dp i I'Iodinei.
d /, i
i
;
.
I
I
.
'1' i . I
d ; ! .
m
c:
d
j
I !
0
a. 10' ci 30 50 10 50 10 50 10
II!
IV
a: Hyd fLowmlhnin 0
1ti

let

Fig. 3.11. Halogen response profiles at various hydrogen flow-rates and bead
bores. Electrode height, 2 mm. Triplicate injections of 1 ~l of a 1% solution
of chlorobenzene, bromobenzene and iodobenzene. Nitrogen flow-rate, 50 ml/min.
(From ref. 39.)

3.6. RESPONSE TO INDIVIDUAL HETEROATOMS

3.6.1. Phosphorus oompounds

A high sensitivity to phosphorus compounds was found with the earliest ver-
sions of the AFID. The minimum detectability and selectivity of all types of AFIO
are highest for phosphorus compounds 1,2,14,16,21,27,36,47,48 (e.g., 1.5.10- 14 g/
sec 11 , 1.10- 12 g/sec 13 , 5.10- 13 g/sec 9; see also Tables 3.1 and 3.2). In com-
parison with the FlO, the sensitivity is 3-4 orders of magnitude higher.
Similarly to with compounds containing other heteroatoms, the sensitivity of
the AFID to phosphorus compounds depends mainly on the hydrogen flow-rate
(background ionization current); the response increases with increasing back-
ground ionization current (hydrogen flow_rate)l,7,8,27,36. The response is
positive, attaining a maximum at a certain ba~kground ionization current (hy-
drogen flow-rate). The course of the decrease in response at high hydrogen
flow-rates in characteristic, for instance, of the Pye three-electrode system 17
(Fig. 3.12) (an analogous course was also reported for other heteroatoms 49 - 51 ).
The molar response varies between 10 2 and 10 3 C/mole of the compound for all
of the AFID types 14 ,18,27,36 and the selectivity varies from 10 3 to 10 5•
34

+10
l~

+5
/

- 5

-10

Fig. 3.12. Var3ation of the response with hydrogen flow-rate. 1 = Background


current (x 10- A); 2 = peak height of trimethyl phosphate (x 10-9 A); 3 = noise
level (x 10- 11 A); 4 = peak height of hexadecane (x 10-9 A). (From ref. 17.)

3.6.2. Nitrogen aompounds

In 1966, Ruyle et al. 52 and Wells 53 found that the AFIO yielded a selec-
tive response to nitrogen compounds. The response is positive and 2-3 orders
of magnitude higher than that of the FI0 11 ,48,51,54, depending on the detector
design. The AFIO requires the presence of potassium, rubidium or caesium salts
to give a selective response, no increase in ionization current being found
when a sodium salt is used 36 ,48,52-54 (Fig. 3.13). Rubidium has been found to
be the most suitable alkali metal for nitrogen compounds (mainly due to the
stability, source life and simplicity of the preparation of the source, partic-
ularly that of the jet tips)11 ,23,28,30,49,51 ,52,54-56. The AFID response to
nitrogen compounds is negative 27 ,39 (see Fig. 3.7) at high hydrogen flow-rates
(high background ionization currents) or at low air flow-rates.
The response to nitrogen compounds is greatly affected by the distance of
the electrode from the alkali metal source 52 ,54,55 (Fig. 3.13), this relation-
ship being affected by the hydrogen flow-rate (Fig. 3.14). With increasing
hydrogen flow-rate (background ionization current), the response of an AFID
35

...c>.
...
...
.Q

UJ
(/I
z
o
CL
(/I
IJJ
0::

2 3 4 5 6

ELECTRODE POSITION tnm abave bead)

Fig. 3.13. Dependence of nitrobenzene response on the collector electrode posi-


tion. (Reproduced from ref. 54 with permission.)

c:
::J

...>.
...CI
-...
.c
C

w
(/I
z
o
a.
VI
IJJ
0::

2 3 4 5 6
ELECTRODE POSITION (mmabove bead)

Fig. 3.14. Effect of hydrogen flow-rate on nitrobenzene response profile.


(Reproduced from ref. 54 with permi.ssion. J

provided with a jet tip increases to a maximum, than decreases again 10 ,11,27,29,
36,51,56. A similar relationship also holds with variations in the air flow-
rate 39 ,56 (cf., section 3.8.6).
Under the conditions stipulated by the manufacturer, the minimum detectability
of the Pye detector is 1.1.10- 11 g/sec of azobenzene 51 , the selectivity for
36

2 3
hydrocarbons is 10 4-10 5 and that for phosphorus compounds is 1:10 to 1:10 (see
also Table 3.2).
12
The minimum detectability of the detector provided with a jet tip is 1.10- gl
3
sec 11 ,55,57 and its N:C selectivity is 10 (ref. 55).

3.6.3. Ha~ogen aompounds

Giuffrida and Karmen found that the AFIO yielded selective responses to
chlorine, bromine and iodine compounds for both single- 1,2 and double-flame 21
detectors. This response is higher than that to hydrocarbons and about an order
of magnitude higher than that of the FlO. The positive response (increase in
the ionization current of the detector in the presence of halogen compounds)
when using a sodium salt is higher with bromine compounds than with chlorine
and iodine compounds2,21 ,26,37,47; when employing potassium, rubidium and
caesium salts, the order of the positive response levels depends on the detector
design and the distance of the electrode 38 ,47. The level of the positive re-
sponse depends on the cation used; it increases in the order47 Na+<K+<Rb+<Cs+
with the double-flame detector, whereas with the single-flame detector this
relationship is more complex and depends on the hydrogen flow-rate.
With the single-flame detector, mainly with the AFIO type provided with a
jet tip, the response to chlorine, bromine and iodine compounds is negative
under certain conditions8 ,9,36-39. With this type of AFIO, the character and
the level of the response to halogen compounds depend, above all, on the back-
ground ionization current (hydrogen flow-rate) and the cation used (Fig. 3.5)37.
The response to compounds containing all the three halogens is positive when
a sodium salt is used, and it is higher than that of the FlO for chlorine and
bromine compounds. With the use of a potassium salt (Fig. 3.58), the response
to iodine compounds is positive over the whole range of background ionization
currents and it surpasses that of the FlO, whereas the response to chlorine
and bromine compounds becomes negative at a certain background current. With
caesium salts (Fig. 3.5C) and simi~arly with rubidium salts, the response to
compounds containing all three halogens changes to negative. The lower the
atomic number of the halogen, the smaller is the background ionization current
(hydrogen flow-rate) at which the change in polarity of the response occurs.
The course of these relationships depends on the height of the collector elec-
trode 38 ,39 (see section 3.8.2) and on the diameter of the jet tip bore 39 • If
follows from Fig. 3.11 that the larger the diameter of the jet tip bore, the
lower is the hydrogen flow-rate (background ionization current) that induces
a negative response, as mentioned in section 3.5. The larger the bore diameter,
the ·lower is the flame. Hence it heats the major part of the surface of the
37

surface of the salt source and, consequently, the background ionization current
is higher at equal flow-rates of the gases. The flow-rate dependences of the
response related to the bore of the jet tip again qualitatively follow the
relationship between the response and the background ionization current.
In accordance with the operating conditions, the selectivity of the AFlO
response to halogen compounds is determined by an increase in their response
in comparison with hydrocarbons or by a negative response that may be up to
three orders of magnitude higher than that for hydrocarbons 38 . The detection
limit with the negative response is about 1 ng 38 and the minimum detectable
mass rate with a positive response is 1 • 10- 9 g/sec 27 •
Fluorine compounds behave in the same manner as hydrocarbons in the AFlO,
i.e., fluorine does not increase the ionization current. An exception to this
rule is the lower response of the double-flame detector when using a caesium
salt 21 ,47. Karmen and co-workers 58 ,59 described a double-flame AFlD in which
CaC1 2 is situated on a screen between the two systems, below the source of the
alkali metal salt. HF is produced from the organic fluorine compounds in the
lower flame and it reacts with CaC1 2 to give HC1. The HCl vapour then causes
an increase in the ionization current in the upper system.

3.6.4. HydPoaarbons

The sensitivity of the AFlO to compounds that contain no heteroatom to which


the AFlO responds (hydrocarbons, alcohols) is, at best, as high as that of the
Fl0 14 - 17 ,21,36,39,47. The response and character are again primarily functions
of the background ionization current (hydrogen flow-rate). With increasing
background ionization current (hydrogen flow-rate), the response decreases 1,8,
9,17,36 below the response of the FlO and becomes negative8 ,9,17,36 (Fig. 3.11),
starting at a certain value of the background current (hydrogen flow-rate). The
relationship between the response and increasing flow-rate of the carrier gas 39
(see Fig. 3.6) displays the same course. The degree of the decrease in the
positive response of the AFlO in comparison with that of the FlO also differs
in accordance with the detector design, amounting to 1 - 3 orders of magni-
tude 14 - 16 ,36,54, whereas the negative response is approximately commensurate
with the positive response8 ,9,17,36. The increase in the selective response
to compounds with a heteroatom compared with hydrocarbons in thus given both
by the increase in the response to compounds with heteroatoms compared with
the FlO response and by a decrease in the hydrocarbon response compared with
the FlO.
38

3.6.5. SuLphur aompounds

Dressler and Jan~k42 found that the response of an AFID with a jet tip of a
compacted alkali metal salt could be negative. The sensitivity depends on the
background ionization current and on the kind of alkali metal, and it increases
with increasing background current and atomic number of the cation 42 ,45. The
course of the changes in the polarity of the response and peak shape for sulphur
compounds as a function of the background ionization current is described in
section 3.5 (Fig. 3.8). When using K2S04 and at a background ionization current
of 6 • 10- 9 A, the molar response of thiophene is 6 C/mole42 ,43 and the minimum
detectable mass rate is 1 • 10- 10 g/sec (detector noise 4 • 10- 12 A). The AFID
response to sulphur compounds is negative under these conditions, whereas it
is positive for other heteroatoms (except lead and tin) (Fig. 3.9). The AFID
response is about 20 times higher than the FID response. The negative peak of
sulphur compounds displays tailing in some instances, which seems to be con-
nected with the character of the surface of the jet tip. It was observed when
the upper surface of the jet tip was flat, but not with conical shapes 45 .
A positive response (RbCl) has been described for the pye three-electrode
AFID (see Fig. 3.1d)41,49. The dependence of the response on the hydrogen flow-
rate reaches a maximum at higher flow-rates compared with the situation with
nitrogen compounds 43 , or at lower air flow-rates 18 ,41. The best selectivity
over phosphorus and nitrogen compounds (1:1 and 10:1, respectively) occurs at
a low air flow-rate (100 ml/min)18,41; the minimum detectable mass rate is about
3 • 10 12 g/sec (noise 5 • 10- 14 A) under these conditions 41 . A positive response
was also found with the double-flame detector when potassium, rubidium and
caesium salts were used (the response ratios, compared with the FID, were 2.6,
11.6 and 13.8, respectively)47.

3.6.6. Arsenia aompounds

Ives and Giuffrida 48 reported the response of a Single-flame AFID to arsenic


compounds to be five times (KCl) to thirty times (esCl) greater than that of
the FID (Fig. 3.1a). The double-flame detector 16 also responds to arsenic com-
pounds. The response is lower than that of the FID (about 3 • 10- 1 C/mole), but
is approximately ten times greater than that to hydrocarbons. A response is
also obtained with arsine, for which the minimum detectable mass rate is
2.2 • 10-8 g /sec.
39

3.6.7. Boron oompounds

At the optimum hydrogen flow-rate (which is substantially higher than the


optimum flow-rate for phosphorus compounds; when using CsBr it is 42 ml/min
for boron compounds and 32 ml/min for phosphorus compounds), the pye three-
electrode AFlD yields a response to boron compounds that is about 50 times
greater than that of the FlD. The minimum detectable mass rate for decaborane
is 1.6 • 10- 12 g/sec and the response ratio of phosphorus and boron compounds
is 330:1 50 •

3.6.B. Tin and lead oompounds

The dependence of the response of an AFlD equipped with a jet tip of com-
pacted alkali metal salt on the background ionization current for tin and lead
compounds displays a course similar to that for sulphur compounds (Fig. 3.15).
The negative response increases with increasing background current (hydrogen

AFlD/FID
B
A

AFID/F1D

-10'

10"111 1(f' 10'" A A

Fig. 3.15. Dependence of the response on the background current. A, Sodium salt;
B, potassium salt. 1 = Triethoxysilane; 2 = tetraethyllead; 3, tetraethyl tin.
(From ref. 12.)
40

FlO

I-
0:
~
VI

I-
AFIO (Na)
0:
<{
I-
UJ VI
VI
Z
o
a.
VI
UJ
0:

TIME

Fig. 3.16. Chromatograms of tetraethyltin (1). Column, stainless steel (68 cm x


6 mm I.D.), packed with 20% PEG 1500 on Chromosorb W, 55 0 C. (From ref. 12.)

flow-rate) and, at a constant background current, it increases with increasing


atomic number of the cation used (with a sodium salt the response to lead com-
pounds is still positive at a background ionization current of 10- 10 A). Tin
compounds have higher molar responses than lead compounds (Table 3.1). The peaks
of compounds containing these two heteroatoms show tailing, and the elution
time is prolonged (Fig. 3.16; cf., negative response 3.5). After repeated
injections of larger amounts of these compounds, a black deposit appears on the
surface of the jet tip while the background ionization current is decreasing 12
A negative response to tin was also reported for the double-flame detector;
the response is commensurate with the values obtained with the single-flame
detector, but the reproducibility of the response is poor 15

3.6.9. Sirieon evmpounds

The response of the AFID with a jet tin is positive within the range of back-
ground ionization currents from 1 • 10- 10 to 2 • 10-8 A12. Except for the sodium
salt and background currents up to 1 - 10- 9 A, the AFID response always surpasses
that of the FID (Fig. 3.15). The decrease in response, in comparison with the
FID, is due to the fact that, in the range of background ionization currents
mentioned, the detector still functions as an AFID (see section 3.7), but with
41

a hydrogen flow-rate that is not optimal. A comparison of the molar responses


with those of compounds containing other heteroatoms is given in Table 3.1.

3.7. INFLUENCE OF COMPOUND STRUCTURE ON DETECTOR RESPONSE

In the quantitative interpretation of chromatograms obtained by means of


the AFID, it has often been assumed that the response is due to the heteroatom
only and that, consequently, the response is proportional only to the percent-
age of the heteroatom contained in the molecule of the sample compound. However,
the results reveal that this is not always so and that the detector design and
operating conditions seem to play an important role.
Aue et al. 40 showed that the response of the AFID with a jet tip (Rb 2S04)
to phosphorus, chlorine, bromine, iodine and sulphur compounds is invariably
proportional to the percentage of the heteroatom contained in the molecule with
both a positive and a negative response. Nitrogen compounds do not behave in
this manner, however40 ,60. In addition, it was found for another version of this
type of AFID that the structure of sulphur compounds affected the molar re-
sponse 43 . Ives and Giuffrida 48 also reported for a type a detector (Fig. 3.1)
that the response is proportional to the number of nitrogen atoms in the
molecule and independent of the structure. Karmen 21 found a proportionality of
the responses to mono-, di-, tri- and tetrachloromethane with a double-flame
detector, but this proportionality was not found with another double-flame
detector 15 . The decrease in the contributions of the second and the other
chlorine atoms to the total molar response of multi-substituted chlorine com-
pounds is attributed to the decrease in the number of chlorine atoms that reach
the upper system as a result of diffusion processes 34 . The molar responses of
various bromine compounds are not identical; the responses to phosphorus com-
pounds also depend on the structure, and the presence of two phosphorus atoms
in the tetraethyl pyrophosphate molecule does not substantially increase the
molar response in comparison with compounds containing one phosphorus atom 15 •
The response of the pye three-electrode detector (Fig. 3.1d) also depends
.
on the solute structure for nltrogen, phosphorus 51 ' 61 and boron 50 compounds.
However, in the series of dimethyl-, diethyl- and diisopropylnitrosamines, the
response is proportional to the percentage content of nitrogen 30 • The response
of the AFID in which the alkali metal is brought into the flame in the gaseous
phase also depends on the structure of the compound 14 •
The frequently conflicting conclusions reached in studies of the effects of
the structure of sample compounds on the molar response may be due to the fact
that with an AFID with a jet tip, the contributions of the AFID and FID responses
are combined for compounds with heteroatoms where the AFID response is not too
42

different from the FlO response (up to one order of magnitude difference in
the response levels). Hence, in addition to the heteroatom, the carbon skeleton
also contributes to the total molar response of the compound. The smaller the
amount of the alkali metal in the flame (lower background ionization current),
the greater is the contribution of the carbon skeleton, and the molar responses
of compounds with the same heteroatom differ considerably. The contribution
of the carbon skeleton decreases only at higher background ionization currents
(the FlO contribution to the total response) and the molar responses of various
bromine compounds are about the same (Fig. 3.17). When substracting the con-
tribution of the carbon skeleton from the total response of a compound, the
contribution of the halogen atoms is invariably the same even with multi-sub-
stituted compounds (Table 3.4)37.

1.5

to

Fig. 3.17. Dependence of the relative response on the background current using
a sodium salt. 1 = Bromocyclohexane; 2 a bromotoluene; 3 = bromocymene. (From
ref. 37.)

Fig. 3.18 shows the dependence of the responses of compounds with an SH


group on the heat of combustion. The decomposition of the compounds in the flame
leads to increased flame temperatures and to an increase in the ionization
current that, to a certain extent, compensates for the variations in the ioniza-
tion current caused by specific interactions of the particular heteroatom62 •
43

TABLE 3.4

MOLAR RESPONSES OF HALOGEN COMPOUNDS

From ref. 37.

Compound Ionization efficiency


Clmol compound Clmol Cl
Chlorobenzene 1.38 1.38
1,4-Dichlorobenzene 2.79 1.40
1,3,5-Trichlorobenzene 4.10 1.37
1,2 ,4 ,5-Tetrachlorobenzene 5.53 1.38
Bromocyclohexane 2.4
4-Bromotoluene 2.5
B-Bromostyrene 2.2
Bromobenzene 2.4
2-Bromocymene 2.5

elmol

o 2000 4000 k J Imol

Fig. 3.18. Dependence of the ionization efficiency of some mercaptans on increase


in the heat of combustion. (From ref. 62.)

3.8. INFLUENCE OF MAIN OPERATIONAL PARAMETERS ON DETECTOR RESPONSE

3.8.1. Voltage and polarity of the electrodes

The AFID response to phosphorus compounds increases with increasing voltage


applied to the detector electrodes 33 ,35,39,47,63,64, commonly attaining a
constant value at a certain voltage (saturation current range), after which the
current increases again. The course of the potential-current characteristic of
44

the detector (interrelation between ionization current and applied voltage)


depends, however, on the background ionization current35 ,63 (which is determined
by the hydrogen flow-rate or by electrical heating of the alkali metal source);
the higher the background ionization current, the higher is the potential causing
the saturation of the current (Fig. 3.19). For compounds with other heteroatoms,
the course of the potential-current characteristic of the AFIO depends on the
detector design. In double-flame detectors, the potential-current characteristic
for chlorine, bromine and iodine compounds is similar to that for phosphorus
compounds 64 . With a single-flame AFIO with the jet tip made of compacted alkali
metal salt, the response to halogen compounds also increases with increasing
voltage, but it is negative (depending on the alkali metal used) (see section
3.6.3); the response to nitrogen compounds remains approximately the same 39 •

+ I.1Q-l0( .10- 8 ) I Al
15 (4.5) , ___ 2
/ / 3
10(3)/ /
I /"
5 I .
/ ,/
,,,-I /

".
-E IV 1200 100/~' 100 200 +EIVI
I 1
;~
.1/
,I /
10 ( 3
.I
,I I
/
I )

-I-~ 15 (4.5)
I
- 1.10-10 (.10- 8 ) I Al

Fig. 3.19. Potential-current characteristics of AFIO with jet tip salt. 1 = H2


11.7 ml/min, range 5.10-10-15.10- 10 A; 2 = H2 = 12.9 ml/min, range 5.10- 0_
15,10-10 A; 3 = H2 = 13.7 ml/min, range 1.5,10-8-4.5,10- 8 A. (From ref. 63.)

The polarity of the double-flame detector fundamentally affects both the


level and shape of the responses of both the upper system (AFIO) and the lower
system (FlO) of this type of detector. The electrical systems affect each other;
the least interaction occurs with a + - + - connection (jet and collector
electrode of the lower system, jet and collector electrode of the upper system).
With a + - - + connection minor distortion of the peaks in the AFIO system
45

occurs and in connection with a negative polarity of the FlD jet distortion
and inversion of the peaks in the FlD system occur 16 ,47.
Peak distortion and inversion also occur with single-flame detectors if the
system contains a third "gate" electrode (cf., Fig. 3.2), depending on the
voltage applied (Fig. 3.20). The voltage that induces peak distortion and
inversion depends on the structure of the compound 46 .

inversion zone

specific inversion potential

A.,:
Ju I

-240
potential IV)

Fig. 3.20. Peak s~ape deformation and inversion as a function of gate electrode
potential. (From ref. 46.)

Different electrical configurations inside the Carlo Erba detector (Fig. 3.3)
affect the response level and polarity of phosphorus and nitrogen compounds
and hydrocarbons 20 (see Fig. 3.21).

3.8.2. Height and shape of the aolleator eleatrode

With other operating conditions of the detector constant, the sensitivity of


the jet-tip AFlD for phosphorus, nitrogen, sulphur, chlorine, bromine and
iodine compounds depends on the distance of the collector electrode from the
surface of the salt source 18 ,23,34,38-40,54,55,65. The dependence of the re-
sponse on the electrode distance is different for each heteroatom. At hydrogen,
nitrogen and air flow-rates of 33, 50 and 215 ml/min, respectively, and with
a jet tip of Rb 2S0 4 with a 1-mm bore diameter, the negative response of sulphur
compounds decreases with increasing electrode distance; the negative response
to chlorine compounds first increases and then decreases; for other heteroatoms
46

0\ 0\
C C
N
d CIJ
C
c CIJ
o N
:c C
CIJ
o .0
"5 o
N
:I: «

N mode Pmode NPmode

N N

0\
C
0
0
~
<Xl
U
P In
Lll ~
N In
X

Fig. 3.21. Analysis of a test sample in the NPD-40 working modes. (From ref. 20.)

the response increases with increasing distance of the electrode. However, the
character of these interrelations depends on the operating conditions of the
detector: the hydrogen flow_rate 38 ,39,54 (Fig. 3.14), the cation used 54 (Fig.
3.13), the shape of the collector electrode and the shape of the jet tip66. As
the dependence of the response on the electrode distance from the source differs
for each heteroatom, the selectivity of a compound containing a certain element
compared with a compound with another element attains its maximum value at a
certain distance.
As noted in section 3.4, under certain conditions the AFID response can be
a function of the background ionization current. Comparisons of the responses
of phosphorus, chlorine and sulphur compounds at different electrode distances
47

from the salt source (Table 3.3) have shown that if the background ionization
current is kept constant for the AFIO with a jet tip when changing the distance
of the electrode (by changing the hydrogen flow-rate), the response of these
compounds also remains constant 34 . Hartmann 11 also reported that for this type
of AFIO the position of the electrode in the detector is not critical for the
responses of phosphorus and nitrogen compounds, but that the flow-rates of the
gases have to be adjusted so as to attain the optimum response.
Mostly, data on the relationships between the response and the position of
the electrode have not been accompanied by data on the background ionization
current. However, it follows from Figs. 3.6 and 3.10 that, also for this type
of design, the AFIO response to compounds with various heteroatoms is a certain
(always different) function of the background ionization current (given by the
hydrogen flow-rate), and that these relationships do not differ in quality from
those established by Dressler and Janak 37 (see Fig. 3.5). It can be concluded
from Table 3.3 that the background ionization current increases to its maximum
with increasing distance of the electrode at constant flow-rates and that the
course of this relationship varies for different shapes of the electrode and
different hydrogen flow-rates. So, considering these variations in background
ionization current with varying distance of the electrode from the salt source,
the variations in the responses to compounds with various heteroatoms at dif-
ferent distances of the electrodes can be related to the corresponding varia-
tions in the background ionization current at the respective points in the
flame. There is a qualitative resemblance of the dependence of the response on
the background ionization current to that on the electrode distance. The
influence of the hydrogen flow-rate on this relationship can also be related
to the background ionization current. At higher hydrogen flow-rates the above
relationships shift into the region of higher background ionization currents.
Again, qualitative agreement can be observed. With AFIO designs in which the
mobile polarizing electrode is situated above the detector jet, the background
ionization current decreases with increasing distance of this electrode from
the salt source, while the response ratio of phosphorus and nitrogen compounds
shifts with respect to nitrogen compounds. In contrast, the P:N response ratio
increases at a close distance of the electrode to the source (higher background
current)65.
The lateral distance of the electrode from the flame also affects the re-
sponse to nitrogen compounds. Increased distances result in decreased responses
to nitrogen compounds, whereas the response to phosphorus compounds remains
unaltered 23 . The position of the electrode relative to the flame seems to be
related to the dependence of the AFIO response on the diameter of the collector
48

10 Phosphorus

o

.D
~
~
51c:
o

1.1O-11~..,
~
Q)

a: 0.1 Background
~
::J
a.
g..,
..,
to
::J
0.01 1 . 10-12 -;
0~--~---7~~3~--~4

Probe gap, mm

Fig. 3.22. Variation of detector response with probe height. Flow-rates: air,
100 ml/min; nitrogen, 30 ml/min; hydrogen, 30 ml/min. (From ref. 18.)

electrode. The response is higher for larger diameters of the electrode 34 ,66;
however, this relationship also depends on other operating parameters, such as
the hydrogen flow-rate and the distance of this electrode from the source. As
a result, the response can also decrease with increasing diameter of the
electrode under different operating conditions. In this instance, the relation-
ships also follow the variations in the background ionization current and, at
a constant background current, the response may be constant 34 .
Fig. 3.22 shows the dependence of the response to phosphorus, nitrogen and
sulphur compounds and that of the background ionization current on the distance
of the probe electrode for the Pye three-electrode detector. These relation-
ships vary considerably depending on the flow-rates of hydrogen, air and
nitrogen, even though the basic character depicted in Fig. 3.22 remains un-
altered. The maximum responses (attained at various ratios of the flow-rates
of the gases) for the compounds with all the mentioned heteroatoms decrease
with increasing distance of the electrode form the flame 18 .

3.B.3. Diameter of the tip bore

The diameter of the jet tip bore affects the shape of the flame and, thereby,
the size of the salt surface that is contacted by the flame and hence the flame
temperature at the point of contact with the salt. The larger the diameter of
the jet bore, the broader is the flame at its base where it covers a larger
49

, " , " ! , ! I
mm

Fig. 3.23. Alkali salt designs and approximate flame configurations. (From ref.
66. )

surface of the salt. A greater number of alkali metal atoms are introduced into
the flame, and the background ionization current is increased at constant flow-
rates of the gases 27 . With a constant background ionization current (due to the
decreased hydrogen flow-rate with increasing diameter), the response to nitrogen
compounds increases with increasing diameter of the jet bore (by a factor of
about 4 from 0.6 to 3.00 mm). With halogen and phosphorus compounds the de-
pendence of the response on the diameter of the jet bore produces a maximum for
both positive and negative responses (halogens)39. The effect of the diameter
of the jet tip bore on the negative response is described in section 3.5.
The detector response to organophosphorus compounds with salt sources having
a large flame-to-salt (Rb 2S04 ) contact surface (Fig. 3.23, II, IV and V) is
about 50 times greater than that for salt sources with a small contact surface
[the hydrogen and air flow-rates are adjusted so as to result in maximum re-
sponse; the hydrogen flow-rate is always lower for types II, IV and V (27-30
ml/min) than for the other types (45-50 ml/min)6~.

3.8.4. Cations and anions of the atkati metat satt

The vapour pressure of alkali metal compounds increases with increasing


atomic number of the alkali metal and the ionization potential of the alkali
metal decreases in the same order. This means that, with a constant hydrogen
flow-rate, different amounts of salts of the individual alkali metals are
01
o

TABLE :1.5

DEPENDENCE OF AFrO: FlO RESPONSE RATIOS ON THE TYPE OF ALKALI METAL AT CONSTANT GAS FLOW-RATES

(A) Sulphates; (B) chlorides.

Compound Li Na K Rb Cs

(A) Double-flame detector (from ref. 47):


Fluorobenzene 0 0 0.35 0.35 7.80
Chlorobenzene 0.26 2.30 15.00 30.81 117.50
Bromobenzene 0.16 3.85 5.00 14.83 67.60
lodobenzene 6.93 2.82 8.35 69.10 242.20
Diisopropyl methanephosphonate 7.59 144.70 247.20 418.80 751.40
Thiophene 0.12 0.76 2.62 11.63 13.80
Pyridine 0.13 0.11 0.98 1.18 3.10
(B) Single-flame detector, type a (Fig. 3. 1) (from ref. 48) :
Triphenylamine 30 60 150
Triphenylphosphine 10 000 11 000 12 000
Triphenylarsine 5 10 30
51

brought into the flame and, in addition, the proportion of the ionized atoms
of the individual alkali metals is also different. Therefore, if all other
operating parameters of the detector remain constant, the background ionization
current will increase in the order Na<K<Rb<Cs. As previously noted, the AFID
response depends on the amount of alkali metal present in the flame; the sen-
sitivity of the AFID should also increase in the above order of the alkali
metals. Table 3.5 shows that this is so both with detectors having the salt
melted on a carrier48 and with double-flame detectors 47 . In addition to the
effect on the response, the purity of the salt employed affects the detector
noise and thereby also the minimum detectability.
The monitoring of the responses when using various alkali metals at the same
background ionization current (by reducing the hydrogen flow-rate in the order
Na>K>Rb>Cs) allows us to compare the responses with equal amounts in the system.
Also in this instance the sensitivity of the AFID with a jet tip for phosphorus,
nitrogen, sulphur, silicon, lead and tin compounds increases in the order
Na<K<Rb<Cs (see Figs. 3.10 and 3.15) for both positive and negative responses 12,36
In a detector having the salt melted on a coil, Giuffrida et al. 25 found an
increase in response for phosphorus compounds in the order Na 2S0 4 < Cs 2S0 4 <
Rb 2S0 4 < K2C0 3. However, carbonates yield a greater response than sulphates with
the same background ionization current (see Table 3.6). Probably for this reason
the response is highest when using potassium carbonate.

TABLE 3.6

DEPENDENCE OF AFID RESPONSE ON THE ANION OF THE SODIUM SALT

Response level is relative to the response for Cl-. N2 flow-rate, 60 ml/min;


air flow-rate, 660 ml/min. (From ref. 36.)

Anion AFID response H2 flow-rate


Chlorobenzene Diisopropyl (ml/min)
methanephosphonate
Cl- 1. 00 1. 00 54.0
Br- 1.47 2.56 36.0
S02- 1.60 2.23 50.0
4
C02- 2.28 4.62 36.5
3
N0 3 2.02 5.05 33.0
52

The same dependence on the atomic number of the alkali metal as for compounds
with other heteroatoms is valid for the detector sensitivity in the case of
halogen compounds. However, the type of alkali metal also affects the polarity
of the response (Fig. 3.5). When using a sodium salt, the AFrO response to
chlorine, bromine and iodine is positive over the whole range of the observed
background current (10- 10 _10- 8 A). When a potassium salt is used, the response
is positive within this range of background currents only to iodine compounds,
the response to bromine and chlorine compounds becoming negative at certain
values of the background current. With rubidium and caesium salts the responses
are negative for all three halogens 37 . The higher the atomic number of the
alkali metal employed, the lower is the background ionization current for which
the response becomes negative 12 ,36,37. For the double-flame detector, an approx-
imately constant ratio of the responses to phosphorus and chlorine compounds
has been reported for all alkali metals at the same background current67
For the AFrO with a jet tip, the influence of the salt anions on the response
at a constant beckground ionization current can be seen from Table 3.6. The
response for phosphorus and chlorine compounds increases 36 in the order Cl- <
Sr - <S042- <C0 32- <N0 3- . With a detector having the salt melted on a carrier, the
dependence 32 is OH - < C0 32- < P0 3- - 2- -
4 < Sr < SO% < Cl at optimum hydrogen flow-rates
for each salt. Giuffrida and co-workers 2 ,48 found that for an AFrO with the
salt melted on a carrier, the response to halogen compounds is suppressed when
using KCl and KSr.

3.B.5. Detector temperature

With increasing temperature of the detector, the background ionization cur-


rent and the detector response to phosphorus and nitrogen compounds increase,
whereas the response to hydrocarbons decreases. With increasing temperature
the selectivity increases by a factor of about 5 in the range from 200 to
400 0 C19 • With the Pye three-electrode detector the response decreases, starting
from a temperature of about 250 0 C29
With a detector having the jet tip moulded into the detector jet, the re-
sponse to phosphorus compounds also decreases, at optimum N2 : H2 flow-rate
ratios, with increasing temperature of the detector (by almost three orders of
magnitude within the range 200-300 0 C), whereas it does not vary very much
for nitrogen compounds. The N2 : H2 flow-rate ratio at which the maximum sen-
sitivity is obtained decreases with increasing detector temperature 1D
53

,). 8. 6. Carrie1' gas, a'ir

When helium is used as the carrier gas, the sensitivity of the AFID for
phosphorus and nitrogen compounds is greater than the sensitivity in nitro-
gen 4 ,48,57,65 (using a jet of about 1 mm I.D.; this difference levels off when
a larger diameter is used, e.g., 1.5 mm48 ) , it decreases with chlorine com-
pounds 4 (using KC1) and hydrocarbons 65 and, therefore, the selectivity of the
responses to phosphorus and nitrogen compounds is higher when helium is used.
The minimunl detectable mass rate is also lower when employing helium65 • The
optimum hydrogen flow-rate with respect to the sensitivity of the AFID is
different, being lower with helium65 .
With nitrogen compounds, detectors with jet tips of compacted salt yield
negative responses in helium at lower hydrogen flow-rates than they do in
nitrogen. For this reason, the positive response is lower at equal flow-rates
of the gases with helium (particularly at higher hydrogen flow-rates) and the
negative response is roughly the same. With helium, the negative response (at
the optimum air flow-rate) is about 1.5 orders of magnitude higher than the
positive response and the two responses are roughly the same with nitrogen
(Fig. 3.7). With halogen compounds, both the positive and negative responses
with helium are higher than those with nitrogen (Fig. 3.6)39.
The positive response to phosphorus, nitrogen, chlorine, bromine and iodine
compounds and hydrocarbons initially increases with decreasing flow-rate of
the carrier gas (Figs. 3.6 and 3.7). At a certain flow-rate (depending on the
kind of heteroatom and the hydrogen flow-rate), the response attains a maximum
and then changes gradually with further decrease in the flow-rate to reach a
negative response that subsequently continues to increase. The positive re-
sponse to chlorine and iodine compounds is about 2-3 orders of magnitude
and that to hydrocarbons about one order of magnitude lower than the maximum
negative response 39
The increased flow-rate of the carrier gas cools the flame and causes a
decrease in the background ionization current. This means that, at the same
time, the mentioned relationships for the response variations in connection
with the decreased flow-rate of the carrier gas follow the increase in the
background ionization current. Negative responses for halogen compounds occur
at flow-rates of the carrier gas that are lower (higher background currents)
the higher is the atomic numbers of the halogen. The same is true for the
dependence of the response on the background ionization current. The higher
the hydrogen flow-rate, the higher is the flow-rate of the carrier gas at which
a negative response occurs. The same holds for nitrogen compounds and hydro-
carbons (Fig. 3.7).
54

A comparison of the responses obtained at equal background currents (by


regulating the hydrogen flow-rate) has shown that, also in this instance, de-
creased flow-rates of the carrier gas lead to increased responses to nitrogen,
phosphorus and chlorine compounds' ,27, However, the variations are substantially
lower than those at a constant hydrogen flow-rate,
The dependence of the negative response for sulphur compounds on the flow-
rate of the carrier gas displays a maximum that, for various hydrogen flow-rates,
approximately matches the same flow-rate of the carrier gas (about 60 ml/min)40,
A maximum has also been observed for the dependence of the response on the flow-
rate of the carrier gas with the Pye three-electrode detector (65 ml/min with
nitrogen compounds)56,
At a constant background current, the response increases' with increasing
air flow-rate at low flow-rates (roughly up to 250 ml/min) where the supply of
oxygen is inadequate, The response subsequently remains constant' ,27 at flow-
rates up to about 500 ml/min, after which it increases again with increasing
flow-rate 27 , The dependence of the response on the air flow-rate exhibits a
maximum (475 ml/min) with the three-electrode detector (type d, Fig, 3,,)56,
This dependence also shows a maximum for detectors in which the alkali is
supplied to the flame in the gaseous phase'4,
With double-flame detectors, the dependence of the response on the air flow-
rate is similar to that of the FID, i ,e" the response increases for flow-rates
up to about 500 ml/min and remains constant at higher flow-rates 47 ,
The use of oxygen instead of air in the double-flame detector leads to a
suppression of the response to phosphorus compounds and an increase in the
response to halogen compounds 22

3,9, DETECTION MECHANISM

Similar difficulties to those experienced in attempts to generalize the


experimental results obtained form the studies on the AFID are encountered when
trying to arrive at an unambiguous explanation of the response mechanism with
this detector, The existing theories can be classified into the following groups,

3.B.1. Solid-phase peaations

2
Karmen ' found in one of his first studies on the AFID that, on introduction
of chloroform into the lower burner of the double-flame detector, both the
ionization current and the light emission characteristics of the alkali metal
used increased in the upper system. When a halogen compound was introduced into
the upper burner, no emission occurred 23 , Karmen concluded that halogen compounds
55

or their decomposition products increase the volatility of the alkali metal


salts. In the flame the ionization current is increased owing to the ionization
of newly formed atoms of the alkali metal. The AFIO responds to halogen compounds
only if the detector allows-contact of the halogen compounds with the alkali
metal salt. Hence the generation of the AFIO response is attributed to the
reactions of the heteroatom-containing combustion products with the alkali metal
in the solid phase.
However, the presence of halogen and phosphorus compounds in the flame of
the single-flame AFIO also leads to the reduction of the light emission of the
alkali meta1 24 ,34. Depending on the hydrogen flow-rate, the emission of the
flame is higher (at low flow-rates) or lower (at high flow-rates) in the pres-
ence of halogen compounds (except fluorine)68,69. In contrast, the emission is
reduced for phosphorus compounds at low hydrogen flow-rates 69 . The variation
in the ionization current of the AFIO during the passage of halogen compounds
through the flame also depends on the hydrogen flow-rate, in that the positive
AFIO response (i.e., the increase in ionization current) changes to a negative
response at higher hydrogen flow-rates (see section 3.6.3). Owing to the dis-
similar courses of the variations in emission and ionization current (see also
section 7.6), these variations cannot be compared directly, however.
Brazhnikov et al .32 studied the weight variations of the alkali metal salt
source of the AFIO when the flame was burning and, later, also when? constant
amount of tributyl phosphate vapour was supplied to the flame. The evaporation
rate of caesium bromide (characterized by the weight loss of the source with
time) remained constant regardless of the presence or absence of the phosphorus
compound (Fig. 3.24).
During the elution of tin and lead compounds from the chromatographic column,
the time-related course of the signal of an AFIO with a jet tip of compacted
salt differs from that of the FI012. With the AFIO, both the retention value
and the standard deviation of the tetraethyl tin peak exceed those of the FlO
(Fig. 3.16). Therefore, it is believed that with certain compounds reactions
take place on the surface of the alkali metal salt (see section 3.5 for attempts
to explain the negative responses of the other heteroatoms).
According to Olah et al. 70 , a mechanism describing the ionization process
as a catalytic reaction can be applied for the AFIO in some instances (see
section 4.4.2).
56

a
III

ai
III
U

-....
a
1.64600

..c
Cl
41
3 1.64500 ......
.........
5 10 15 20 25 30 35 40

Time (h)

Fig. 3.24. Dependence of salt tip weight on its performance time. Flow-rates:
hydrogen, 33 ml/min; nitrogen, 30 ml/min; air, 250 ml/min. (From ref. 32.)

3.9.2. Gaseous-phase peactions

An alkali metal from the source of the alkali metal salt is brought into the
AFID and ionized in the flame, generating a constant ionization current either
thermally:

(3.1)

or by a three-body reaction:

(3.2)

where A is the atom of the alkali metal, and M represents a general flame-gas
molecule.
In explaining the mechanism of the AFID response by reactions in the gaseous
phase, Page and Woolley?1 proceeded from the assumption that in the oxygen-
hydrogen flame there is a higher concentration of hydrogen atoms than that which
would correspond to the equilibrium state (the influence of the slow three-body
recombination of hydrogen atoms):

(3.3)

H + H + M- H2 + M (3.4)
57

The ionization of the alkali metal in the flame is accomplished according


to reaction 3.1, because reaction 3.2 is slow. However, the rate of reaction
3.2 can be increased (in proportion to the amount of the heteroatom, X) in the
presence of a compound containing the heteroatom X. The degree of ionization
of the alkali metal increases in the range from 1 to y2, where y is the ratio
between the true concentration of hydrogen atoms and the equilibrium concentra-
ti on, y = [H] / [}le;J :

(3.5)

X + H ~HX + e (3.6)

(3.7)

According to this theory, the ionization current increases in the presence


of a compound with a heteroatom owing to the increased ionization of the atoms
of the alkali metal that is already present in the flame. The decrease in light
emission is caused by the decrease in the concentration of neutral atoms due
to this increased ionization. Baldwin 72 ascribed the increased emission (observed
at higher alkali metal concentrations in the flame) to self-absorption. As during
the passage through the flame of a compound with a heteroatom the neutral atoms
of the alkali metal are decreasing, self-absorption decreases and, consequently,
light emission increases.
Brazhnikov and Shmidel 13 explained the increase in ionization current by the
production of heavy ions during the combustion of phosphorus and nitrogen
compounds. The mobility of heavy ions is about 1000 times lower than that of
light ions; they combine with the ions of the alkali metal giving rise to even
heavier ions. The alkali metal salts are inhibitors of combustion. The con-
centration of the alkali metal salts decreases owing to the production of heavy
ions, thereby reducing the inhibiting effect of this salt (the flame temperature
increases). The temperature increase leads to greater ionization of the alkali
metal atoms and, thereby, to an increase in ionization current.
Sevcik 62 explained the AFrO response by specific interactions. With halogen
compounds, the electrons produced as.a result of ionization in the flame are
captured and the ionization current decreases owing to the increased recombina-
tion of ions. The reaction of the decomposition products with the alkali metal
atoms that gives thermostable compounds is dominant for sulphur compounds. The
concentration of the alkali metal in the flame decreases and, for this reason,
the ionization current also decreases. The two preceding mechanisms are not
substantial for phopshorus compounds. Various phosphorus compounds with a low
ionization ~otential ar~ produced in the flame and the ionization current
increases.
58

Add P

c
No P

a CPIB~------

6 1. 2 6
min

Fig. 3.25. Effect of addition of phosphorus vapoufi to the upper flame. CPIB =
chlorophenoxy isobutenate; a = lower flame (3'10- A f.s.); b = upger flame
(10- 8 A f.s.); c = lower flame (10- 8 A f.s.); d = upper flame (10- 7 A f.s.).
(Reproduced from ref. 23 with permission.)

Experiments have shown that the ionization current of the AFID increases in
the presence of phosphorus and nitrogen compounds, even if no reactions of the
combustion products of the compounds with the source of the alkali metal salt
can occur, i.e., when the alkali metal is supplied to the detector in the
gaseous phase 14 ,23,24,71,72. The response of a detector provided with this
means of supplying the alkali metal to the flame is comparable to the response
of other types of AFID, viz., about 5 C/g of phosphorus and 0.3 C/g of nitrogen.
Karmen 23 also attributed the response of phosphorus compounds to reactions
in the gaseous phase. If a constant concentration of phosphorus vapour is sup-
plied to the upper burner of the double-flame detector, the response of the
upper system to halogen compounds brought to the lower burner is higher than
if no phosphorus is supplied to the upper burner (see Fig. 3.25). This increase
in response is ascribed to the increased volatility of the alkali metal of the
source caused by the halogen compound in connection with the increased ioniza-
tion of the alkali metal in the upper flame owing to the presence of phosphorus.
This effect was not observed when phosphorus compounds were supplied to the
lower burner.
59

3.9.3. Photoeffects

Brazhnikov et al .73 assumed that the light generated from the burning phos-
phorus compounds is responsible for the photoevaporation of the salt, i.e., the
photons emitted by the flame are absorbed on the salt surface and lead to
evaporation of the alkali metal salt. These atoms of the alkali metal are
ionized in the flame.

3.9.4. Negative response

The causes of negative responses are even more obscure than those of positive
responses. Maier-Bode and Riedmann 19 explained the negative response by changes
in the shape of the flame during elution of the compounds. During the growth
of the flame, when the compounds elute, the salt source is actually situated
in the zone of the flame in which the temperature is lower than it was before
the elution of the compound. Consequently, less salt is evaporated into the
flame and this results in a decrease in the ionization current. ~evtfk62 ex-
plained the negative response to halogen compounds by electron capture and that
to sulphur compounds by the generation of stable compounds (see section 3.9.2).

REFERENCES

1 L. Giuffrida, J. Ass. Offic. Agr. Chern., 47 (1964) 293.


2 A. Karmen and L. Giuffrida, Nature (London), 201 (1964) 1204.
3 L. Giuffrida and N.F. lves, J. Ass. Offic. Agr. Chern., 47 (1964) 1112.
4 J.H. Ford and M. Beroza, J. Ass. Offic. Anal. Chern., 50 (1967) 601.
5 J.R. Wessel, J. Ass. Offic. Anal. Chern., 50 (1967) 430.
6 D.R. Coahran, Bull. Environ. Contam. Toxicol., 1 (1966) 141.
7 D.M. Oaks, K.P. Dimick and C.H. Hartmann, Aerograph Phosphorus Detector,
W-122, Varian Aerograph, Walnut Creek, CA, 1966. .
8 C.H. Hartmann, Aerograph Research Notes, Varian Aerograph, Walnut Creek, CA,
Summer 1966.
9 C.H. Hartmann, Bull. Environ. Contam. Toxicol., 1 (1966) 159.
10 M. Mraz, R. Nemetek, V. ~edivec and J. Flek, Chern. Listy, in press.
11 C.H. Hartmann, Varian Aerograph Tech. Bull. 136-69, Varian Aerograph, Walnut
Creek, CA.
12 M. Dressler, V. Martinu and J. Janak, J. Chromatogr., 59 (1971) 429.
13 V.V. Brazhnikov and E.B. Shmidel, J. Chromatogr., 122 (1976) 527.
14 V.V. Brazhnikov, V.M. Poshemansky, K.K. Sakodynskii and V.V. Chernjankin,
J. Chromatogr., 175 (1979) 21.
15 J. Janak, V. Svojanovsky and M. Dressler, Collect. Czech. Chern. Comrrmn., 33
(1968) 740.
16 V. Svojanovsky, J. Janak and M. Dressler, Collect. Czech. Chern. Commun., 31
(1966) 3925.
17 F.P. Speakman and C. Waring, Co Zumn , 2, No.3 (1968) 2.
18 R.A. Hoodless, M. Sargent and R.D. Treble, J. Chromatogr., 136 (1977) 199.
19 H. Maier-Bode and M. Riedmann, Residue Rev., 54 (1975) 113.
20 G.R. Verga, J. Chromatogr., 279 (1983) 657.
60

21 A. Karmen, Anal. Chem., 36 (1964) 1416.


22 K. Abel, K. Lanneau and R.K. Stevens, J. Ass. Offic. Anal. Chem., 49 (1966)
1022.
23 A. Karmen, J. Chromatogr. Sci., 7 (1969) 541.
24 W.A. Aue, D.L. Stalling, C.W. Gehrke, R.C. Tindle and S.R. KOirtyohann,
Organophosphate-Alkali Interaction, 5th National Meeting of the Society for
Applied spectroscopy, Chicago, IL, June 1966.
25 L. Giuffrida, N.F. Ives and D.C. Bostwick, J. Ass. Offic. Anal. Chem., 49
(1966) 8.
26 D. Jentzsch, H.G. Zimmermann and I. Wehling, Z. Anal. Chem., 221 (1966) 377.
27 M. Dressler and J. Jan~k, Collect. Czech. Chem. Commun., 33 (1968) 3960.
28 D.F.K. Swan, Column, No. 14 (1972) 9.
29 R. Greenhalgh and J. Dokl~dalova, Column, No. 14 (1972) 4.
30 T.A. Gough and K. Sugden, J. Chromatogr., 86 (1973) 65.
31 R.F. Coward and P. Smith, J. Chromatogr., 61 (1971) 329.
32 V.V. Brazhnikov, M.V. Gurev and K.I. Sakodynskii, Chromatographia, 3 (1970)
53.
33 H. Beckman and W.O. Gauer, Bull. Environ. Contam. Toxicol., 1 (1966) 149.
34 M. Dressler, Alkali Flame Ionization Detector, Thesis, Institute of Analytical
Chemistry, Brno, 1969.
35 R.A. Mees and J. Spaans, Z. Anal. Chem., 247 (1969) 252.
36 M. Dressler and J. Janak, Collect. Czech. Chem. Commun., 33 (1968) 3970.
37 M. Dressler and J. Janak, J. Chromatogr., 44 (1969) 40.
38 S. Lakota and W.A. Aue, J. Chromatogr., 44 (1969) 472.
39 K.O. Gerhardt and W.A. Aue, J. Chromatogr., 52 (1970) 47.
40 W.A. Aue, K.O. Gerhardt and S. Lakota, J. Chromatogr., 63 (1971) 237.
41 R.A. Hoodless, M. Sargent and R.D. Treble, Analyst (London), 101 (1976) 757.
42 M. Dressler and J. Janak, J. Chromatogr. Sci., 7 (1969) 451.
43 M. Dressler and J. Janak, Collect. Czech. Chem. Commun., 34 (1969) 1797.
44 W.H. Stewart, Anal. Chem., 44 (1972) 1547.
45 J. Novotny and A. MUller, J. Chromatogr., 148 (1978) 211.
46 F.K. Martens, M.A. Martens, T. Soylemozoglu and A.M. Heyndrickx, J. Chro-
matogr., 140 (1977) 86.
47 .J. Janak and V. Svojanovsky, in A.B. Littlewood (Editor), Gas Chromatography
1966, Institute of Petroleum, London, 1967, p. 166.
48 N.F. Ives and L. Giuffrida, J. Ass. Offic. Anal. Chem., 50 (1967) 1.
49 N. Mellor, J. Chromatogr., 123 (1976) 396.
50 R. Greenhalgh and P.J. Wood, J. Chromatogr., 82 (1973) 410.
51 R. Greenhalgh and M. Wilson, Column, No. 15 (1972) 10.
52 C.D. Ruyle, W.A. Aue, C.W. Gehrke, D.L. Stalling and R.C. Tindle, Application
of the Alkali Flame Detector to Nitrogen Containing Compounds, 5th National
Meeting of the Society for Applied Spectroscopy, Chicago, 1L, June 1966.
53 C.E. Wells, Application of the Potassium Chloride Thermionic Detector to the
Analysis of Nitrogen Containing Drugs, U.S. Food and Drugs Administration
Pesticide Workshop, Kansas City, MO, 1966.
54 W.A. Aue, C.W. Gehrke, R.C. Tindle, D.L. Stalling and C.D. Ruyle, J. Gas
Chromatogr., 5 (1967) 381.
55 C.H. Hartmann, J. Chromatogr. Sci., 7 (1969) 163.
56 M. Butler and A. Darbre, J. Chromatogr., 101 (1974) 51.
57 W. Ebing, Chrornatographia, 1 (1968) 382.
58 A. Karmen and E.L. Kelly, Anal. Chern., 43 (1971) 1992.
59 A. Karmen and H. Haut, J. Chrornatogr., 99 (1974) 349.
60 J.F. Palframan, J. Macnab and N.T. Crosby, J. Chrornatogr., 76 (1973) 307.
61 R. Greenhalgh and W.P. Cochrane, J. Chrornatogr., 70 (1972) 37.
62 J. Sev~ik, Chrornatographia, 6 (1973) 139.
63 V. Svojanovsky, M. Dressler and J. Janak, in H.G. Struppe (Editor), Gas
CJwornatographie 1968, Deutsche Akademie der Wissenschaften DDR, Berlin,
1968, p. 545.
64 A. Karmen and H. Haut, Anal. Chern., 45 (1973) 822.
61

65 G.R. Verga and F. Poy, J. ChPomatogr., 116 (1976) 17.


66 H.K. Deloach and D.O. Hemphill, J. Ass. Offic. Anal. Chem., 52 (1969) 533.
67 A. Karmen, J. GaD ChY'OmatogY'., 3 (1965) 336.
68 A.V. Nowak and H.V. Malmstadt, Anal. Chem., 40 (1968) 1108.
69 W.A. Aue and R.F. Moseman, J. ChPomatogY'., 61 (1971) 35.
70 K. Olah, A. Sz6ke and Z. Vajta, p. ChY'omatogY'. Sci., 17 (1979) 497.
71 F.M. Page and D.E. Woolley, Anal. Chem., 40 (1968) 210.
72 J. Baldwin, Alkali Metal-Halogen InteY'actions in the Sensitized Flame
Ionization/Flame PhotometY'ic DetectoY', Thesis, University of Illinois,
Urbana, Il, 1968.
73 V.V. Brazhnikov, M.V. Gurev and K.I. Sakodynsky, ChPomatogY'. Rev., 12
(1970) 1.
This Page Intentionally Left Blank
63

Chapter 1

FLAMELESS ALKALI SENSITIZED DETECTORS

CONTENTS

4.1. Introduction . . • • • • • • 63
4.2. The Perkin-Elmer detector •• 64
4.3. The Hewlett-Packard detector 72
4.4. The Tracor detector •••• 73
4.5. The Varian detector . • . • . • . . . • . • 74
4.6. The Detector Engineering Technology detector. 78
4.7. The chemi-ionization detector ••• 83
4.8. Detector life and reproducibility of response 84
4.9. Detectors for halogen compounds 87
References • . • • . . • • • • . . • • • • • 90

4.1. INTRODUCTION

Detectors that utilize the different ionization of phosphorus and nitrogen


compounds in the presence of alkali metals are commonly called alkali flame-
ionization detectors (AFIDs), thermionic detectors (TIDs) or nitrogen-phosphorus
detectors (NPDs), frequently regardless of whether a flame is present in or
absent from the detector system. If detectors bringing about ionization of com-
pounds in a flame are involved, the term (AFID) (Chapter 3) should be used. For
other detectors, in which ionization occurs in the absence of a flame, the term
flameless alkali sensitized detectors (FASDs) has been used recently. With some
types of these detectors, the thermal emission is thought to be the basic process
taking place in the system. It is not possible, however, to consider this process
as a general explanation of the response mechanisms in all FASDs, because other
supposed processes exist with some types. Therefore, the term FASD actually seems
to be the most suitable for flameless detectors with an alkali metal.
Compared with the AFID, the FASD is distinguished by the following main
features: (1) although hydrogen and air are mostly required, the amount (flow-
rate) of the hydrogen used (1-6 ml/min) is not sufficient to sustain an ordi-
nary flame; (2) FASDs of the type described so far exhibit selective responses
only to nitrogen- and phosphorus-containing compounds (for this reason they are
also called NPDs) and are not sensitive to halogen-containing compounds; (3) the
alkali metal source is heated electrically.
64

4.2 THE PERKIN-ELMER DETECTOR

The detector developed by Kolb and Bishoff 1 and manufactured by Perkin-Elmer


represents a transition from an alkali flame-ionization to a flameless alkali
sensitized detector. The hydrogen flow-rate is 35 ml/min for detection in the
P mode and therefore should belong to AFIDs. On the other hand, the hydrogen
flow-rate is low (2 ml/min) in the NP mode and, consequently, the presence of
a flame as classically conceived is not likely.

COLLECTOR
ELECTRODE
RUBIDIUM
r I
,VENT

I ~ f A{HEATI"d_
BEAD
JET

AIR -~PtH,L-q~

COLUMN I POLARITY
EFFLUENT SWITCH
Fig. 4.1. Simplified schematic diagram of the Perkin-Elmer detector. (Reproduced
from ref. 2 with permission.}

A schematic diagram of a three-electrode detector is shown in Fig. 4.1. The


source of the alkali metal is glass containing an alkali metal (rubidium sili-
cate). A negative voltage is always applied to the source, which is heated
electrically. The jet polarity can be changed; if it is positive the detector
operates in the P mode, and if it is negative it operates in the NP mode. The
function of this detector can be seen from Fig. 4.2, which shows the electrode
connections and compares the AFID and FID chromatograms. In Fig 4.2a the alkali
metal source is not heated and the detector operates as an FID, not being af-
fected by the mere presence of the source.
Electrical heating of the source leads to alkali metal emission and to in-
creased sensitivity to phosporus compounds (Fig. 4.2b). Earthing of the jet
(Fig. 4.2c) causes the detector to become selective to phosphorus compounds;
the decomposition products of the phosphorus compounds react with the alkali
metal, giving rise to electrons above the source with a negative potential, which
are sensed by the collector electrode. The electrons resulting from the decom-
position of substances in the flame below the alkali metal source cannot pass
65

P
a FI ,A, b FI+P

0 rB:@ r~@
x
~

tV
(X)
til-e ~e

FlO

~
r-1-@
0
x
tV
fU-e
c.J1
en

P
.J.... C P d
P N+P

rtt~ r:.1=~
.J....

tll-G> u.-Le
N

Fig. 4.2. Operational modes of the Perkin-Elmer detector, exemplified by the


analysis of the ester oil with nitrogen- and phosphorus-containing additives.
Glass column (1 m) packed with 3% OV-1 on Chromosorb WHP. Temperature, program-
med from 270 to 300 0 C at 80 C/min. Compounds: P = P-containing additives (tri-
eresyl phosphates); N = N-eontaining additives (first N peak: diphenylamine).
In (a)-(d) with FID; H2 flow-rate = 35 ml/min; air flow-rate = 400 ml/min.
(a) NPD, FI mode: H2 flow-rate = 35 ml/min; air flow-rate = 400 ml/min; heating
position, "off". (b) NPD, P + FI mode: H2 flow-rate = 35 ml/min; air flow-rate
= 400 ml/min; heating position, 250. (e) NPD, P mode: H2 flow-rate = 35 ml/min;
air flow-rate = 400 ml/min; heating position, 280. (d) NPD, NP mode: H2 flow-
rate = 2 ml/min; air flow-rate = 100 ml/min; heating position, 710. (From ref.
3. )
66

the barrier of the negative potential of the source and therefore the two groups
of electrons are separated. Selectivity for nitrogen compounds is attained by
reducing the hydrogen flow-rate to about one tenth of the value used with the
FID and P-AFID and the air flow-rate to about 100 ml/min (Fig. 4.2d). This low
hydrogen flow-rate (1-3 ml/min) can neither heat the alkali metal source nor
sustain a stable flame burning at the jet; the hydrogen burns around the glowing
source (the latter is electrically heated). The energy of this flame is not high
enough to induce the ionization of hydrocarbons.
Kolb and Bischoff 1 described the response mechanism of this three-electrode
detector in the following way. Nitrogen compounds enter a relatively cool and
diluted zone of the flame around the alkali metal source and are pyrolysed
therei n:

I
- C - N= -+ ·C = NI (4.1)

The radical formed takes an electron from the excited atom of the alkali metal,
giving a cyanide ion:

Rb* + ·C;; NI [IC ;; NI r (4.2)

The cyanide ion moves to the collector electrode, releasing an electron that is
sensed by the collector electrode. The reaction scheme can be seen from Fig.
4.15 (see section 4.7). This explanation of the selective response to nitrogen
compounds is supported by the zero (or low) response to compounds having the
nitrogen bonded in a way that does not allow the formation of a radical [barbi-
turates with a -CO-NH-CO structure, amides (-CO-NH 2 ) and nitrate esters (0-N0 2 )].
l'li th these compounds, the oxygen atom cannot be extracted from the carbon atom;
however, if the nitrogen atom in barbiturates is alkylated, the R-CH 2-N= struc-
ture allows the formation of the 'C ;; NI radical and the response is normal. It
has been reported 2 , as another supporting fact for the above mechanism of the
response of nitrogen compounds, that no selective response to nitrogen compounds
occurs in a hot oxidation flame that favours the formation of nitrogen oxides
(the response to nitrogen compounds in the P mode is about two orders of magni-
tude lower, in fact 4 ).
An analogous reaction mechanism giving PO or P0 2 radicals is considered 1 to
exist with phosphorus compounds:

I· P = 0) + e- -> [(P = 0) r (4.3)

(0 ='p O~ + e- -> [~O = P - QI]-


°/r (4.4)
67

The response of the detector is approximately proportional to the nitrogen


content in the compound. However, the response is also dependent on the compound
However, the response is also dependent on the compound structure, as is obvious
from Table 4.1, which gives relative responses to nitrogen compounds converted
to the mass content of nitrogen.
An alternative theory that explains the long-term stability of the Perkin-
Elmer NPD was eleborated by Ol~h et al. 6 . The mechanism for a nitrogen compound
is as follows:

k'
N-comp·(vapour) ~
+ N-comp·(ads) (adsorption) (4.5)
k

N-comp·(ads.) ~
CN(ads) (degradation) (4.6)

CN(ads) ~ products (desorption) (4.7)

Rb + CN(ads) Rb+ +
~
CN(adS) (ionization) (4.8)

CN(adS) +
CN(gas) (emission) (4.9)

Rb+ + e + Rb (recombination) (4.10)

The essential parts of the reaction sequence are the following:


4.5 and 4.6: non-ionic, voltage-independent reaction, rate-determinating step
at high voltages and limiting current;
4.7: process concurrent with 4.8, branching to non-ionic desorption and de-
gradation, at least to CO 2 and N2 ;
4.8 and 4.9: ionization and ion emission, activated and exponentially becom-
ing a voltage-dependent process, rate-determining step at low voltages;
4.10: Rb atom recombination on the bead surface; the electrons originate from
the platinum wire, migrating through the glass phase.
The charge carriers are electrons and emitted negative ions. The former migrate
inside the glass from the platinum to the surface and the latter move from the
glass surface through the gas phase to the collector. According to this theory,
the ionization process is catalytic, the Rb atoms being only surface catalysts
of the electron transfer. The alkali glass is not a source, because the Rb atoms
do not leave the glass surface.
The idea of electrically heating the alkali metal source originated in order
to make the source temperature independent of the gas flow-rates and to provide
for setting optimum flow-rates of these gases. As already mentioned in Chapter 3,
en
00
TABLE 4.1

RELATIVE RESPONSES OF NPD AND FID TO DIFFERENT TYPES OF NITROGEN COMPOUNDS


(From ref. 5 with permission.)

Compound Blended weight ratio Relative area Relative response factor,


response ratio weight/area
Nitrogen Carbon NPD FID NPD FID

Pyridine 1.278 0.7988 1.617 0.7122 0.7904 1.122


2,6-Dimethylpyridine 0.9494 0.8308 1.120 0.8450 0.8477 0.9832
4-Methylpyridine 1.076 0.8082 1.564 0.7534 0.6880 1.073
2,4,6-Trimethylpyridine 0.8720 0.8759 1.109 0.9438 0.7863 0.9281
4-tert.-Butylpyridine 0.7551 0.8515 1.145 0.8995 0.6595 0.9466
Quinoline 1.008 1.133 1.101 1.037 0.9155 1.093
2-Methylquinoline 0.9857 1.235 1.054 1.2S2 0.9352 0.9864
6-Methylquinoline 0.7154 0.9154 0.8450 0.8880 0.8656 1.031
2,6-Dimethylquinoline 0.6445 0.8853 0.6985 0.9255 0.9227 0.9566
2,4-Dimethylquinoline 0.9146 1.257 0.9732 1. 318 0.9398 0.9537
Indole 1.000 1.000 1.000 1.000 1.000 1.000
3-Methylindole 1.017 1.145 0.9628 1.184 1.056 0.9671
2,3-Dimethylindole 0.7061 0.8815 0.6260 0.8730 1.128 1.010
N-Ethylcarbazole 0.5466 0.9567 0.5942 0.9525 0.9199 1.004
N-r~thylcarbazole 0.6129 0.9981 0.6596 0.9257 0.9292 1.078
5,6-Benzoquinoline 0.7819 1.272 0.8492 1. 284 0.9207 0.9907
l,2,3,4-Tetrahydrocarbazole 0.6224 0.9360 0.5204 0.9078 1. 196 1.031
6-Methyl-1.2,3.4-tetrahydrocarbazole 0.5481 0.8928 0.4831 0.8889 1.135 1.004
Carbazole 0.7472 1.120 0.6621 1.072 1.128 1.045
69

the background ionization current increases with increasing flame temperature


owing to the higher temperature of the source and the greater ionization of the
alkali metal. When increasing the temperature of the source by electrical heating,
without changes in the flow-rates of the gases, the background ionization current
increases exponentially depending on the vapour pressure. The response then also
increases with increasing background ionization current 4 ,7,8.
If the background ionization current of a three-electrode detector is increased
by increasing the hydrogen flow-rate with a constant heating current, the responses
to nitrogen compounds also increase, but they do so faster than when the back-
ground ionization current is changed by heating it electrically at a constant
hydrogen flow-rate. Even if the background current is kept constant (by changing
the heating current), the response increases with increasing hydrogen flow-rate.
This means that the variations in response caused by the hydrogen flow-rate can-
not be compensated for by changing the heating current so as to maintain a con-
stant background current. The same is also true for the flow-rates of the carrier
gas and air7. However, on changing the heating current so as to keep the back-
ground ionization current constant, the decrease in response with time can be
compensated for (see section 4.7).
In contrast to Lubkowitz et al. 7, who reported an increased sensitivity when
the hydrogen flow-rate was increased in the NP mode with a constant background
current (adjusted by changing the heating current), a decrease in sensitivity was
also found 1,2,8.
In the P mode, the detector can also respond to nitrogen compounds: about
2 • 10- 3 C/g of azobenzene, this sensitivity being essentially independent of
the hydrogen flow-rate. By reducing the nitrogen flow-rate the response can be
increased (relative to the optimum for phosphorus compounds) to a value as high
as 1.3 . 10- 2 C/g 4. At a constant heating current, this optimum is about
74 ml/min for phosphorus compounds 4• In the NP mode, the dependence of the sen-
sitivity on the flow-rate of the carrier gas reveals a maximum at low flow-rates,
about 7 ml/min (similarly to the background current 7). The replacement of nitro-
gen with helium leads to changes in the background current owing to a difference
in the thermal conductivity of the two gases 9 . The dependence of the sensitivity
on the air flow-rate shows a maximum at flow-rates of about 200 ml/min (the same
as for the background current) in both the P mode 4 and the NP mode 7.
The basic parameters of the detector are listed in Table 4.2. Table 4.3 gives
the basic parameters of a pure rubidium-quartz source of the same detector con-
struction prepared from ground quartz mixed with rubidium hydroxide 8.
TABLE 4.2

CHARACTERISTIC OPERATION SPECIFICATIONS OF THE PERKIN-ELMER DETECTOR IN THE THREE MODES OF OPERATION
(Reproduced from ref. 2 with permi ss ion. )

Parameter FI mode P mode NP mode


P N

Sensitivity 5.10- 3 C/g C 1 C/g P 5 C/g P 0.5 C/g N


Minimum detectable mass rate 3.10 12 g C/sec 5.10- 14 9 P/sec 1.10 14 g P/sec 1.10- 13 g N/sec
Linearity 2.10 7 10 5 10
5
105
Selectivity 10 6 g C/g P 1.25.10 5 g C/g P 2.5.10 4 g C/g N
Noise ca. 5.10- 14 A ca. 7.10- 14 A
TABLE 4.3

PERFORMANCE CHARACTERISTICS OF RUBIDIUM/QUARTZ BEADS


(Reproduced from ref. 8 with permission.)

Compound Atom X Sensitivity Selectivity mmin Noise


(C/g X) (g X/g C) (g X/sec) (A)
2.4% Rb 7.6% Rb 19.6% Rb 2.4% Rb 7.6% Rb 19.6% Rb (19.6% Rb)

Diethylphenyl 6.10 4
phosphorothioate P 3.1 4.1 9.2 1.8.10 5 6.1.10 5 1.1.10- 15 1.1.10- 14
Ch 1orpyri fos P 3.6 3.3 6.4 7.10 4 1. 4.10 5 5.4.10 5 1. 7.10- 15
Atrazine N 0.4 0.2 0.3 7.10 3 9.10 3 2.5.10 4 3.4.10- 14
Lindane Cl 2.9.10- 3 1.2.10- 3 2.2.10- 3 57 50 1.9.10 2 4.4.10- 12
Hexadecane C 5.2.10- 5 2.3.10- 5 1.2.10- 5 1 1 1 7.2.10- 10
72

4.3 THE HEWLETT-PACKARD DETECTOR

The NPD manufactured by Hewlett-Packard uses a ceramic cylinder coated with


an alkali metal salt activator as a source of alkali metal. The source is elec-
trically heated. A voltage of -240 V is applied to the cylindrical collector
electrode enclosing the source (shown schematically in Fig. 4.3). A low-tem-
perature plasma is generated around the source 10 . The basic parameters of this
detector are given in Table 4.3. The use of an 8% mixture of hydrogen in helium,
instead of pure hydrogen, improves the sensitivity of the detector. The detector
also responds to those nitrogen compounds which do not contain HCN bonding (cf.,
Fig. 4.2). The response mechanism described by Kolb and Bischoff 1 for the
Perkin-Elmer detector does not apply to the Hewlett-Packard detector, because
the detector potential is negative. In addition, it responds fairly well to
barbiturates and pesticides which contain vicinal carbonyl groups and no HeN
bonding lO •

POLARIZING VOL TAG E

TO RECORDER

lOj r::J :lECTRODES


AIR- "'-
(1 "ALKALI METAL SOURCE
HYDROGEN I JET

COLUMN EFFWENT

Fig. 4.3. Schematic diagram of the Hewlett-Packard detector. (From ref. 11.)

TABLE 4.4

PERFORMANCE CHARACTERISTICS OF THE HEWLETT-PACKARD DETECTOR (From ref. 10.)

Parameter Compound
Nitrogen Phosphorus
t1inimum detectable mass rate:
8% H2-He <1.10- 13 g N/sec P/sec
H <4.10- 13 g N/sec P/sec
Sel~ctivity (g X/g C) 3.4.10~
Linear dynamic range >10
Background current (A) 1.6.10- 11
Flow-rate of hydrogen (ml/min): 1.0-5.0
8% H2-He 15 -60
Air 30 -100
73

4.4 THE TRACOR OETECTOR

The source of alkali metal in the Tracor detector is a mixture of alkali metal
salts in a silica gel matrix 12 • The detector is provided with a quartz jet, above
which the alkali metal source and the collector electrode are positioned. The
heating element for the alkali metal source is a short loop of platinum wire,
which constitutes one arm of the Wheatstone bridge 13 • As the resistance of
platinum is directly proportional to its temperature, the voltage drop across
the source can be related to the power required to heat the source in order to
maintain a fixed resistance (temperature) 14. Theoretically, the rapid response
of the alkali metal source heater makes the source temperature independent of
the carrier and plasma gas flow-rates. In fact, small variations in the source
temperature occur, however13.

TABLE 4.5

PERFORMANCE CHARACTERISTICS OF THE TRACOR DETECTOR


(From ref. 12.)

Parameter Compound
Nitrogen Phosphorus

Minimum detectable mass rate


Selectivity (g X/g C)
Linear dynamic range

The basic parameters of this detector are listed in Table 4.5. The sensitivity
to phosphorus and nitrogen compounds, hydrocarbons, the background current and
noise increase as the source setting (temperature) is increased. The sensitivity
to nitrogen compounds and hydrocarbons decreases with increasing hydrogen flow-
rate (1-7 ml/min) at a constant background current, whereas the sensitivity
to phosphorus compounds increases. Therefore, nitrogen compounds can be distin-
guished from phosphorus compounds at different hydrogen flow-rates or due to a
change in the negative polarizing voltage. A decrease in voltage reduces the
phosphorus response by 30% while maintaining the normal response for nitrogen.
74

4.5 THE VARIAN DETECTOR

Fig 4.4 is a schematic diagram of the detector manufactured by Varian. A


ceramic alkali metal source containing an alkali metal salt in its entire volume
is placed above the detector jet, through whi ch the effl uent together wi th
hydrogen leaves the chromatographic column. The source is heated electrically
and a cylindrical collector electrode is situated around it •

~~~~~~~~~~:i~~~~t-

_ _~SCREEN
-I~I---COLLECTOR WITH

ERAMICBOTTOM
BEAD WITH
HEATER COIL

FLAME TIP

l-----DETECTOR BASE

1 1 1 4 - - - - GC COLUMN

CARRIER GAS

Fig. 4.4. Schematic cross-section of the Varian detector. (Reproduced from


ref. 16 with permission.)
TABLE 4.6

PERFORMANCE CHARACTERISTICS OF THE VARIAN DETECTOR


(From ref. 16.)

Parameter Compound

Nitrogen Phosphorus

Minimum detectable mass rate 5'10~14 9 P/sec


Selectivity (g X/g C) 2·10
Linear dynamic range 10 4
Background current (A) 4.10- 12
Noise (A) 1_10- 14
Flow rates:
Hydrogen (ml/min) 4.5
Air (ml/min) 200
75

The sensitivity of the detector to nitrogen and phosphorus compounds is about


five orders of magnitude higher than that to other compounds (see Table 4.6).
\~ith a low negative voltage applied to the alkali metal source, the response
(negative ions) is high at a low background ionization current. On passing from
a negative to a positive voltage the polarity of the response is changed, and
the response is relatively low at a relatively high ionization current 15 ,16 (see
Fig. 4.5). A voltage of about -4 V is applied to operate the detector 16 •
The temperature of the source is the main parameter affecting the response of
the detector (Fig. 4.6). The detector response drastically increases with in-
creasing heating current of the source, with both phosphorus and nitrogen com-
pounds. However, the background ionization current and, consequently, the detec-

z
UJ
Z
~
z~ -22V
~~ z 12
-6J, • 1(f A
OUJ
~~ a..
U
UJ
za..
o
~
a..
UJ
I

-1.85V
12
+ 64·10· A
-1.80 V
_12
+64'10 A

Fig. 4.5. Chromatograms of detector test sample at different settings of bead bias
voltage. IB = background current. Column, glass, 200 cm x 2 mm 1.0., packed with
5% OV-IOI on Gas-Chrom W. (Reproduced from ref. 17 with permission.)
76

1D

0.9

0.8

0.7

0.6

0.5

0.4

t- 0.3
Z
UJ
c::: 0.2
c:::
:J
U
0.1
Z
Q 0
Cl 2.50 2.70 2.80 2.90
UJ
N
BEAD CURRENT (A)
~ toO
~ 0.80
c:::
0.60
~ 0.40

0.20

0.10
0.08
0.05
0.04

0.02
0.115 0.120 0.125 0.130 0.135 0.140' 0.145
(BEAD CURRENTr 2 (A- 2)

Fig. 4.6. Variations of the background current (I B), peak height of azobenzene
(IN) and peak height of mal~thion (Ip) as a functlon of bead heating current
and (bead heating current)- • In tne top graph the ion current scale in linear
and in the bottom graph it is logarithmic. (From ref. 15.)

tor noise increase more rapidly, so that the minimum detectable mass rate in-
creases. It is recommended, therefore, that one should work at a source tempera-
ture that is low enough to be still suitable for the decomposition of the sample;
in fact, a background ionization current of 4'10- 12 A is appliedl~.
The hydrogen flow-rate is low and, therefore, it exerts a negligible influence
on the temperature of the source surface 15 ,16. However, changes in the hydrogen
flow-rate can result in substantial changes in the radical concentrations in the
boundary layer of the gases (see later); hence the hydrogen flow-rate strongly
affects the response level and the selectivity of the detector (Fig. 4.7). At
higher hydrogen flow-rates, the detector response to hydrocarbons increases
whereas that to phosphorus and nitrogen compounds decreases. At increased air
flow-rates (100-200 ml/min), the background ionization current and, therefore,
77

the detector response decrease. If nitrogen is used as the carrier gas, the
decrease occurs more rapidly than with helium. A similar dependence also holds
for the flow-rates of the carrier gas (10-60 ml/min); the decrease in response
with increasing flow-rate is faster in helium. However, if the background ioniza-
tion current is kept constant by changing the heating current of the alkali metal
source, the response is constant at varying flow-rates of air and carrier gas.

1.0

0.8

0.6
r
Z 0.4
UJ
0::
0:: 0.2
~
u
z 0
0
1.0
0
UJ
N 0.8
He
::::i
<!
L 0.6 ,,
0::
0
Z 0.4
,,
,,
;f
I

"
/ ,,
0.2 p"
"
..... ~' "
0
0 2 4 6 10
H2 FLOW (mllminl

Fig. 4.7. Variations in the background current (IB)' peak height of azobenzene
(IN), peak height of malathion (Ip) and hydrocarbon response current (IC) as a
function of H2 flow-rate for He and NZ carrier gases at 25 ml/min. Bead heating
current held constant. (From ref. 15.)

The sample leaving the chromatographic column impinges on the hot surface of
the alkali metal source (700-900 oC), where it decomposes. A boundary layer of
high-temperature gases is produced around the hot source. The layer consists of
radicals, similarly as in a flame (H, OH, 0), which play an important role in the
decomposition of the compounds and in subsequent ionization. In this detector,
the ionization mechanism is ascribed to surface ionization processes 15 ,16. Nega-
tive ions are emitted from the surface of the alkali metal source. The work
function of the beads is about 3.4 eV, whereas it drops to some 2.3 eV in the
presence of nitrogen and phosphorus compounds (as calculated from the dependence
78

of the response and the background ionization current on the heating current of
the source and on the detector temperature). Hence the detector response to
nitrogen and phosphorus compounds is due to the reduction of the work function
of the bead. This reduction is attributed to the formation of decomposition
products that are highly electronegative (probably CN, N0 2 and P0 2 ) and to the
subsequent generation of gas-phase negative ions owing to extraction of elec-
trons from the surface of the heated source. Both the background ionization
current and the detector response depend on the temperature of the source sur-
face (see Fig. 4.6) in a way that can be expected for thermionic emission 15 •

4.6 THE DETECTOR ENGINEERING TECHNOLOGY DETECTOR

In 1982, Patterson et al. 17 described another detector, developed for the


Spectra-Physics gas chromatograph. A significant difference between this detec-
tor and the previous version 15 ,16 is that the source is in the form of a cylin-
drical collector electrode. This detector has been further modified 18 ,19 so that
the alkali metal source possesses a separate non-alkali metal/ceramic sublayer
and an alkali metal/ceramic surface layer (Fig. 4.8). The sublayer protects the
metallic heating wire in the source from chemical attack by corrosive alkali
metal atoms or by corrosive decomposition products of the sample compounds. The
detector is provided with a built-in temperature-sensing element, which is used
in an electronic control circuit that varies the heating current so as to main-
tain a constant source temperature.
Two different surface coatings were described for the detector 17 ,19. One of
them contains a low concentration of a caesium compound that is used for opera-
tion in dilute hydrogen/air environment. The other coating contains a much
higher concentration of the caesium compound and nitrogen is the only gas sup-
plied to the detector. The mechanism of detection is thought to be the same as
already described in section 4.5 for the Varian detector, i.e., a surface ioniza-
tion process that can be described schematically as follows:

Sample + Hot surface + Chemically + Electronegative decomposition (4.11)


reactive products
gases

Electronegative species + Hot surface + Negative ions

NP selectivity is obtained from the very hot surface (600-800 0 C) of moderate


work function (so-called TID-2-H /air). In the TID-I-N 2 mode, the low-work-fonc-
2
tion source operates at a moderate temperature (400-600 0 C). In this instance,
the hydrogen (gas 1) and air (gas 2) flows to the detector are replaced with
79

GAS 2
/n"(/ rfi\
I
I
I
.,. ....

I
I
I
SA~PLE
GAS 1
-'" '\

I
I
I
SAMPLE CONDUIT

Fig. 4.8. Schematic diagram of the TID detector. (From ref. 19.)

nitrogen flows (both gas 1 and gas 2). The gaseous boundary layer of the source
is no longer chemically reactive. Therefore, if samples are to decompose into
electronegative products, the electronegative functional groups must originate
within the sample itself l7 , i.e., the detector is selective to certain compounds
containing electronegative functional groups:

Sample + Hot surface + Electronegative decomposition products


(4.12)
Electronegative species + Hot surface ~ Negative ions

The source is biased at a negative potential (-15 V for the Spectra-Physics,


-5 V for the TID-2-H 2/air, and -15 or -45 V for the TID-1-N 2 ).
For the TID-2-H 2/air, the basic parameters of the detector at flow-rates of
hydrogen = 3 ml/min and air = 60 ml/min are as follows 19 : sensitivity, 0.35 C/g
of N (azobenzene) and 1.1 C/g of P (malathion); minimum detectability, 1,10- 13 g
of N/sec (azobenzene); linear dynamic range, 10 5 ; N:C selectivity, 2.10 4 ; P:C
selectivity, 4'10 4. Also with this type of detector the above basic parameters
depend on the temperature of the source (heating current) and on the hydrogen
flow-rate. I~ith i ncreas i ng source temperature the background current ; ncreases,
the selectivity slightly increases and the minimum detectability decreases. The
dependence of the selectivity and minimum detectability for nitrogen compounds
on the hydrogen flow-rate displays an optimum value in the vicinity of flow-rates
80

FID TID H2 =6 ml/min


1024 • 1(j"12 A

N.P

c
c
N

N N.P P

TID H2=3 ml/min


128' 1012A
TID H2 = 9 ml/min
11
512 '10 A
~
C
a...
Z· a...
z
0

N,P
P
N

I
0 2 4 6 MIN 0 2 4 6 MIN

Fig. 4.9. Comparison of FlO and Spectra-Physics TIO chromatograms of the detector
test sample. (Reproduced from ref. 17 with permission.)

of 5-6 ml/min. At a hydrogen flow-rate above 5.5 ml/min, the background current
increases very rapidly and the detector loses its selectivity and starts to behave
like an Fl0 20 • Fig. 4.9 shows chromatograms of the FlO and those of the above de-
tector at different hydrogen flow-rates. As can be seen, the phosphorus compounds
behave in a different way to nitrogen compounds 17 when the hydrogen flow-rate is
changed. At a hydrogen flow-rate of 6 ml /mi n the sens iti vity for phosphorus com-
pounds is enhanced as compared with that at a flow-rate of 3 ml/min, but that for
nitrogen compounds is suppressed. The detector becomes much less selective at
hydrogen flow-rates of 9 ml/min and above.
The basic parameters of the TlO-1-N 2 at a total nitrogen flow-rate of 50-
200 ml/min are as follows 19 : noise, 1'10- 14 _2'10- 14 Ai sensitivity, 0.6 C/g of
81

2,4-dinitrotoluene; minimum detectability, 1.10-13 g of 2,4-dinitrotoluene/sec;


selectivity, >10B (methylparathion/C 15 ); linear dynamic range, 10 3_10 4• The
minimum detectability is again dependent on the heating current, detector tem-
perature and other instrumental parameters 21 Peak tailing was observed when
using a TID-1-N 2 (heating current 3.2-3.4 A).

TID (L-Cs)
H2/AIR
128 .1O-12 A

Cl. ii:_
~

TID(H-Cs)
N2
16·1Q-11A

Fig. 4.10. Comparison of chromatograms of the test sample for the two different
modes of TID operation. L-Cs, low-concentration Cs source; H-Cs, hiqh-concentra-
tion Cs source. (Reproduced from ref. 17 with permission.)

The sensitivity of the above detector depends not only on the presence of an
electronegative atom or group in the molecule of the compound (Fig. 4.10), but
also on the structure of this molecule. This is obvious from Fig. 4.11, which
shows the chromatograms obtained from the analysis of base neutrals that are of
concern as water pollutants. The catalytic flame ionization detector (CFID) is
a detector that resembles the TID-1 and TID-2 in design, but its source consists
of a nickel-ceramic composition for both the sub-surface and surface layers. The
hydrogen and air flow-rates to the detector are 25 and 100 ml/min, respectively.
The CFID provides universal responses to most organic compounds and is similar
to but not identical with an FID lB ,19. The TID-l-N 2 mode yields a selective re-
sponse only to 2,6-nitrotoluene and 3,3'-dichlorobenzidine. If the detector gas
environment of the TID-l source is changed from nitrogen to oxygen, the TID-l-0 2
mode yields an enhanced relative response to chloro compounds and a diminished
relative response to nitro compounds. The TID-2-H 2/air mode responds to all nitro-
82

BASE NEUTRALS 200 ng


CFID 16'10-10 A
a: a:
UJ UJ
:J: :J:
....
UJ
....
UJ
UJ
Z
::;
,.. UJ
:J:
:J: ....
....
UJ ,..
...J
Z
..:
o x a:
a: UJ o
g Z
UJ
:I: ~

:J:
UJ
N ,..
...J ...J
U.

~ Z
UJ
:I:
.... E
~
CD
o
a:
.... II)
a;
Z

I J
TID-1-N2
8' 10- 9 A

J_'-----------'~~
I ~~ TI .D1~1_~OO~
~~l---------.J!~' - TID -2-H2/AIR
16 .10- 9 A

Fig. 4.11. Chromatograms of base neutrals. (From ref. 19.)

gen compounds. Fig. 4.12 depicts other modes of the TID-I. Comparison of the
response to 4-nitrophenol and 2-nitrophenol illustrates a significantly greater
TID-I-N 2 sensitivity for the isomer with the nitro group in the 4-position. If
the source is operated in gas environment composed of approximately equal flows
of nitrogen and air. the sensitivity to certain compounds is suppressed whereas
that to others is enhanced. If both detector gases 1 and 2 are air. responses
are obtained for all of the chloro- and nitrophenols. with the dominant response
for dinitro compounds (FiJ. 4.13). The sensitivities to 2- and 4-nitrophenol are
commensurate. For nitro 'aromatics the TIO-I-N 2 provides a negligible response to
nitrobenzene and a selective high response to 2,4- and 2,6-isomers of dinitro-
toluenes with about a tw;'~e as high response to the 2,4_isomer 1B ,19 The sen-
sitivity for nitrated polycyclic aromatic hydrocarbons is also very different.
It differs from solute to solute by as much as a factor of 100 (ref. 21).
83

CFI D 3· 10-10 A

...J
o
z
UJ
:t
"-
o
II:
....

TI D-l-N
3 '10- 8 A

TID-l-N/AIR
L
3 ·10-s A

Fig. 4.12. Chromatograms of phenols. (From ref. 19.)

4.7 THE CHEMI-IONIZATION DETECTOR

A detector that yields a selective response in a nitrogen environment when an


alkali metal is present was described by Scolnick 22 . In this detector (called the
chemi-ionization detector), caesium bromide in the gas phase is brought into the
reaction space of the detector. The detector gives a selective response to phos-
phorus compounds and this response is 20-60 times smaller than that of the
AFID. A schematic diagram of the device is presented in Fig. 4.14. The effluent
from the column enters the reaction zone of the detector (temperature 800-850 0 C),
where it mixes with nitrogen saturated with CsBr vapour in the saturation zone
at about 500 0 C. The tubular cathode is earthed and the collector electrode is
biased at +50 V. The results of the measurements showed that the ionization re-
actions between the alkali metal salt vapour and the phosphorus compounds can
occur in the gas phase and in the absence of combustion products or hydrogen and
oxygen radicals.
84

TIO-l-AIR
8 • 10-9 A

-'
0
-'
0
Z
'"
UJ
It:
UJ
:I: 'r
IL
0
I'
0
8• 10-1OA It:
.... It: 0
-'
t: z
-'
0 Z z UJ
;:; :I:
~
Z IL
UJ
:I: -'
.!t 0
It:
IL 0 N -i 0
0
-'
0
It:
0 '"
UJ
It:
-'
:I:
<..)
Z -' <..)
UJ :I: ....'z"
.,
..J
-'
0
Z
J:
IL
<..)
iE E
I UJ
0
Z
UJ 0 0 IL UJ
:I: It: It: :I:
IL
0
0
-'
:I:
'"N
-t' 0
-'
IL
0
It:
It:
0 <..)
:I:
y.., ....
~ ~
-'
:I:
Y
N N'

Fig. 4.13. Chromatograms of phenols. Both TID gases 1 and 2 are air. (From ref.
19. )

4.8 DETECTOR LIFE AND REPRODUCIBILITY OF RESPONSE

As has already been mentioned in Chapter 3, the popularity of FASDs over AFIDs
is due mainly to the long service life of the alkali metal source in the FASD.
The life of the FASD has been reported to amount to 1000 h of operation.
A high long-term stability has been reported for the Perkin-Elmer three-
electrode detector and is ascribed to the following mechanism 1• Transport of the
alkali metal cation to the gas phase from a source having a negative potential
is impossible, and this transport is also unlikely with the silicate of an alkali
metal. Hence only the neutral atom of the alkali metal, which can be formed by
the alkali metal ion accepting an electron (at the temperature applied the glass
behaves like an electrolyte), can be transported to the gas phase. Owing to ioni-
zation in the flame, alkali metal ions are produced, which are immediately col-
lected by the negatively charged source, neutralized and re-evaporated (see Fig.
4.15). The relative standard deviation of the response for methamphetamine (NP
mode) was 1.5% over 30 days9.
Lubkowitz et al. 7 showed, however, that the service life of the above detector
in the NP mode, particularly with background ionization currents exceeding
5.10- 12 A, is markedly shorter. When the source is heated electrically, the back-
85

INSULATING
MATERIAL BIAS
VOLTAGE

....=1---11-+-1111-- E LE CTROMETER

VYCOR
TUBING

CsBr+
GLASS BEADS - - -...

COLUMN
EFFLUENT

Fig. 4.14. Schematic of the chemi-ionization detector. (Reproduced from ref. 22


with permission.)

Flame space

Surface of
the bead

Fig. 4.15. Alkali metal recycling mechanism in the Perkin-Elmer detector. e- =


electron; A' = alkali metal atom; A* = excited alkali metal atom; A+ = alkali
metal ion; R' = eN or P02' (Reproduced from ref. 2 with permission.)

ground ionization current decreases from 51'10- 12 to 0.8'10- 12 A within 16 days,


and higher heating currents are required to attain the same background ionization
current. Table 4.7(A) reveals the rapid decrease in response during the first
6 h, simultaneously with the reduction in the background ionization current when
the heating current is kept constant. By maintaining the background ionization
current constant by increasing the heating current, a relative standard deviation
of 0.94% can be obtained [Table 4.7(8)J.
86

TABLE 4.7

REPRODUCIBILITY STUDIES OF PEAK AREAS


(From ref. 7.)

Time elapsed Heati ng current Bead current Area response *


(h) (dial setting) (pA) (arbitary units)

A. Constant heating current:


o 6.600 14.3 7320
6 6.600 11.3 5823
24 6.600 7.20 3462
48 6.600 4.17 1685
B. Constant bead current:
0"* 7.236 14.3 4622
1 7.242 14.3 4606
2 7.264 14.3 4718
3 7.278 14.3 4659
4 7.298 14.3 4690
5 7.320 14.3 4704
6 7.345 14.3 4705
Av~rage 4672
R.S.D.*** 0.94%

*Constant azobenzene injection of 11.8 ng.


"*Initial setting not changed over a period of 48 h.
***R.S.D. = relative standard deviation.

A smaller decrease in sensitivity with time (to about a quarter after 30


days) was found by Schulte and Shive 23 . The inconsistency of the determined
servi ce 1i fe wioth the mechani sm of the a1ka 1i metal ci rcul ati on has been ascri bed
to the fact 4 that ageing o~ the source is caused by changes in the properties
of the platinum wire carrying the silicate. The reducing flame that burns around
the alkali metal source induces embrit°tlement and cracking of the platinum wire.
In contrast to the NP mode, the flame burns at the jet in the P mode and there
is no reducing atmosphere. The response does not change during 7 days4. The lowest
reproducibility of the response was found for sources with high boron and sodium
contents 24 •
The stability of the Hewlett-Packard detector has been reported to be excel-
lent 10 ; the standard deviation of 199 analyses of samples containing nanogram
amounts of amphetamine and methamphetamine was less than 5%.
The service life of the ceramic alkali metal source used by Varian is several
thousand hours. However, the detector sensitivity decreases during the operation
of this source and changes in the selectivity of the response occur. This de-
87

crease can be offset by adjusting the heating current of the source 15 (Fig.
4.16). Detectors with sources consisting of an inert ceramic core with a surface
coating of an alkali metal-ceramic activating material also exhibit decreases in
background current, sensitivity and selectivity with operating time 20 • The de-
crease in sensitivity and selectivity can be restored to some extent by periodic
adjustments of the source heating current. A decay of background current and
sensitivity was also observed with the Tracor detector 13 • Therefore, it is ob-
vious that almost all commercial FASDs (NP mode) also suffer some decrease in
sensitivity with operating time.

1.0
w
VI
Z
~ 5
VI
W
0::

w
Z 0
1 - - - - - - - - ACETONE SOLVENT -----_+_ ACETONE
AND
CHLOROFORM
50:50
W
U.
U.
~ 1.0
o
W
N
:; 5
<{
~
0::
o
Z -----+--METHANOL EXTRACT OF COLA DRINK-
o 50:50

o 200 400 600 BOO 1000 1200 1400 1600 lBOO 2000
OPERATING TIME (H)

Fig. 4.16. Change of detector response as a function of bead operating time.


Discontinuous steps in data correspond to increases in bead heating current to
the values indicated. (Reproduced from ref. 16 with permission.)

4.9 DETECTORS FOR HALOGEN COMPOUNDS

All the types of FASDs described in the previous sections, except for the
TID-2-N2 (section 4.6), are selective for phosphorus and nitrogen compounds.
However, as this chdpter deals with the FASD it should also include the surface
ionization detector, which is selective for halogen compounds and is also called
the thermo-ionization detector in the literature. In 1961, Cremer et al. 25 em-
ployed a surface ionization detector as a halogen-sensitive device in the search
for leakages 26 and as a detection device in the gas chromatography of halogen
88

To heater

To anode

Glass envelope

Silicone-rub
sleeve

IIII!I--- Brass adapter

From column

Fig. 4.17. Halogen-sensitive detector. (From ref. 27.)

compounds. The detector (Fig. 4.17) consists of an anode containing an alkali


metal (ceramic body, alkali metal-coated platinum) and of a cathode usually in
the shape of a cylinder enclosing the anode. The anode is electrically heated
(800 oC). If the electron affinity of the anode material is higher than the ioni-
zation potential of the alkali metal, emission of positive ions from the heated
anode occurs 28 ,29. The ion emission increases in the presence of halogens. This
phenomenon is attributed 30 to the formation of a K-O complex on the surface of
the anode metal that splits off K+ at higher temperatures. The probability of
cleavage of the complex increases in the presence of halogens. A Cl-K-O complex
is produced that considerably alters the work function.
The sensitivity of the detector depends on the heating current of the anode
and on the detector temperature 25 ,31. The response to halogen compounds increases
with increasing heating current 27 ,31-33 (Table 4.8) and, at the same time, the
response to hydrocarbons 27 , the background ionization current and the noise also
increase 31 ,33. The response also increases with increasing voltage of the anode,
and the detector noise also increases again 27 • The response increases with in-
creasing carrier gas flow-rate (40-120 ml/min)34. The response is higher for
chlorine compounds than for other halogen compounds 25 , and is about four orders
of magnitude higher than for halogen-free compounds 25 ,32. The ratios of the molar
89

TABLE 4.8

CHANGE IN RESPONSE OF A PLATINUM THERMIONIC DIODE TO SOME PESTICIDE COMPOUNDS


WITH INCREASING INCREMENTS OF FILAMENT HEATER CURRENT
(From ref. 33.)

Pesticide Amount Response- as peak height (mm)


injected
(ng) lo7A * 1.8 A* 1.9 A*

Lindane 0.18 10 30 91
Heptachlor 0.38 15 32 116
Aldrin 0.36 7 17 72
Tel odri n 0.36 11 26 97
Kelthane 0.11 7 14 75
Heptachlor epoxide 0.36 11 22 95
p,p'-Dichlorodichlorophenylethylene 0.36 4 9 40
Dieldrin 0.36 4 9 39
Methoxychlor 9.0 13 30 90
Parathion 4.0 0 0 0
8.0 0 0 0
Melathion 8.0 4 8
16.0 13
Trithi on 4.0 7 17 41
8.0 81

*Filament current.

responses of fluorine, chlorine, bromine and iodine compounds are 0.9 : 1 : 0.8 :
0.1 31 . The ratios of the molar responses of mono-, di-, tri- and tetrachloro-
methane are 1 : 2 : 3 : 431 and those of mono- and trichlorobenzene are 1 : 2 to
1 : 2.2 32 • The response decreases with operating time and the ratio of the molar
responses of the individual halogens also changes (but that for chlorine com-
pounds is always the highest)31. The detection limit is 1.10- 10 g of chlorine
compounds 27 ,33 and the linear dynamic range of the response covers 4-5 orders
of magnitude 32 ,34,35. The distance between the electrodes is critical (usually
it is about 1.5 mm). loJith an increase in distance above 1.5 mm the response de-
creases rapi dly and becomes zero at an electrode di stance of 4 mm 33 .
Cremer and co-workers 32 ,35 applied pre-combustion of the sample prior to its
introduction into the space of the detector proper. Pre-combustion overcomes the
problems related to the variations in the baseline in the presence of excess of
solvent, enhances the reproducibility of the response and correlates the ratio
of the molar responses of the halogen compounds concerned with the number of
halogen atoms in the molecule of the compound.
90

REFERENCES

1 B. Kolb and J. Bischoff, J. Ch~omatogr. Sai •• 12 (1974) 625.


2 B. Kolb, 11. Auer and P. Pospisil, J. Chromatogr. Sai., 15 (1977) 53.
3 B. Kolb, M. Auer and P. Pospisil, J. Chromatogr., 134 (1977) 65.
4 B.P. Semonian, J.A. Lubkowitz and L.B. Rogers, J. Ch~togl'., 151 (1978) 1.
5 O.K. Albert, Anal. Chern., 50 (1978) 1822.
6 K. OUh, A. Szoke and Z. Vaita. J. Chromatogr. Sai., 17 (1979) 497.
7 J.A. Lubkowitz, J.L. Glajch, B.P. Semonian and L.B. Rogers, J. Chromatogr.,
133 (1977) 37.
8 R. Greenhalgh, J. MUller and W.A. Aue. J. Chromatogr. Sai., 16 (1978) 8.
9 M.J. Hartigan, J.E. Purcell, M. Novotny, M.L. McConnell and M.L. Lee, J.
Chromatogr •• 99 (1974) 339.
10 C.A. Burgett. D.H. Smith and H.B. Bente, J. Chromatogr., 134 (1977) 57.
11 H.B. Bente, Appliaation Note ANGC 2-77, Hewlett-Packard, Avondale. PA, 1977.
12 Traaol' MOdel 702 N-P Deteator, Austin. TX. 1977.
13 B.J. Ehrlich, Ind. Ree. Dev •• April (1980) 107.
14 R.C. Hall, CRC Crit. Rev. Anal. Chem., 8 (1978) 323.
15 P.L. Patterson, J. Chromatogr •• 167 (1978) 381. .
16 P.L. Patterson and R.L. Howe, J. Chromatog~. Sai., 16 (1978) 275.
17 P.L. Patterson. R.A. Gatten and C. Ontiveros, J. Chromatogr. sai., 20 (1982)
97.
18 P.L. Patterson, Chromatographia. 16 (1982) 107.
19 DET Deteatore, Detector Engineering Technology, Walnut Creek, CA.
20 A. Nohl, Speat.ra-Phyeiae Chromatogr. Rev •• 9, No.1 (1983) 11.
21 C.M. White, A. Robbatt, Jr •• and R.M. Hoes, Anal. Chem •• 56 (1984) 232.
22 M. Scolnick. J. Chromatogr. Sai •• 8 (1970) 462.
23 B.K. Schulte and L.W. Shive. Anal. Chern., 54 (1982) 2392.
24 J.A. Lubkowitz. B.P. Semonian, J. Galobardes and L.B. Rogers, Anal. Chern.,
50 (1978) 672.
25 E. Cremer, T. Kraus and E. Bechtold, Chern.-Ing.-Teah., 33 (1961) 632.
26 C.W. Rice. U.S. Pat., 2 550 498 (1951).
27 R. Goulden. E.S. Goodwin and L. Davies, Analyet (London), 88 (1963) 951.
28 K.H. Kingdon and I. Langmuir, Phye. Rev., 21 (1923) 380.
29 I. Langmuir and K.H. Kingdon, Proa. R. Soc. London, Sera A. 107 (1925) 61.
30 H. Moesta and P. Schuff, Ber. Bunaengee Phye. Chern., 69 (1965) 895.
31 E. Bechtold, Theeie, Innsbruck. 1962.
32 E. Cremer, H. Moesta and K. Hablik, Chem.-Ing.-Teah., 38 (1966) 580.
33 H.A. McLeod and W.P. McKinley, J. Ass, Offia. Anal. Chern., 50 (1967) 641.
34 G.G. Oevjatych. N.C. Agliulov and V.V. Lucinkin, Zavod. Lab., 33 (1967) 901.
35 E. Cremer, J. Gae ChromatogF., 5 (1967) 329.
91

Chapter 5

FLAME-IONIZATION DETECTOR

CONTENTS

5.1. Introduction . . • • • . . • • . • • . . • . . • . . . . • • . . " 91


5.2. Hydrogen atmosphere flame-ionization detector . • . . . . . . . " 92
5.3. Hydrogen atmosphere flame-ionization detector for silicon compounds 103
5.4. Flame-ionization detector with hydrocarbon background 105
5.5. Selective detection of halogen compounds 106
References . . . . . . . • . . . . . . . . . . . . . 106

5.1. INTRODUCTION

The basic reactions taking place in the oxygen-hydrogen flame can be described
by the following equations t :
H + O2 ~ OH + 0 (5.1)
o + H2 ~ OH + H (5.2)
OH + H2 H20 + H
~ (5.3)
The radicals formed in the reaction zone of the flame react exothermically as
follows (M represents a general flame-gas molecule):
H + H + M~ H2 + M (5.4)
OH + H + M ~ H20 + M (5.5)
and, if organic compounds are present, chemi-ionization occurs:
CH + 0 ~ CHO+ + e- (5.6)
Various substituents on the individual carbon atoms in the molecule of a
compound affect ionization in different ways (commonly reducing the ionization
compared with hydrogen). This effect has led to the determination of the so-
called number of effective carbon atoms in a molecule of a compound 2 , the
ionization of an alkane carbon atom being taken as a reference, i.e., a value
of unity is ascribed to the response due to an alkane carbon atom. For this
reason, the flame-ionization detector (FID) is expected to respond only to
organic compounds containing an effective (ionizable) carbon. Inorganic gases,
carbon oxides and carbon disulphide should yield no response (the effective
carbon number for C=O and C=S being zero). It has been found, however, that,
92

at increased hydrogen flow-rates 3 and with an electrode geometry that causes


the electrodes to be heated by the flame 4 , carbon disulphide gives a response
that is a thousandth to a hundredth of that of hydrocarbons. At increased
hydrogen flow-rates, the detector gives a response to nitric oxide 5 that is
commensurate with the methane response. Under these conditions, the detector
gives a lower response 5 ,6 also to carbon dioxide, carbon monoxide, nitrous
oxide, nitrogen dioxide, helium, oxygen and hydrogen sulphide.
None of the above-mentioned responses has the character of a selective
response, because the molar responses to organic carbon compounds differ from
each other only in the number of the effective carbon atoms in the solute
molecules; the slightly anomalous responses to inorganic gases are small com-
pared with the hydrocarbon responses and they display the same polarity (the
number of ions is increasing).
The aim of this chapter is to describe modifications of the flame-ionization
detector that give a selective response to certain types of compounds with
regard to a higher response and opposite polarity. In fact, this chapter should
also include the alkali flame-ionization detector (AFIO), because it is a
modification of the FlO. However, the modifications of the AFIO cover a broader
spectrum (e.g., additional electrode, the method of heating the alkali metal
source) and, essentially, the AFIO is so widely and manufactured distributed
that it has been considered as a separate type of detector.
In 1969, an FlO was described 7 the electrodes of which were made of 0.051 mm
diameter high-purity platinum wires. These microelectrodes were adjustable and
it was with their varying distances from the detector jet that a selective
response, i.e., different responses for different compounds, was obtained. The
responses to ketones and aliphatic and aromatic hydrocarbons displayed maximum
values at a vertical distance of the electrodes from the jet of ca. 6 mm,
whereas the response to polychloroalkanes was maximal at a distance of ca. 2 mm.
When the electrodes were placed in the region of the flame corresponding to the
optimum response for aliphatic or aromatic hydrocarbons, the chloroalkane
response was four orders of magnitude lower.

5.2. HYDROGEN ATMOSPHERE FLAME-IONIZATION OETECTOR

Aue and HillS found in 1972 that, when the hydrogen and air flows were
interchanged in the flame-ionization detector, i.e., when air was introduced
into the jet while hydrogen was flowing around it, this detector gave a selective
response to organometallic compounds. In this instance the flame burns in a
hydrogen atmosphere and the hydrogen flow is an order of magnitude higher than
the air flow (hence the name hydrogen atmosphere flame-ionization detector,
93

Igniter

Collector electrode ~~~--90V


assembly

from GC

Fig. 5.1. Cross-section of a silane-doped HAFID. (From ref. 10.)

HAFlD). Further investigations showed that the introduction of silicon compounds


into the detection system (e.g., by the carrier gas from the stationary phase,
the silicon septum, etc.) is essential for the HAFlD response to organometallic
g
compounds to occur .
On the basis of these findings, a detector was designed 10 to which gaseous
silane was supplied (Fig. 5.1). The carrier gas from the column and an oxygen-
nitrogen mixture enter the 1.2 mm quartz jet of the detector, which is 70 mm
from the base of the detector body. The hydrogen and silane inlet is situated
at the detector base near the jet. The collector electrode is of the shape of
a 5-mm diameter ring, 50 mm from the jet. Under these conditions, the HAFlD
responds to organometallic compounds·and this response, depending on the presence
of a metal in a molecule of the compound, is several orders of magnitude higher
than the response to hydrocarbons (see Table 5.1). The detector is most sensi-
tive to manganese and aluminium compounds, the minimum detectable mass rate for
these compounds and for iron, tin and chromium compounds being lower than or
commensurate with the FlD minimum detectability of hydrocarbons (1.7 I 10- 14 g/
sec of Mn for manganese"). The background current of the HAFlD is higher than
that of the FlD'2,13. Therefore, the noise of the HAFlD is increased by approx-
imately an order of magnitude'3. The detection limit and selectivity are also
evident from Figs. 5.2 and 5.3. A comparison of the gasoline chromatograms
TABLE 5.1

RESPONSES OF MODEL COMPOUNDS

Selectivity against tetradecane. taken as the ratio of injected amounts of tetradecane and the organometallic that produce
the same peak area of 1.10- 11 C.
A minus sign preceding a selectivity value indicates a negative response. (From ref. 10. )

Compound Column temperature Sel ecti vity Detection 1 imit


(oC) (g)

Aluminium(III) hexafluoroacetylacetonate 50 3.2 • 10 5 6.0 • 10- 13


Ferrocene 145 1.9 • 10 45 2.1 • 10- 12
Tetravinyltin 110 2.6 • 10 1.0 • 10- 11
Tetrapropyltin 140 1.9 • 10 4 7.8 • 10- 12
Tetraethyl ti n 90 1.6 • 10 4 2.1 • 10- 11
Chromium(III) hexafluoroacetylacetonate 60 1.6 • 10 4 1.5 • 10- 11
Chromium hexacarbonyl 30 1.5 • 10 4 2.3 • 10- 11
Chromium(III) trifluoroacetylacetonate 170 1.1 • 10 44 1.7 • 10- 11
Tetraethyllead 100 1.1 • 10 3 5.1 • 10- 11
Tetrabutyll ead 190 6.0 • 10 7.5 • 10- 11
Triphenylantimony 250 1. 2 • 10 3 1.3 • 10- 10
Tungsten hexacarbonyl 50 8.6 • 102 5.3 • 10- 10
Molybdenum hexacarbonyl 30 5.0 • 10 2 3.5 • 10-9
Tetrabutylgermane 140 3.5 • 10 2 5.5 • 10- 10
Tri-n-butyl phosphate 190 -3.2 • 10 2
2 1.1 • 10-9
Chlorobenzene 70 -3.2 • 10 1.4 • 10-9
Bromobenzene 80 -3.2 • 10 2 1.4 • 10-9
Triphenylarsine 250 -3.2 • 10 2 2.3 • 10- 9
Triphenylbismuth 250 1.3 2 6.3 • 10- 9
• 10
Piaselenole 165 -9.2 • 10 1 5.7 • 10- 9
Iron(III) trifluoroacetylacetonate 175 8.0 • 10 1 4.1 • 10- 9
Di-n-butyl disulphide 140 -2.0 • 10 1 2.3 • 10-8
Fluorobenzene 30 7.1 • 10 0 6.7 • 10-8
Nitrobenzene 110 3.1 • 10 1 1.8 • 10- 7
Di ethyl mercury 70 2.9 • 10 0 1.3 .10- 7
Tetraethylsilane 50 1.9 • 10 0 4.1.10- 7
Tetradecane 140 1.0 • 10 0 3.7 • 10- 7
96

(Fig. 5.3) obtained by means of an HAFlD and a FlD clearly shows the advantage
of the HAFlD in detecting organometallic compounds in mixtures with hydrocarbons.
The molar response of the compounds seems to be given by the number of hetero-
atoms in the molecule of the compound 13 (Table 5.2). Similarly to the FlD, the
linear dependence of the molar response on the number of effective carbon atoms
in the molecule is retained l2 .
The wide gap between the jet tip and the base of the detector body is of
great importance with regard to the selectivity of the HAFlD response 13 • Table
5.3 compares the response selectivities and the detection limits for a common
massive jet and a jet tip modified with quartz or stainless-steel tubing (10 mm),
with the same distance of the collector electrode from the jet tip. Compared
with a normal jet, the response of a detector with a jet tip is substantially
lower for hydrocarbons and, therefore, the selectivity of the response to
hydrocarbons is higher. The reduction in the response to hydrocarbons in the
HAFlD compared with the FlD is believed to be due to an oxidation step that
occurs in the oxygen-rich precombustion zone of the HAFlD I2 ,13, in addition to

a b c

Fig. 5.2. Chromatograms of model compounds: (a) 5 pg of aluminium(lll) hexa-


fluoroacetylacetonate, 50 0 C; (b) 10 pg of ferrocene, 145 0 C; (c) 50 pg of
chromium(lll) trifluoroacetylacetonate, 170 0 C. Glass column (1 m) packed with
6% OV-17 on Chromosorb W, temperature as indicated. Flow-rates: hydrogen 1600
ml Imi n pl us 7 lJ 1Imtn silane, oxygen 125 ml 1m; n, ni trogen 100 ml Imi n, ni trogen
fron1 a GC column 40 ml/min. (From ref. 10).
97

FlO FlO
«
~
'0

iii
~ HAFID HAFID
«
en
'0 4
~
6 3
~ 5

2
6

.J IJ' J>. A
I I I I II

30 50 00 120 ·C 50 90120 ·C

Fig. 5.3. HAFIO and FlO analyses of alkyl lead compounds in gasoline. Left, regular
gasoline; right, premium gasoline. Glass column (6 ft. x 2 mm 1.0.) packed with
aO-l00 mesh Ultrabond 20 M. Temperature, programmed from 300C (5 min) to 120 0C
(5 min) at 20 0C/min. HAFIO flow-rates: hydrogen 1600 m1/min, silane 34 ppm,
oxygen 150 m1/min, air 120 m1/min. (From ref. 11 with permission.)

the effect of the position and potential of the collector electrode. The massive
jet dissipates the heat from the precombustion zone, decreasing the oxidation
therein. The jet tip then reduces the cooling effect and enhances oxidation 13 .
The linearity of response of the detector in which oxygen enters the detec-
tion space at the detector base covers about three orders of magnitude for
aluminium, chromium, tin, lead and iron compounds 10 • If hydrogen is supplied to
the detector above the jet tip only, the response is not a linear function of
the amount of compound 11 .
\D
(Xl

TABLE 5.2

EVALUATION OF RESPONSE

(From ref. 13 with permission.)

Compound Detection limit (pg) Sensitivity (C/mol) Selectivity


FlD HAFID FlD HAFID FID HAFID
Dodecane 300 48,000 1.4 0.034 1.00 1.00
Diallyldibutyltin 800 800 0.38 12'" 0.27 3.5 • 10 2
Tetraethylti n 2000 38 0.52 29 0.37 8.5 • 10 2
Tetrabutyltin 250 70 1.8 31 1.3 9.1 • 10 2
Tetra-n-propyltin 1400 22 1.4 38 1.0 1.1 • 10 3
Hexabutylditin 760 13 2.1 55 1.5 1.6 • 10 3

* Due to the column.


TABLE 5.3

COMPARISON OF JET TIPS

(From ref. 13 with permission.)

Compound Massive jet Stainless-steel jet tip Quartz jet tip


Detection Sel ecti vity Detection Sel ecti vity Detection Selectivity
1 imit 1imit 1imit
Tetraethyl tin 38 pg 110 15 pg 1. 2 • 10 4 10 pg 5.9 • 10 4
Dodecane 48 ng 1 0.2 llg 1 0.5 llg 1

\0
\0
100

-0
.... _.-----. ~ 1
,
t 2
4
5
3
2

-l~j

- 1
104

~i~i-ri,i-'i-i~i-'i-ri'i-'i-i~i-'i-'i'i-'i-'-i
-600 -500 -400 -300 -200 -100 o +100 +200 V

Fig. 5.4. Background current and responses of model compounds at different


collector electrode potentials. 1, Background current; 2, tetradecane (2 ~g);
3, tetrabutyll ead (lOng); 4, tetrabut}'lti n (10 ng); 5, ferrocene (1 na). h =
Peak height (arbitrary units, logarithmic scale). (From ref. 10.)

Fig. 5.4 shows the relationships between response and collector electrode
potential for tin, lead and iron compounds and for a hydrocarbon. The negatively
polarized electrode is clearly more advantageous; compared with the positive
collector electrode, the response to organometallic compounds in higher and the
hydrocarbon response and background current are lower. The responses of all of
the investigated compounds and the background ionization current increase to
maximum values as the potential increases. The optimum electrode potential with
regard to the selectivity and the minimum background current is -90 V.
For a constant concentration of a silicon compound entering the detector,
the response of the HAFID depends on the distance of the collector electrode
from the jet9 ,10. The response to organometallic compounds increases to maximum
values with increasing distance of the electrode. This response enhancement is
1.5 orders of magnitude for a distance ranging from 10 to 50 mm. On the other
hand, the hydrocarbon response decreases by approximately an order of magnitude
within this range. The background current also decreases with increasing
distance of the collector electrode from the jet.
The dista~ce of the collector electrode from the jet also affects the amount
of silane that has to be supplied to the detector. As mentioned above, the HAFID
gives a selective response to organometallic compounds only in the presence of
a silicon compound. The HAFID response and the background current increase to
101

ELECTRODE
0.4 HEIGHT

0.3 50 mm
'"52 70 mm
« 0.2 90 mm
l-
I
l!)
30 mm
u:i 0.1 110 mm
I
~
4:
LJ.J
a.. 0.0
10 mm
-0.1

Fig. 5.5. HAFID response as a function of silane concentration for various


electrode heights. Flow-rates: hydrogen 1500 ml/min, oxygen 120 ml/min, air
165 ml/min. (From ref. 14.)

maximum values with increasing concentration of silane in the detector. The


response to hydrocarbons and the background current attain this maximum value
at higher silane concentrations than does the response to organometallics 10
The optimum silane concentration, however, is different for each distance
between the collector electrode and the jet 14 (Fig. 5.5). The greater the
electrode distance, the lower is the optimum silane concentration. With optimum
silane concentrations the HAFID response to organometallics increases with
increasing distance of the electrode. Greater electrode distances (70 mm) are
advantageous for this reason, because both minimum detectability for organo-
metallics (improved sensitivity of response, lower background current and noise)
and selectivity (higher sensitivity of response to organometallics and reduced
sensitivity to hydrocarbons) are increased.
As can be seen from Fig. 5.5, the response of an HAFID of the type used is
negative at a distance of the collector electrode from the jet of 10 mm 14 This
negative response is related to the shape and position of the collector elec-
trode. The electrode was pin-shaped and situated out of the jet centre in this
instance. When using a ring-shaped electrode situated above the jet, the
response is also positive for a distance of 10 mml0. The optimum distance for
negative response depends on the entire detector design 15 . Negative responses
also occur at low positive potentials of the collector electrode (see Fig. 5.4)
and at high silane concentrations 10 . The detection limits for tetraethyltin and
tetraethyllead in the negative mode are lower by about one order of magnitude
than those in the positive mode 15 . However, parameter settings for the negative
response must be carefully controlled and a relatively high silane concentra-
tion may prove to cause long-term degradation of response due to silicon dioxide
102

deposits on the electrode. Thus for routine analyses the positive mode seems to
be more suitable 15 •
The HAFID requires substantially higher hydrogen flow-rates than the FlO.
The HAFID response to organometallics increases to a maximum at about 3 1/min 14
with increasing hydrogen flow-rate (4 ppm of silane, 70 mm electrode distance),
but the hydrocarbon response drastically increases with flow-rates below 650
ml/min 9 • The means of supplying hydrogen to the detector affects the amount of
silane needed to achieve maximum response, both in positive 11 and negative mode 15.
In the positive mode, if hydrogen enters the detector 21 mm above the jet tip
this concentration is 34 ppm, and if hydrogen is supplied to the detector base
the optimum silane concentration is below 10 ppm (see Fig. 5.5).
The dependence of the response of an organometallic compound on the percent-
age of nitrogen in the air-oxygen mixture supplied also shows a maximum (at a
1:1 nitrogen to oxygen ratio). Therefore, for a total flow-rate through the jet
of 305 ml/min and a nitrogen carrier gas flow-rate of 20 ml/min, the optimum
air and oxygen flow-rates are 165 and 120 ml/min, respectively14. With higher
nitrogen to oxygen ratios the response to organometallics decreases, the
hydrocarbon response increases and, consequently, the detection selectivity is
diminished 10 .

5.3. HYDROGEN ATMOSPHERE FLAME-IONIZATION DETECTOR FOR SILICON COMPOUNDS

If a silicon compound is introduced into the HAFID, it gives a selective


response to metallic compounds. If an inverted system is used, i.e., if an
organometallic compound is supplied to the detector, the HAFID responds to
organic silicon compounds. Hence this is again an HAFID, but an iron compound,
commonly ferrocene, is supplied in the hydrogen flow in this instance 16 •
The detector geometry plays an important role in peak tailing 17 • Pronounced
tailing occurs with larger inner diameters of the detector (above 11 mm).
Hydrogen is introduced into the detector space below the jet level and oxygen
is supplied to the detector jet together with the carrier gas from the column.
The collector electrode (-90 V) is 110 mm from the jet.
The amount of iron compound supplied to the detector fundamentally changes
the response16~18-21 (Fig. 5.6). With increasing amounts of an iron compound
the response to silicon compounds increases to a maximum (about 5 ppm in this
instance) and then decreases, passing to a negative response (at about 12 ppm).
The specific course of this dependence is affected by the distance of the point
of the inlet of the iron compound from the detector flame 20 (the inversion
from a positive to a negative response is shifted to higher amounts of the iron
compound at larger distances) and by the type of iron compound supplied 18 ,20
The maximum negative response (about 35 ppm of ferrocene) is three times greater
than the maximum positive response. The detector noise increases at the same
103

~ 20
I
0
~
15
4:
Q.o
0/1
10
c:
0
a.
0/1
5
Q.o
L- Amount of ferrocene (ppm)
0
III
I
Cl -5 30 40
~
4:
x 10

-15

-20

Fig. 5.6. Effect of amount of ferrocene on the HAFID-Si response. (From ref. 18
with permission.)

TABLE 5.4

DETECTION LIMITS FOR THE HAFID-Si

(From ref. 21.)

i·lode of HAFI D Coumpounds containing


operation
Silicon Phosphorus Iron Chlorine
Non-doped 4 ng (ref. 16) 7 ng 0.6 ng 50 ng
Positive mode 50 pg (ref. 17) 0.5 ng 1 ng 2.5 ng
Negative mode 1 ng (ref. 18) 1 ng 14 ng 9 ng

time, so that the detection limit is smaller by a factor of about 20 in the


negative mode (see Table 5.4). Linearity of detection covers a range of about
three orders of magnitude in both mOdes 21 .
Hence, the selectivity of response of this detector,· called the HAFID-Si, is
conditioned by the response level and response polarity (Fig. 5.7). With smaller
amounts of doping iron compounds in the positive mode, the HAFID response to
silicon compounds is higher than that of a detector in the absence of a doping
agent and approximately commensurate with the response of a common FID, whereas
the response to hydrocarbons is reduced. The selectivity of response of Si to C
is 18,21 about 10 4 :1. With higher concentrations of iron compounds in the nega-
tive mode, the response to silicon compounds is negative whereas that to
hydrocarbons remains positive. The change from a positive to a negative mode
is easy to accomplish, by increasing the flow-rate of the gas containing the
104

124 FlO
a b

5 8

HAFIO lSi)
9
positive mode

3 5
5 2

70 B4 98 112 126 "C 70 84 98 112 126 "C

1 C
2

HAFIO lSi)
ne9ative mode
5

Fig. 5.7. Chromatograms of equal amounts of compounds using different detection


methods. Chromatograms obtained with a standard FID (a); a HAFID doped with
5 ppm of ferrocene (b); a HAFID doped with 30 ppm of ferrocene (c). Peaks: 1 =
Hexane (solvent); 2 = chlorobenzene; 3 = decane; 4 = octanol; 5 = triethyl
phosphate; 6 = dodecane; 7 = decanol; a = ferrocene; 9 = tetradecane. Methyl-
silicone-coated fused-silica capillary column (10 m x 0.2 mm I.D.). Temperature,
programmed from 70 to 150 0 C at aOC/min. Helium flow-rate: 1.2 ml/min, HAFID gas
flow-rates: hydrogen 1600 ml/min, oxygen 130 ml/min. FID gas flow-rates: hydro-
gen 30 ml/min, air 240 ml/min. (From ref. 21.)

doping iron compound. A selective enhancement of the response to silicon


compounds compared with hydrocarbons was also observed 18 when other metals were
used as doping agents, such as aluminium, nickel, chromium, molybdenum, brass,
platinum, copper and magnesium. The lifetimes of the sources of these metals
are short, however.
105

The HAFIO-Si gives a response approximately commensurate with that of silicon


compounds also to phosphorus, iron and chlorine compounds '8 ,2' (see Table 5.4).
The response to the latter compounds also depends on the amount of doping agent
applied. The course of this dependence for phosphorus and iron compounds is
similar to that for silicon compounds, i.e., the response to these compounds
is negative for larger amounts of dopants supplied to the detector (Fig. 5.7).
Alcohols, ketones, ethers and nitrogen and fluorine compounds give a low posi-
tive response 20 .
The HAFIO responds to organic silicon compounds '6 ,17 and to phosphorus, iron
and chlorine compounds 21 even if no iron compound is supplied to the detector.
In this instance, the sensitivity of detection is lower than in the presence
of ferrocene in the positive mode (see Table 5.4), being approximately two to
three orders of magnitude lower than that of the FlO. The detector does not
differ in design from that with ferrocene '7 , the Si to C selectivity is about
3.5 orders of magnitude and the detection limit is 4 ng tetraethylsilane.

5.4. FLAME-IONIZATION OETECTOR WITH HYDROCARBON BACKGROUND

Fritz et al. 22 found with an FlO that the response to silicon compounds showed
an inversion starting from a certain mass flow-rate of the silicon compound.
They ascribed this effect to two contradictory phenomena: the production and
the loss of charged particles in the system. Based on these findings, they
created conditions causing the ion loss to prevail in the detector. They used
a hydrogen-hydrocarbon flame (by supplying acetylene to the detector burner)
and found that the detector response to organic silicon compounds was negative
when the hydrocarbon concentration in the detector exceeded a certain value.
On modifying this detection system by introducing methane into the hydrogen
flow 23 (tenths of a millilitre per minute of methane) to the FID (oxygen-hydrogen
flame), the negative response of this detector attained a value of 0.2 C/g of
silicon. This value considerably exceeds the response of the common FlO as
expressed in coulombs per gram of carbon.
The FlO to which a hydrocarbon is supplied gives, at certain hydrogen flow-
rates, a negative response to inorganic gases to which the common FlO mostly
is not sensitive (see section 5.1). On supplying methane to the detector, either
in the hydrogen or in the carrier gas in amounts causing the background current
due to hydrocarbon ionization to attain values of ca. 3 • 10- 10 A, the response
of this detector 24 to hydrogen sulphide, sulphur dioxide and carbon disulphide
is about 2 • 10- 4 C/mole (which represents a detection limit of 1 • 10-8 mole),
the response to oxygen and nitrous oxide is about 4 • 10- 5 C/mole 25 and the
response to carbon oxide and carbon dioxide is about 3 • 10-6 C/mole 6 . When
106

perfluoromethane was used as the carrier gas, the response to nitrogen, helium
and carbon dioxide was about 5 • 10-6 C/mole 26 .

5.5. SELECTIVE DETECTION OF HALOGEN COMPOUNDS

Under certain conditions (higher hydrogen flow-rate, the electrode positioned


near the flame), the sensitivity of the flame-ionization detector for halogen
compounds may be higher than it is under common conditions 4 When hydrogen is
employed as the carrier gas and oxygen instead of air, the sensitivity of the
FID towards halogen compounds is approximately two orders of magnitude higher
than that of the common detector 27 , The response is in direct proportion to the
number of chlorine atoms in the molecule. For instance, with flow-rates for
hydrogen of 100 ml/min and oxygen of 50 ml/min, and an applied potential of
+170 V, the molar responses of mono-, di- and trichlorobenzene were 23, 45 and
69 C, respectively, i.e., in the ratio 1:2:3. However, the background current
and the detector noise are also several orders of magnitude higher than they
are with the standard FID28 (1 • 10- 9 and 5 • 10- 10 A, respectively).

REFERENCES

1 T.M. Sugden, Ann Rev. Phys. Chem., 13 (1963) 369.


2 J.C. Sternberg, W.S. Gallaway and D.T.L. Jones, in N. Brenner, J.E. Callen
and M.D. Weiss (Editors), Gas Chromatography, Academic Press, New York, 1962,
p. 231.
3 B.L. Walker, J. Gas Chromatogr., 4 (1966) 384.
4 M. Dressler, J. Chromatogr., 42 (1969) 408.
5 P. Russev, T.A. Gough and C.J. Woolam, J. Chromatogr., 119 (1976) 461.
6 B.A. Schaefer and D.M. Douglas, J. Chromatogr. Sci., 9 (1971) 612.
7 R.W. McCoy and S.P. Cram, J. Chromatogr. Sci., 7 (1969) 17.
8 W.A. Aue and H.H. Hill, Jr., J. Chromatogr., 74 (1972) 319.
9 H.H. Hill, Jr. and W.A. Aue, J. Chromatogr. Sci., 12 (1974) 541,
10 H.H. Hill, Jr. and W.A. Aue, J. Chromatogr., 122 (1976) 515.
11 M.D. DuPuis and H.H. Hill, Jr., Anal. Chem., 51 (1979) 292.
12 J.H. Wagner, C.H. Lillie, M.D. DuPuis and H.H. Hill, Jr., Anal. Chem., 52
(1980) 1614.
13 D.R. Hansen, T.J. Gilfoil and H.H. Hill, Jr., Anal. Chem., 53 (1981) 857.
14 J.E. Roberts and H.H. Hill, Jr., J. Chromatogr., 176 (1979) 1.
15 D.R. Hansen and H.H. Hill, Jr., J. Chromatogr., 303 (1984) 331.
16 H.H. Hill, Jr. and W.A. Aue, J. Chromatogr., 140 (1977) 1.
17 M.A. Osman, H.H. Hill, Jr., M.W. Holdren and H.H. Westberg, Anal. Chem., 51
(1979) 1286.
18 M.A. Osman and H.H. Hill, Jr., J. Chromatogr., 213 (1981) 397.
19 M.A. Osman and H.H. Hill, Jr., J. Chromatogr., 232 (1982) 430.
20 M.A. Osman and H.H. Hill, Jr., Anal. Chem., 54 (1982) 1425.
21 M.A. Osman and H.H. Hill, Jr., J. Chromatogr., 264 (1983) 149.
22 D. Fritz, G. Garz6, T. Szekely and F. Till, Acta Chim. Hung., 45 (1965) 301.
23 G. Garz6 and D. Fritz, in A.B. Littlewood (Editor), Gas Chromatography 1966,
Institute of Petroleum, London, 1967, p. 150.
107

24 B.A. Schaefer, Anal. Chem., 42 (1970) 448.


25 B.A. Schaefer, J. Chromatogr. Sci., 10 (1972) 110.
26 W.C. Askew, Anal. Chem., 44 (1972) 633.
27 A.E. Karagozler, C.F. Simpson, T.A. Gough and M.A. Pringuer, J. Chromatogr.,
158 (1978) 139.
28 C.F. Simpson and T.A. Gough, J. Chromatogr. Sci., 19 (1981) 275.
This Page Intentionally Left Blank
109

Chapter> 6

PHOTOIONIZATION DETECTOR

CONTENTS

6.1. Introduction • • • • • • • • • • • • • . • • • • • 109


6.2. Response model • • • • . • • • • • • • • • • • • • 111
6.3. Sensitivity of response and minimum detectability 112
6.4. Selectivity of response 118
6.5. Carrier gas 126
References ••• . • • • • • 131

6.1. INTRODUCTION

When a compound AB is irradiated with UV light, ionization occurs if the iozi-


zation potential of the compound is equal to or smaller than the photon energy,
hv. The photoionization process is initiated by the absorption of UV light:

(6.1 )

At present, commercial photoionization detectors (PIDs) (Fig. 6.1) are equipped


with a radiation source, viz., a UV lamp separated by a wirdow from the ionization
chamber. Interchangeable lamps with various energies of the emitted photons (com-
monly 9.5, 10.2 and 11.7 eV) have been employed as radiation sources. The window
material consists of MgF 2 and LiF (for photon energies exceeding 10.9 eV). The
ionization chamber contains two electrodes to which a voltage is applied. The
ions formed flow to the collector electrode in the electric field, and the gen-
erated current is measured.
In earlier types of PIDs, a glow 2- 7 or microwave 8 discharge with no separa-
tion of the radiation source and the ionization chamber (see Fig. 6.2) was used
as the source of UV radiation. The discharge occurred at sub-atmospheric pres-
sure; therefore, the ionization chamber was run at a'low pressure and the chro-
matographic column terminated in a space at sub-atmospheric pressure. These de-
tectors have several other drawbacks 9 : they need a continuous flow of highly puri-
fied gas into the discharge zone, they are of a cumbersome design (necessitating
a vacuum pump), contamination of the electrodes resulting in aging occurs and
the temperature limit is 170 oC8. The source of UV radiation produces 9 , in addi-
tion to photons, ionized particles and excited molecules. The ionization of the
110

Fig. 6.1. Schematic diagram of a PIO manufactured by HNU Systems. 1 = Ceramic


internals; 2 = ionization chamber; 3 = glass-lined inlet; 4 = glas-lined ex-
haust; 5 = UV lamp. (From ref. 1.)

I
Fig. 6.2. Schematic diagram of a glow discharge PIO. 1 = Gas inlet to source of
UV radiation and discharge cathode; 2 = discharge anode; 3 = sensing chamber
cathode; 4 = carrier gas inlet and anode of sensing chamber; 5 = outlet to
pump. (From ref. 2, with permission.)

substance can be initiated not only by photons and therefore photoionization is


not the only process occurring in the detector, which leads to difficulties in
interpreting the mechanism of detection.
A PIO provided with a separated source of UV radiation lO eliminates most of
the disadvantages that occur with the windowless detector. This detector can be
run at atmospheric pressure. A combination of this PIO with an electron-capture
detector (ECO) (see also Chapter 11) has been described 11 • In this instance, the
detector works as a PIO or as an ECO based on photoionization.
III

6.2. RESPONSE MODEL

The processes taking place in the PIO are represented by the following kinetic
scheme lO :

AB + h ->- AB" Rl =I0 - I (6.2)


AB" ->- AB+ + e- R2 = K2 [AB"] (6.3)
AB* ->- A+ B R3 = K3 [AB*] (6.4)
e + anode + current R4 = K4 [e-] (6.5)
AB+ + cathode + AB R5 = K5 [AB+] (6.6)
AB+ + e- + C ->- AB + C R6 = ex [AB+] [e-] (6.7)
AB" + C ->- AB + C R7 = K7 [AB"] [C] (6.8)

where R is the reaction rate, [AB] is the concentration of substance AB, C is the
carrier gas of concentration [e], I O is the initial photon flux, I is the intensity
of transmitted radiation, I O - I is the number of photons absorbed per second and
ex is the recombination coefficient.
The following expression applies for the equilibrium state:

(6.9)

The probability of the photon being absorbed (eqn. 6.2) depends on the absorp-
tion cross-section cr of the substance (defined by the Beer-Lambert law). The
probability of ionizing of the excited state depends on the photoionization ef-
ficiency, n, where

(6.10)

If the recombination reaction (eqn. 6.7) is suppressed by a high enough volt-


age, we can write for the ionization current (il of the detettor 12 ([C] ~ 1):

(6.11 )

where N is Avogadro's number, F is Faraday's constant, and L is the path length.


For a given detector and lamp, the m0lar response of the PIO is proportional
to the molar concentration and the photoionization cross-section, cr n:

R = if [AB] = kan (6.12)


112

There is a direct dependence of the photoionization cross-section on the


photon energy and ionization potential 13 . Even though this is a complex depen-
dence, the FlO signal is related to the ionization potential by means of the
photoionization cross-section 12 •

6.3. SENSITIVITY OF RESPONSE ANO MINIMUM OETECTABILITY

The molar response of the PIO depends on the photon energy14 (Fig. 6.3). The
dependence is given by the relationship between the photoionization efficiency
and photon energy (Fig. 6.4). As can be seen, the low alkane response at a photon
energy of 10.2 eV is associated with the low photoionization efficiency; similar-
ly, the high response to aromatic hydrocarbons is related to the high ionization
efficiency at this energy. Near the ionization potential, alkenes, cyclic alkenes
and alkylbenzenes also show a similar rapid increase to the maximum ionization
efficiency. This course is typical of photoionization involving TI electrons 15 .

RMR

o
10 15 20 eV

Fig. 6.3. Relative molar responses per mole of carbon (RMR) of an alkane
(hexane) and an aromatic hydrocarbon (benzene) as a function of the photon
energy. 1 = Hexane; 2 = benzene. (From ref. 14.)

As mentioned above, with a constant-energy source (e.g., a 10.2-eV lamp), the


sensitivity of response of the photoionization detector depends on the ioniza-
tion potential of the compounds and increases as their potential decreases 2 ,6,7,
10,12,14,16. Fig. 6.5 shows this dependence for aromatic hydrocarbons together
with the dependence of the response on the number of carbon atoms in the molecule
of the compound. Fig. 6.6 shows the dependence of the PIO sensitivity on the
number of carbon atoms in the molecule for other homologous series. The detector
response increases with increasing number of carbon atoms in the molecule of the
compound. The idea 17 that the molar response of the photoionization detector may
be related to the carbon number of the compound (correlation coefficient 0.67),
113

9.0 9.5 100 eV

Fig. 6.4. Relationship between the photoionization efficiency (n) and the photon
energy. E indicates the energy of the UV lamp. I, Alkane (hexane); 2, aromatic
(benzene). (From ref. 14.)

\
55 0

a::
50
\1F 0

:l:
a::
45

40--~--~~ ____ ~
o~o
______ ______- L__
~ ~~~

8.5 8.7 89 9.1 9.3


IP eV

6 7 8
c 9

Fig. 6.5. Relative molar response (RMR) of aromatic hydrocarbons as a function


of the ionization potential (curve IP) and the number of carbon atoms (curve C)
in a molecule. RMR of benzene = 1. c = number of carbon atoms. (From ref. 12.)

as with the flame-ionization detector, does not agree with the results, however.
The increase in the detector response with increasing number of carbon atoms in
the molecule invariably applies to particular homologous series, and is different
for different homologous series 12 ,18. Obviously, it is mainly due to the decrease
114

60

Fig. 6.6. Relative molar response (RMR) as a function of the number of carbon
atoms (C) in a molecule. R/4R of benzene = 1. 1 = Styrene; 2 = xylene; 3 a aro-
matics; 4 = cyclohexanes; 5 = cyclohexenes; 6 = 1-alkenes; 7 = n-alkanes. (From
ref. 12.)

in the ionization potential with increasing carbon number of the compound {see
Fig. 6.5)12. In addition, the relationship between the pro sensitivity and the
carbon number of the members of a homologous series is not always linear.
When comparing, for instance, the responses to hydrocarbons with six carbon
atoms in the molecule, the response is found to increase with increasing degree
of unsaturation: benzene> cyclohexene > cyclohexane (a linear dependence of
the response on the number of double bonds in the molecule is valid for CI8
fatty aCidsl ). For different types of compound, the sensitivity increases in the
following orders l8 :
alkanes < alkenes < aromatics;
alkanes < alcohols < esters < aldehydes < ketones;
non-eye 1i c compounds < eye 1i c compounds;
fl uorine-substituted < chl ori ne-substi tuted < bromi ne-substituted < i odi ne-
substituted compounds;
benzenes substituted with electron-withdrawing groups < benzene < benzenes
substituted with electron-releasing groups.
The theoretical relationship between the photoionization cross-section and the
ionization potential is complex, including other factors such as molecular geo-
metry and symmetryl9,20. For instance, the relative molar responses (RMR) (re-
lative to benzene) for 0- and m-xylene, which have the same ionization potential
(IP) of 8.56 eV, are 1.14 and 1.15, respectively18,19. The RMR for the more sym-
metric p-xylene, with an IP of 8.445 eV, is 1.2. The RMR of the isomeric ethyl-
115

3
R
2

a
7 8 9 10 11 12

IP eV

Fig. 6.7. Molar response (R) as a function of the ionization potential (Lang-
horst's data). (From ref. 19.)

benzene, with a much higher IP of 8.76 eV, is 1.16, similar to those of 0- and
m-xylene. In spite of the complexity of the overall situation, the ionization
potential is believed to be the most important factor determining the PIO re-
sponse 19 • The dependence of the molar response on the IP for 50 compounds from
the Langhorst set 18 of response data gives a linear regression line with a cor-
relation coefficient of 0.89 (Fig. 6.7). The assumption 21 that the number of n
electrons may be a factor important to the response of the individual compounds
seems to be less acceptable 19 , because the mentioned dependences yield low cor-
relation coefficients in this instance (0.25 for the dependence from Fig. 6.7).
The large discrepancy can also be caused by the fact that the 10.2-eV source is
not monochromatic 20 (see Table 6.5).
The response level also depends on the lamp intensity setting. The response
improves with increasing intensity but, at the same time, the lamp life decreases.
The detector noise also increases with increasing intensity.
Table 6.1 lists the basic parameters of various types of PIOs. As can be seen,
the commercial PIO with a separated radiation source manufactured by HNU Sys-
tems 9 ,22,23 displays the most favourable detection limit (2.10- 12 g of benzene),
detector noise (2.10- 14 A) and dynamic linear range (10 7) (similar parameters
apply to pros manufactured by Tracor 24 ). The sensitivity of the HNU Systems de-
tector is17 0.3 C/g for aromatic hydrocarbons, i.e., about 24 C/mole for benzene
(flow-rate of helium carrier gas 30 ml/min). Hence the response is about 20 times
greater than that of the flame-ionization detector (FlO). A ca. 10-fold greater
molar resprnse was found for the Photovac detector, which operates with a high-
......
......
0'1

TABLE 6.1

COMPARISON OF PID OPERATING CHARACTERISTICS AND DESIGN

From ref. 9.

Ionization Background Linear Detection Noise Type of Sealed Maximum


efficiency current dynamic limit (pA) discharge lamp temperature
(pA) range (pg) (0C)

Lovelock 2 10- 4 500 10 4 Ar No


Price et al. 7 10- 3 200 10 5 1 0.1 Ar No
Freeman and Wentworth 8 10- 3 6000 10 4 40 40 He No 250
Krysl and Sevcik 10 1.6.10- 4 90 105 2 0.3 H2 Yes 100
Locke and Meloan 6 500 10 5 200 1 Ar No
Ostojic and Sternberg 16 5.10- 5 300 10 5 300 0.5 Ar, H2 Yes 350
Driscoll and Spaziani 9 8.10- 4 8 10 7 2 0.02 H2 Yes 250
117

frequency electrodeless source 25 . However, its noise is also about ten times
higher, so that the minimum detectability is approximately comparable to that
obtained with other detectors. This detector, operating at ambient temperature,
is mounted in a portable gas chromatograph designed mainly for analyses of atmos-
pheric samples.
The detection sensitivity for other compounds is indicated only as the detec-
tion limit for the individual compounds or it is compared with the sensitivity
of the FID. Table 6.2 shows the values for a 10.2-eV lamp. The sensitivity in-
dicated in this way is greater than that for the FID in all instances, the magni-
tude of the PID:FID ratio depending on the structure of the compound. The high
sensitivity to inorganic compounds is of prime importance. The'detection limit
is in the range of tens of picograms for H2S, PH 3 , 12 , AsH 3, CS 2 and NO, and is
200 pg for NH 3 • Thus, all mentioned compounds can be analysed using one detector
whose sensitivity surpasses by several orders of magnitude that of the only other
detector that could be used for all of these compounds, viz., the thermal con-
ductivity detector. Compared with other selective detectors (flame photometric
detector, Hall detector), the detection limit is approximately one order of
magnitude lower.

TABLE 6.2

DETECTION LII1ITS FOR PID

Compound Detection limit (pg) Ref.

Aromatic hydrocarbons 1-10 1


Barbiturates 100-1000 1
Polyaromatic hydrocarbons 50-100 1
Tetraethyllead 150 26
( CH 3)2 S 20 26
( CH 3 )2 S2 22 26
CH 3SH 20 26
30 26
~SS 15 26
P~3 20 26
AsH 3 25 26
NH3 200 26
NO 52 26
12 25 26

The dynamic linear range is claimed to be about 6.5, 5.75, 5 and 4.5 orders
of magnitude for photon energies (lamps used) of 10.2, 9.5, 11.7 and 10.9 eV,
respectively. This is due to the fact that the detection limit increases for the
individual lamps in the mentioned sequence, while the upper limit remains approx-
118

imately the same22 • However, the upper limit of the linear response for n-octane
was found to be about 1 ~g for the 10.2-eV source 20 •

6.4. SELECTIVITY OF RESPONSE

Under certain experimental conditions, the photoionization detector is also a


selective detector. Its selectivity is due to the ability of the substances to
be photoionized. This ability varies with the energy of UV radiation, i.e., it
differs when different lamps are used. Fig. 6.8 gives a schematic illustration
of the detector response when using 11.7-, 10.2- and 9.5-eV lamps. The detector
invariably responds to compounds that have ionization potentials that are
lower than the photon energies of the individual lamps. Hence the use of dif-
ferent lamps makes it possible to achieve the selective detection of various com-
pounds. The lamp with the lowest energy displays the highest selectivity. The
relative responses of a PID for about 30 compounds when using 9.5- and 10.2-eV
lamps, normalized to benzene (R = 100) for both lamps and related to the ioniza-

R 11.7 eV

~.romatics

Amines
NR
Mercaptans

Formic acid

Formaldehyde ---1--- Alkenes (except C H4)


2
Ac rylonit ri le

12 11 10 9 eV

Fig. 6.8. PID response for various UV lamps. (From ref. 1.)
119

Ii'
100

50

10

100 9.2
IP eV

Fig. 6.9. Relative sensitivity of a PID with a 9.5-eV lamp and a 10.2-eV lamp
as a function of the ionization potential of the compounds. R = relative res-
ponses (normalized to benzene = 100 for both lamps). (Reprinted from ref. 28,
with permission.)

TABLE 6.3

COMPARISON OF SOME HYDROCARBON RESPONSE RATIOS FOR PIDs WITH 10.2- AND 9.5-eV
LAt1PS AS A FUNCTION OF THE IONIZATION POTENTIAL

Normalized to benzene (for the low-boiling aromatics) or benzophenone (for the


higher-boiling polyaromatics) for 9.5- or 10.2-eV lamp. (Reprinted from ref. 27,
with permission.)

Compound Response rati 0 Improvement


in sel ecti vi ty,
IP (eV) PID (10.2 eV) PID (9.5 eV) 9.5 eV/1O.2 eV

Benzophenone 9.45 6.4 6.4 1.0


Benzene 9.25 1.0 1.0 1.0
Toluene 8.82 8.4 11.6 1.4
Xylene 8.45 5.3 22.3 4.2
Naphthalene 8.12 12.1 89.3 7.4
Anthracene 7.23 3.8 34.2 9.0

tion potential, are listed in Fig. 6.9. The responses (relative to benzene) of
several compounds are given in Table 6.3 and indicate the possibility of using
the response ratio of the two lamps in order to obtain some structural informa-
tion. As can be seen from Table 6.4, the 10.2-eV lamp can also yield a selective
response (see also Fig. 6.10). The selectivity of the response for the 10.2-eV
source is independent of the intensity of the source, whereas this is not so
for the 9.5-eV source 20 •
120

TABLE 6.4

COt1PARISON OF RESPONSES TO UNSATURATED COMPOUNDS FOR VARIOUS DETECTORS

Normalized to benzene. (Reprinted from ref. 27. with permission.)

Detector Response
Benzene Cyclohexene Cyclohexane

PID (lO.2-eV lamp) 1.0 0.30 0.15


PID (11.7-eV lamp) 1.0 1.05 1.00
FIO 1.0 1.00 1.00

2 6
9.SeV
7

1a.2eV

1tZeV

Fig. 6.10. Comparison of normalized PID responses (R) for seven selected naphtha
peaks. (Reprinted from ref. 27. with permission.)
121

Fig. 6.10 illustrates the case of using three different lamps for identifica-
tion purposes. It shows the results of analysing a naphtha sample, depicting the
responses of seven selected peaks normalized to benzene (10.2-eV lamp). Peak re-
sponses 1 and 5 are the same for the three lamps, which indicates an aromatic
hydrocarbon (the 10.2- and 11.7-eV ratios being approximately identical) with
an IP of about 9 eV or higher (the 9.5- and 10.2-eV ratios being approximately
identical). Peaks 2 and 3 have responses with the 11.7-eV lamps that are approxi-
mately double those with the 10.2-eV lamp, the analysed substance thus being an
olefin (cf., Table 6.4). The higher response obtained with the 9.5-eV lamp com-
pared with the 10.2-eV lamp indicates an IP smaller than 9 eV. Peak 6 shows an
11.7-eV to 10.2-eV response ratio of about 1, indicating an aromatic hydrocarbon,
but its high 9.5-eV to 10.2-eV ratio indicates a polyaromatic hydrocarbon. Peak
7 has a high 11.7-eV to 10.2-eV response ratio and a low 9.5-eV to 10.2-eV re-
sponse ratio, so that a straight-chain hydrocarbon can be assumed to be the sub-
stance being analysed 27 •
In practice, the PID also responds to compounds the ionization potentials of
which are slightly higher than the photon energyl2. The following factors cause
this phenomenon: (a) the energy difference between the molecules in excited vibra-
tional states and those in the ground state can be up to 0.4 eV less than the IP
and (b) contamination in the lamp that can give rise to small side-bands with
photon energies between 10.2 and 10.9 eV (for the 10.2-eV lamp). For the Photovac
detector, equipped with a high-frequency source and using photons with an energy
slightly below 11 eV, the response is delivered also by compounds with an IP
greater than 12 eV. The explanation lies in the formation of negative ions and
in some type of functional exchange between these large negative species and the
negative oxygen ions that commonly serve as negative carriers in the ion cell,
and/or in the reaction of the ozone produced in the ion chamber with non-ionizable
species, forming new compounds that can be ionized later 25 .
The narrower the energy spread of the incident photons, the greater is the
selectivity. The characteristic lines of four HNU Systems sources (Model PI 52-02
detector) are listed in Table 6.5. It follows from these data that (1) of the
four different sources, only the 8.3-eV lamp is monochromatic within the range
examined, but the energy of this line is 8.44 eV, not 8.3 eV; (2) the 9.5-eV
source is the worst, as nearly all the light emitted is. at 8.4 eV; of the remaining
ing lines, the most intense is at 9.57 eV, but others are present up to 10.88 eV;
(3) the 10.2-eV source has two sharp lines at 10.03 eV (83%) and 10.64 eV (17%);
(4) the 11.7-eV source again has two sharp lines (11.62 and 11.82 eV), but as
they are closer together than those of the 10.2-eV source, they do not create as
much of a problem when interpreting the detector response; (5) the output of the
lamps does not include an emission continuum, implying that the gases are at low
122

pressure; (6) the gases used are very pure; and (7) the importance of the lamp
window material can be seen in the spectra of the 8.3- and 9.5-eV sources: both
are filled with xenon, but the 8.3-eV source uses a window of a material that
cuts off completely above about 8.5 eV.

TABLE 6.5

CHARACTERISTIC LINES FOUND IN VACUUM UV OUTPUT OF HNU SYSTEMS LAMPS

(From ref. 20.)

Lamp Wavelength Energy Output Lamp type


designation
(eV) (nm) (eV) (%)

8.3 147.0 8.44 100 Xenon


9.5 114.0 10.88 0.03 Xenon
117.2 10.58 0.01
119.3 10.40 0.18
125.0 9.92 0.05
129.6 9.57 2.1
147.0 8.44 97.6
10.2 116.6 10.64 17.1 Krypton
123.6 10.03 82.9
11. 7 104.9 11.82 26.2 Argon
106.6 11.62 71.8
121.6 10.20 2.0

Table 6.6 gives a list of solvents that, owing to their ionization potential,
do not give a positive response, i.e., increase in ionization current, in the
photoionization detector. Fig. 6.11 compares the chromatograms of drug solutions
in acetonitrile obtained by means of a PIO and an FlO. The PIO yields a smaller
(negative) response to acetonitrile than the FlO. However, the drop of the base-
line is pronounced during the passage of the acetonitrile peak 28 ,29. Just as with
oxygen, the drop is due to electron capture by the acetonitrile molecule and to
the subsequent neutralization of the positive and negative ions (see eqns. 6.15
and 6.16). As this acetonitrile peak, and also the peaks of tetrachloromethane,
trichloromethane, methanol and water, diminish the response of the simultaneously
eluting compounds 30 , the latter should be separated from acetonitrile.
The PIO has frequently been used connected in parallel to an FI0 14 ,20,32-34.
The FlO responses to an effective carbon atom in alkanes, alkenes and aromatics
are approximately the same. The PIO response to a carbon atom of aromatic hydro-
carbons is several times greater than that of the FlO. The values found by
123

TABLE 6.6

SOLVENTS USED WITH PID THAT PRODUCE NO POSITIVE DETECTOR RESPONSE

(Reprinted from ref. 28, with permission.)

Lamp
9.S eV 10.2 eV 11.7 eV

H2O H2O H2O


CH 3CN CH 3CN CH 3CN
CH 30H CH 30H Freons (some)
C2HSOH CH 3Cl
CH 3Cl CH 2C1 2
CH 2C1 2 CHC1 3
CHC1 3 CC1 4
CC1 4 C2H4C1 2
C2H4C1 2 Freons
Freons
CSH12
C6H14
C7H16

Driscoll et al. 14 for the PID to FlO response ratio are <2, 2-4 and S-10, for
alkanes, alkenes and aromatics, respectively. This ratio also differs consider-
ably for sulphur-containing and chlorinated pesticides. It has been found to be
12-16 for aromatic compounds and 2 or less for aliphatic compounds. Fig. 6.12,
depicting the chromatograms of automobile exhaust gases, illustrates the ap-
plication of this method of determining compound types. Table 6.7 shows the re-
sponse ratios of other groups of substances under different experimental con-
ditions.
The response ratio of an electron-capture detector (ECD) and a PID has been
used for the selective differentiation of nitro compounds 11 • This ratio differs
by several orders of magnitude for polyaromatic hydrocarbons and their nitro
derivatives (see Table 6.8). Fig. 6.13 again indicates the possibility of
selectively distinguishing certain groups of compounds by simultaneously em-
ploying an FlO and a PID. For nitro toluenes, the PID yields a relative re-
sponse to o-nitrotoluene that is about two orders of magnitude smaller than
TABLE 6.7

PID/FID TOLUENE NORMALIZED RESPONSES (TNR) FOR VARIOUS COMPOUND CLASSES

(Reprinted with permission from ref. 33.)

Class Retenti on times < 17 min Retention times > 17 min


Number of TNR Number of TNR
species (mean ± S.D.) species (mean ± S.D.)
tested tested

Halogenated alkanes 9 o± 0 6 1 ± 1
Simple alkanes 13 3 ± 3 10 12 ± 4
Cycloalkanes + trimethylalkanes 2 3 ± 1 5 27 ± 8
Alkynes 3 3 ± 3 0
Alkenes 23 70 ± 11* 20 55 ± 10
Aldehydes 4 69 ± 10** 1 56
Ketones 2 157 ± 5 2 123 ± 25
Aromatics 0 21 87 ± 18
Chlorinated aromatics 0 7 141 ± 12
Chlorinated alkenes 3 218 ± 180 4 211 ± 97
Sulphur-containing hydrocarbons 2 500 ± 210 2 129 ± 37
*
**Does not include ethylene.
Does not include acetaldehyde.
125

9 10 16

13

PIO
7 11 FlO

16

o 5 10 . 15
min
o 5 10 min 15

Fig. 6.11. Drug chromatograms. Glass column (6 ft. x 2 mm I.D.) packed with 3%
OV-17 on Gas-Chrom Q deactivated with phosphoric acid. 1 = Barbital; 2 = Methy-
prylon; 3 = Aprobarbital; 4 = Allylisobutyl; 5 = Amobarbital; 6 = Pentobarbital;
7 = Secobarbital; 8 = Meprobamate; 9 = Doriden; 10 = Mephobarbital; 11 = Pheno-
barbital; 12 = Darvon; 13 = Methaqualone; 14 = Primidone; 15 = Dilantin; 16 a
Valium. Temperature programmed from 200 to 285°C at 12 0 C/min. (From ref. 31.)

TABLE 6.8

RELATIVE RESPONSE FACTOR RATIOS ON ECD/PID FOR POLYAROMATIC HYDROCARBONS AND


THEIR NITRO DERIVATIVES
(From ref. 35.)

Compound pair PID/PID ECD/ECD ECD/PID:ECD/PID*

Indan/5-nitroindan 0.46 3.1.10- 5 6.80.10- 5


Fluorene/2-nitrofluorene 1.01 ** **
Naphthalene/2-nitronaphthalene 0.23 1. 70.1O-~ 7.41,10-6
Anthracene/9-nitroanthracene 0.38 2.70·10- 4.88,10- 3
Pyrene/3-nitropyrene 1.00 7.00,10- 3 6.99.10- 3

*Calculations made using peak heights. Relative response factor normalizations


using o-nitrotoluene = 1.00 on both detectors.
**Not possible to obtain any measurable ECD response for fluorene at microgram
levels or above.
126

PIO 26
23 24 29
2~ 27~\ )30
2122
18 1920~
61

1m _ill LL
FlO
11 15
12 14
13
8910
7

56
1

liUill
o 5 10 15 20 25 30 min

Fig. 6.12. Analysis of automobile exhaust. SE-30 silica capillary column (60 m).
1 = ethane; 2 = propane; 3 = isobutane; 4 = n-butane, 5 = 2-methylbutane,
6 = n-pentane; 7 = 2-methylpentane; 8 = 3-methylpentane; 9 = n-hexane;
10 = methylcyclopentane; 11 = 2-methylhexane; 12 = 2,2,4-trimethylpentane;
13 = n-heptane; 14 = methylcyclohexane; 15 = n-nonane; 16 = propylene;
17 = I-butene; 18 = trans- and ais-butenes; 19 = 1-pentene; 20 = 2-methyl-1-
butene; 21 = trans- and ais-pentenes; 22 = cyclopentene; 23 = benzene;
24 = toluene; 25 = ethyl benzene; 26 = m- and p-xylene; 27 = a-xylene; 28 = n-
propylbenzene; 29 a l,3,5-trimethylbenzene; 30 = 1,2,4-trimethylbenzene.
Temperature programmed from -50 0 C (2 min) to 100 0 C at 60 C/min. (Reprinted from
ref. 33 with permission.)

that to dinitrotoluene. The PIO response to dinitrobenzenes is so small that


levels below 1 vg are not detectable. The ECD and PIO response ratios charac-
terize all three groups (Table 6.9).

6.5. CARRIER GAS

Helium has been used as carrier gas because it does not absorb UV radiation
(10-eV photons). The PIO with nitrogen as the carrier gas (absorbing the UV ra-
diation)9 gives a signal similar to that with helium. In this instance, other
reactions can take place in addition to the set of reactions 6.2-6.8:
127

FlO
min

3 4
-
12
PIO min

Fig. 6.13. Gas chromatograms obtained with simultaneous use of an FID and a PID
of four nitro compounds. Glass column (6 ft. x 2.00 mm I.D.) of Ultrabond 20M.
1 = Nitropentane; 2 = nitrocyclohexane; 3 = nitrotoluene; 4 = 2,6-dinitrotoluene.
Temperature programmed from 50 0 C (2 min) to 180 0 C at lOoC/min; split, 70:30
between FID and PID. (From ref. 35.)

N2 + hv + N2 * (6.13)
N2* + AB + N2 + AB+ + e- (6.14)

where reaction 6.14 can be highly efficient. When using argon as the carrier gas,
the PID response is about 40% higher than that when nitrogen or helium is used
(Table 6.10). This is due to the higher drift velocity of the electrons in argon;
the efficiency of electron collection in the PID is higher and the recombination
of electrons and ions decreases (reaction 6.7)36. The response in hydrogen is
also by about 25% greater than that in nitrogen 10 • The response in air is lower
than that in helium 10 ,12,25, and which is due to the more efficient neutraliza-
tion of the positive ions from photoionization by anions compared with that by
electrons 12 (owing to the comparable velocities of the anions and cations). The
oxygen itself decreases the background ionization current (negative response)17:
128

TABLE 6.9

RELATIVE RESPONSE FACTORS FOR NITRO-AROMATICS OBTAINED BY GAS CHROMATOGRAPHY


WITH ECD AND PID

(From ref. 35.)

Compound Relative response factor *


PID ECD ECD/PID

o-Nitrotoluene 1.00 1.00 1.00


m-Nitrotol uene 1.06 1.11 1.05
p-Nitrotoluene 1.56 1.06 0.68
2,3-Dinitrotoluene 2.92 10- 2
0
5.44 186.3
2,4-Dinitrotoluene 9.00.1O-~ 4.86 540
2,6-Dinitrotoluene 2.46·10- 5.47 222.4
3,4-Dinitrotoluene 1. 78.10- 2 4.94 277.5
o-Dinitrobenzene -** 4.86 -**
m-Dinitrobenzene -** 3.31 -**
p-Dinitrobenzene -** 5.81 -**

*Relative response factors were obtained by measuring peak heights and dividing
by the absolute amount reaching the detector. o-Nitrotoluene was assigned an
arbitrary value of 1.00 cm/ng, and other relative response factors were cal-
culated relative to o-nitrotoluene.
**There was no measurable PID response for sub-microgram amounts of the dinitro-
benzenes.

TABLE 6.10

RELATIVE PID RESPONSES

(From ref. 36.)

Carrier gas Relative response


(flow-rate
15 ml/min) Benzene Toluene p-Xylene m-Xylene o-Xylene

Argon 1.401 1.44 1.49 1.40 1.42


Carbon monoxide 1.03 1.02 1.03 1.00 1.00
Tetrafluoromethane 1.010 0.998 0.848 1.098 1. 27
Nitrogen 1.000 1.000 1.000 1.000 1.000
Helium 0.987 0.987 1.126 0.970 0.982
Air 0.542 0.525 0.487 0.456 0.433
Carbon dioxide 0.285 0.269 0.278 0.249 0.179
129

e + O2 + O2 (6.15)
O - + AB+ + AB + O (6.16)
2 2

The response in carbon dioxide is about five times smaller than that in
helium, which, according to Freedman's explanation l2 , is because carbon dioxide
is much more efficient in the deactivation of collision-excited states (see
eqn. 6.8). Carbon monoxide and tetrafluoromethane do not lead to a decrease in
response and, therefore, Senum36 ascribed the lower responses in carbon dioxide
to electron capture, similarly to the case with oxygen (eqns. 6.15 and 6.16).
Table 6.9 lists the relative responses in various carrier gases. A response de-
crease similar to that in air and carbon dioxide has also been found for methane
and nitrous oxide. The absorption coefficient at 10.2 eV is very high for these
two gases (Table 6.11). The responses to the analysed compounds in nitrogen,
helium, carbon monoxide and tetrafluoromethane are commensurate 36 . It is evident
from the above that the use of argon as a carrier gas is advantageous, whereas
the use of methane is disadvantageous (for instance, for the connection of a
PIO and an ECO).

TABLE 6.11

IONIZATION POTENTIALS AND ABSORPTION COEFFICIENTS FOR VARIOUS CARRIER GASES

(From ref. 36.)

Carrier gas Ionization Absorption coefficient


potential at 10.2 eV
(eV) (cm- 1 )

Oxygen 12.063 0.27


Nitrous oxi de 12.894 100
Carbon dioxide 13.769 2
Methane 12.6 400
Tetrafluoromethane 15.67 Negligible
Carbon monoxide 14.013 3
Argon 15.67 Negligible
Helium 24.58 Negligible

In the PIO in which the UV radiation source is not separated from the ioniza-
tion chamb~r and where photoionization is not the only reaction to occur (see
section 6.1), the response decreases in the order He > N2 > Ar > air. Here, the
type of carrier gas also affects the background ionization current and the de-
tector noise 7.
130

TABLE 6.12

CHROMATOGRAPHIC EFFICIENCY (N) AS A FUNCTION OF FLOW-RATE USING THE FID AND


THE PID

(From ref. 37.)

Carrier gas Gas velocity N Di fference


flow-rate (em/sec) in N
(ml/min) PID FlO (%)
(30 m) (30 m)

0.32 5.16 86 917 102 389 17 .8


0.52 7.13 76 990 97 720 26.9
0.91 10.77 47 577 68 182 43.31
1.60 15.23 28 972 45 455 56.89
2.13 18.74 20 508 33 520 63.45

A B

o 5 10 min 0 5 10 min

Fig. 6.14. Comparison of the efficiencies of chromatographic columns with (A) a


PID and (B) an FID. (From ref. 37.)
131

The PID is a concentration detector and therefore the detector response


(peak area) decreases with increasing flow-rate of the carrier gas 9• If the car-
rier gas flow-rate is varied from 10 to 30 ml/min, the ppak area diminishes to
25% or 20% of the initial value 9 ,17,22. Hence the response is lower when using
make-up gas with capillary columns in connection with a PID. For instance, the
PID response increases by about 50% if the flow-rate of the make-up gas is
changed from 30 to 20 ml/min 33 • The optimization of the HNU Systems PID for
capillary columns was described by Jaramillo and Driscol1 37 (for applications
of PIDs with capillary columns see, e.g., refs. 20, 32, 33, 38 and 39). The de-
terioration of the chromatographic efficiency compared with the application of
an FID is a function of the flow-rate of the carrier gas (see Table 6.12). The
decrease in peak separation is very small for low flow-rates (see Fig. 6.14, the
group of peaks with braces). I~hen using make-up gas with a flow-rate of 30 ml/min,
the peak width at half-height is reported to be very similar for both detectors 33 .
The smallest detector volumes vary20,24,32 from about 35 to 50 ~l, the effective
volume of the 225-~1 unmodified HNU Systems detector being only 54 ~l at a flow-
rate of 1.6 ml/min 37 •

REFERENCES

1 HNU Systems, New High Tempepatupe Photoionization Detectop fop Gas Chpomato-
gpaphy, Industrieregler GmbH, Vienna, 1978.
2 J.E. Lovelock, Natupe (London), 188 (1960) 401.
3 t1. Yamane, J. ChPOmatogp., 11 (1963) 158.
4 M. Yamane, J. Chpomatogp., 14 (1964) 355.
5 J.F. Roesler, Anal. Chem., 36 (1964) 1900.
6 D.C. Locke and C.E. Meloan, Anal. Chem., 37 (1965) 389.
7 J.G.W. Price, D.C. Fenimore, P.G. Simmonds and A. Zlatkis, Anal. Chem., 40
(1968) 541.
8 R.R. Freeman and W.E. Wentworth, Anal. Chem., 43 (I971) 1987.
9 J.N. Driscoll and F.F. Spaziani, Res./Develop., 27, May (1976) 50.
10 J. Sevc1k and S. Kr~sl, Chpomatogpaphia, 6 (1973) 375.
11 S. Kapila, D.J. Bornhop, S.E. t1anahan and G.L. Nickell, J. ChPOmatogp., 259
(1983) 205.
12 A.N. Freedman, J. Chpomatogp., 190 (1980) 263.
13 A. Schweig and W. Thiel, J. Chem. Phys., 60 (1974) 951.
14 J.N. Driscoll, J. Ford, L.F. Jaramillo and E.T. Gruber, J. Chpomatogp., 158
(1978) 171.
15 K. Watanabe, J. Quant. Spectposc. Radiat. Tpansfep, 2 (1962) 369.
16 N. Ostojic and Z. Sternberg, Chpomatogpaphia, 7 (1974) 3.
17 J.N. Driscoll, J. Chpomatogp., 134 (1977) 49.
18 M.L. Langhorst, J. Chpomatogp. Sci., 19 (1981) 98.
19 A.N. Freedman, J. Chpomatogp., 236 (1982) 11.
20 J.N. Davenport and E.R. Adlard, J. Chpomatogp., 290 (1984) 13.
21 M.E. Casida and K.C. Casida, J. Chpomatogp., 200 (1980) 35.
22 Instpuction Manual, Model PI-52-02 Photoionization Detectop, HNU Systems,
Newton, MA, 1979.
23 J.N. Driscoll, Amep. Lab., 8 (1976) 71.
24 Tpacop Model 703 Photoioniaation Detectop, Gas Chpomatogpaphy, Tracor
Instruments, Austin, TX.
132

25 R.C. Leveson and N.J. Barker, in Proceedings of the 27th Annual ISA Analysis
Instrumentation Symposium. St. Louis. MO. March 1981. Instrument Society of
America, Research Triangle Park, NC, p. 7.
26 J.N. Driscoll, Ind. Hygiene News. 3, No. I, March (1980).
27 J.N. Driscoll, J. Chromatogr. Sci., 20 (1982) 91.
28 J.N. Driscoll, J. Ford, L. Jaramillo, J.H. Becker, G. Hewitt, J.K. Marshall
and F. Onishuk, Amer. Lab •• 10 (1978) 137.
29 D.B. Smith and L.A. Krause, Amer. Ind. Hyg. Ass. J., 39 (197B) 939.
30 M. Dressler, J. Chromatogr •• in preparation.
31 L.F. Jaramillo and J.N. Driscoll, J. Chromatogr., 186 (1979) 637.
32 S. Kapila and C.R. Vogt. J. High Resolut. Chromatogr. Chromatogr. Commun.,
4 (l981) 233.
33 R.D. Cox and R.F. Earp, Anal-. Chem •• 54 (1982) 2265.
34 J. Winskowski. Chromatographia, 17 (1983) 160.
35 J.S. Krull, M. Swartz, R. Hilliard, K.H. Xie and J.N. Driscoll, J. Chromatogr.,
260 (1983) 347.
36 G.I. Senum, J. Chromatogr., 205 (1981) 413.
37 L.F. Jaramillo and J.N. Driscoll, J. High Resolut. Chromatogr. Chromatogr.
Commun., 2 (1979) 536.
38 W.G. Jennings, S.G. Wyllie and S. Alves, Chromatographia, 10 (1977) 426.
39 J. Meili, P. Bronnimann, B. BrechbUhler and H.J. Heiz, J. High Resolut.
Chromatogr. Chromatogr. Commun., 2 (1979) 475.
133

ChapteY' 7

FLAME PHOTm1ETRI C DETECTOR

CONTENTS

7.1. Introduction • • • • • • • • • • • • • • • • • 133


7.2. Response mOdel • . • • • • • . • • • • • . • • • 136
7.3. Detector sensitivity and minimum detectability 137
7.3.1. Flow-rates of gases •••• 137
7.3.2. Structure of compounds •• 138
7.3.3. Concentration of compounds 139
7.3.4. Detector temperature 142
7.3.5. Interference filter. 142
7.3.6. Photomultiplier ••• 143
7.4. Selectivity of response •• 144
7.5. Tin and germanium compounds 145
7.6. Halogen compounds • • • • . 147
7.7. Other detection possibilities 149
7.8. Linearity of response 150
7.9. Su 1phur background • • • • • . 152
7.10. Response quenching • . • • • • 152
7.11. Flame stability • • • • • • • • • 157
7.12. Other identification possibilities 157
References • . • • 158

7.1. INTRODUCTION

In an oxygen-rich flame, the decomposition of substances that contain a hetero-


atom generates excited species that, during their transition to the ground state,
emit radiation characteristic of the given heteroatom. The principle of the flame
photometric detector (FPD) is based on the measurement of the characteristic
emission in the flame. In gas chromatography, a flame photometric detector for
sulphur and phosphorus compounds has found the widest use and has also been pro-
duced commercially.
As early as in 1869, Salet 1 found that, on introduction of an aerosol of a
solution of a sulphur compound into a low-temperature, fuel-rich hydrogen flame,
an intense blue emission near the surface of the cold object near the flame core
could be observed. Salet's phenomenon became the basis for the technique of de-
tecting sulphur and/or phosphorus compounds by means of a flame photometric de-
tector in a hydrogen-rich flame 2• Dagnall et al. 3 and Syty and Dean 4 developed
this technique for cool flames in flame emission spectrometers, while Brody and
Chaney5 developed it for gas chromatography (Juvet and Durbin 6 for organometallic
compounds in 1963).
134

..... COLUMN
EFFLUENT(N2)

Fig. 7.1. Schematic diagram of flame photometric detector. 1 = Flame-ionization


burner tip; 2 = burner; 3 = mirror; 4 = glass window; 5 = optical filter;
6 = photomultiplier tube. (Reprinted from ref. 5, with permission.)

<0
N
a If)

490 600
WAVELENGTH (nm )

WAVELENGTH ( nm I

Fig. 7.2. Emission spectra of (a) HPO species, slit width 0.3 mm (from ref. 4);
and (b) S2 species, slit width 0.4 mm (from ref. 7).

The design of an FPD for use with a gas chromatograph to detect phosphorus
and/or sulphur compounds is illustrated in Fig. 7.1. The carrier gas at the
chromatographic column outlet is mixed with oxygen while hydrogen is fed directly
135

1
2
PHOSPHORUS
4
3
6
START

~A
x 64 000 1mV
10 min

L ~

START
x 6 400 1mV
r V
5 t
10 min

4
2 6

SULPHUR

Fig. 7.3. Simultaneous detection of phosphorus- and sulphur-containing compounds.


1 = disyston; 2 = methyl parathion; 3 = malathion; 4 = parathion; 5 = methyl
trithion; 6 = ethion. Glass column (6 ft. x 0.16 mm 1.0.) packed with 3% OV-101
on Gas-Chrom Q; temperature, 190 0 C. (From ref. 9.)

A B C 0 E

II II II II
H'~ I /
H/N2 H/AIR
/
N2+AIR
/

t
/
O2/ AIR
I
t I
02+ N2 N2 H2 N2+ H2
°2

Fig. 7.4. Schematic diagram of the gas supply lines leading to the FPD. A, From
ref. 5; B, from ref. 10; C, from ref. 11; 0, from ref. 12; E, from ref. 13.

to the detector body (see also Fig. 7.4A). The compounds decompose in the de-
tector flame giving rise to excited HPO * species (phosphorus compounds) or S2 *
species (sulphur compounds); the spectra of the emitted radiation of the above
compounds are displayed in Fig. 7.2. From the given spectra, wavelengths of
about 526 nm for phosphorus and about 394 nm for sulphur are selected for detec-
136

tion by the FPD. The appropriate interference filter is placed between the
emission chamber of the FPD and the photomultiplier tube. The light emission is
viewed by the photomultiplier tube.
The original construction by Brody and Chaney5, equipped with a single optical
system (526 or 394 nm), was complemented by Bowman and Beroza 8 by an additional
system that made it possible to apply phosphorus and sulphur filters simulta-
neously. At present, this construction is currently also used with commercial
detectors. It allows two simultaneous selective records to be obtained from one
chromatographic analysis (Fig. 7.3).
In addition to the gas inlet .system shown in Fig. 7.1, many other configura-
tions of the FPD have been used (Fig. 7.4). Their potential advantages are dis-
cussed in the respective sections.

7.2. RESPONSE MODEL

The flame in the FPD serves three basic functions in the sulphur response
mechanism 14 ,15:
(a) the initial sulphur-containing molecules are decomposed in the hot regions
of the fl arne;
(b) the sulphur species formed then produce sulphur atoms, either directly:

heat
sulphur compound S atoms (7.1)

or indirectly:

heat
sulphur compound H2S (7.2)

H2S + H HS + H2 (7.3)

HS + H .......-
---'"
S + H2 (7.4)

(c) subsequently, excited S2* species are formed in the cool outer cone. The
excitation energy for S2 is believed to come from three-body recombinations 16 ,17:

S + S + r1 - S2 * + 11 (7.5)

(where r1 is the thi rd body) with the correspondi ng emi ssi on:

(7.6)

Syty and Dean~ believe that the excitation energy for the S2 ~ S2* transition is
due to recomb.ination reactions:
137

(7.7)

or

H + OH + S2 (7.8)

The contribution of reactions 7.7 and 7.8 to the total emission of 52 * species
is thought to be small, however 16 ,17.
The following set of reactions has been proposed for the phosphorus response
mechanism 4:

P + 0 P + PO (7.9)
2

P + OH PO + H (7.10)

H + PO + /·1 ~ HPO * + r~ (7.11)

H + PO ~ HPO * (7.12)

7.3. DETECTOR SENSITIVITY AND MINIMUM DETECTABILITY

The detector response to phosphorus and sulphur compounds is a function of


(1) gas flow-rate; (2) structure of the solute compounds; (3) concentration of
the solute compound in the carrier gas; (4) detector temperature; (5) filter
application; and (6) photomultiplier.

7.3.1. Flow-rates of gases

The dependence of the detector sensitivity and minimum detectability on the


gas flow-rates varies according to the detector design and the configuration of
the conduit leading the gas into the detector 17 ,18-23, and it is difficult to
generalize. Evidently, optimum flow-rates have to be found for each detector
(Table 7.1) and there are substantial differences in some instances. For in-
stance, if the carrier gas is fed outside the jet, the response decreases with
increasing air flow-rate 18 , as with the type E configuration of gas inlets (see
Fig. 7.4)17. For the type C configuration, the response increases 18 with in-
creasing air flow-rate or displays a maximum 24 in the same way as with a detector
with a separate emission chamber 25 • With hydrogen, the dependence of the response
on the increasing flow-rate of this gas can also decrease 17 ,24, increase 25 or
reach a maximum 21 ,23. All these interrelations are affected by the 02:H2 flow-
rate ratio (e.g., refs. 20,23 and 25) and by the type of carrier gas used 23,26
(nitrogen or helium). The minimum detectabilities with various types of FPD are
approximately commensurate at optimum gas flow-rates; however, the gas inlet con-
138

TABLE 7.1

SUMMARY OF OPT! MUr1 FLAME PARAMETERS

Reprinted from ref. 21, with permission.

Source Flow-rate (ml/min)

H2 O2 Air N2 o /H*
2 2

Brody and Chaney5 200 40 0 160 0.20


Mi zany 22 50 20 10 70 0.44
Greer and BYdal~~33 90 18 15 30 0.25
Sugiyama et al. 168 0 195 22 0.23
Pes car and HartT~nn24 80 0 100 0 0.25
Eckhardt et al. 100 24 0 126 0.24
Burgett and Gre g13 50 10 50 60 0.44
Kapila and Vogt2 60-80 6-8 0 0.10

* Including oxygen from the air.

figuration in which the carrier gas is pre-mixed with hydrogen seems to offer
certain advantages. The minimum detectable mass rate ranges from about
1'10- 13 g/sec 20 ,26-28 to 2'10- 12 g/sec 5 ,9,29,30 of P for phosphorus compounds
and from about 2'10- 12 g/sec 20 ,27,29,31,32 to 5'10- 11 g/sec of S28,30-32 for
sulphur compounds.

7.3.2. Structure of compounds

Studies on the effects of the structure of sample compounds on the FPD re-
sponse have produced different results. According to Maruyama and Kakemoto 18 ,
the detector response to sulphur compounds depends solely on the sulphur con-
tent in the molecule of the compound. The response is constant for equal amounts,
of sulphur, i.e., it is independent of the structure of the compound. However,
most workers have demonstrated that the response and, hence, the minimum detect-
ability depend on the structures of both sulphur compounds 19 - 22 ,28,34 and phos-
phorus compounds 20 ,35,36. This means that the response depends substantially on
the efficiency of the production of the emitting HPO* and S2* species. Studies
on the intensity of this influence have led to contradictory conclusions: Mizany22
reported that the response decreases in the order disulphide > sulphate> sul-
phone> sulphide> sulphite, whereas Sugiyama et al. 34 reported the order
buthanethiol > thiolane > thiophene> butyl sulphide > thiophenol > benzothio-
phene > phenyl sulphide> pentyl disulphide. In the former instance disulphide
139

gives a greater response than sulphide, whereas in the latter the contrary is
true. The response decreases in the order 36 phosphane > phosphite> phosphate
phosphate for phosphorus compounds (see, for instance, triphenyl compounds in
Table 7.2). For compounds that contain both a phosphorus and a sulphur atom in
the molecule, the response ratio Rp/RS is given 8 ,9,30 by the phosphorus to
sulphur ratio in the molecule of the compound. The Rp/IRS ratio is recommend-
ed 8 ,9,30 for calculating the phosphorus and sulphur contents. As n (see sec-
tion 7.3.3) does not alway equal 2, and in view of the non-constancy of n, this
relationship cannot be used generally.

7.3.3. Concentpation of compounds

The dependence of the FPD response on the amount of a phosphorus compound is


linear over a range of about four orders of magnitude. Thus, the response per
unit amount of solute (molar response, weight response) is constant within this
range. The dependence of the response on the amount of a sulphur compound is
given by the equation R = k[S]n. As S2 species are responsible for the emission,
the response should theoretically be linearly related to the square of the amount
of sulphur compound. Measurements exist that support this assumption. Maruyama
and Kakemoto 18 reported n a 2, irrespective of the structure of the compound.
Several experimental results have shown, however, that n varies between 1 and 2
(e.g., refs. 20, 31, 34, 37 and 38). In addition, it changes with the hydrogen
flow-rate and the 02:H2 and 02:N2 flow-rate ratios 17 ,19,22, with the structure
of the sulphur compounds 19 ,21,22,34 and with the detector temperature 39 . An ex-
ponential relationship between the response and the amount of the compound was
also found for selenium 40 •41 and tellurium 41 compounds.
Hence the response per unit amount of solute increases with increasing amount
of the solute compound in any case. However. if n varies with the structure of
the compound, the relative responses of different compounds can vary with chang-
ing concentration. Fig. 7.5 demonstrates this effect for dimethyl disulphide
(n = 1.78) and diphenyl sulphide (n = 1.96). With solute supply rates exceeding
about 0.4 ng/sec of S, the response for sulphide is greater than that for di-
sulphide, whereas the disulphide response is greater than the'sulphide response
below this rate. The dependence of n on the structure of the compounds may be re-
sponsible 21 for the discrepancies that occur in the literature for the response
level of individual types of sulphur compounds (see section 7.3.2). Sugiyama et
al. 34 , who found that the detector response to sulphides was greater than that
to disulphides, worked in the range 16-32 ng/sec of S, whereas Mizany22 re-
ported the disulphide response to exceed that to sulphides. When recalculating
the investigated amounts of compounds with regard to the sample charge and the
......
TABLE 7.2 ~
o

RESPONSE OF ORGANOPHOSPHORUS COMPOUNDS IN SINGLE- AND DOUBLE-FLAME PHOTOMETRIC DETECTOR

Reprinted from ref. 36, with permission.

Compound Amount of Amount Single-flame Double-flame


compound of P response (R ): response (R ): 5
injected (ng) peak area x110-5 peak area x210-
(ng)

Tri ethyl phosphate 20 3.4 7.5 28.1 3.7


Tri-n-butyl phosphate 20 2.3 4.2 18.1 4.3
Triphenyl phosphate 20 1.9 2.0 7.2 3.6
Triethyl phosphite 20 3.7 21.1 31.1 1.5
Tri-n-butyl phosphite* 20 2.5 5.8 7.5 1.3
Triphenyl phosphite 20 2.0 11.1 14.1 1.3
Triethyl thiophosphate 20 3.1 28.1 27.9 1.0
Methyl parathion 20 2.3 16.1 16.2 1.0
Parathion 20 2.1 15.1 15.3 1.0
Triphenyl phosphine 20 2.4 18.2 19.4 1.1

*Showed impurity peak.


141

UI
Vl
Z
a
a.
Vl
w
cr:

0.01 0.1 1.0 10 100


ngS ISec

Fig. 7.5. Effect of structure of sulphur compounds on the dependence of the FPO
response on the amount of the compound. (Reprinted from ref. 21, with permission.)

carrier gas flow-rate, it becomes evident that Mizany actually applied about
0.04 ng/sec of S, so that both results appear to be correct. Other discrepancies
are also likely to be explained in this manner; for instance, Mizany22 noted that
the relative response varies with the concentration of the compounds for the
sulphite-sulphide pair. The sulphide response is greater for 5 ng of S whereas
the sulphite response is greater for 10 ng. As n varies with the flow-rate, the
relative response additionally also varies with the gas flow-rates at different
concentrations. For an 02:H2 ratio of 0.44 (2 ng of S) the disulphide response
is greater than the sulphate response and at an 02:H2 ratio of 0.15 the response
i.s the same.
142

7.3.4. Deteator temperature

Whereas the FPD response to sulphur compounds decreases with increasing de-
tector temperature 18 ,24,39, with phosphorus compounds it increases 39 • The charac-
ter of the variations in the response to sulphur compounds differs slightly for
individual types of detector. Whereas Maruyama and Kakemoto 18 found that the de-
pendence of wl'h (W is peak width at half height, h is peak height) on the de-
tector temperature was linear in the range 100-300 0 C with a decrease in the re-
sponse by about a third, Dressler 39 noted a decrease in the peak height by ap-
proximately two thirds in the range BO-160 0 C (see Table 7.3). Under certain ex-
perimental conditions. this decrease changes to an increase in response at de-
tector temperatures exceeding 160 0 C. The above relationships are also slightly
affected by the flow-rates of the gases 39 and by the structure of the compound 24 ,
39. However. the detector noise also increases exponentially with increasing
temperature of the FPD 39 • Hence the minimum detectable mass rate increases with
increasing detector temperature with sulphur compounds, whereas it remains approx-
imately constant with phosphorus compounds.

TABLE 7.3

PEAK HEIGHT OF THIOPHENE OBTAINED USING VARIOUS FPD TEMPERATURES

Amount injected: 1.50.10-10 g. From ref. 39.

TD Peak height Difference in peak heights (%)*


(OC) (arbitrary units) A B C

BO 136 +212.6 +B3.B


100 90 +106.9 +31.6 -33.B
120 74 +70.1 -45.6
140 47.5 +9.2 -35.B -65.1
160 43.5 -41.2 -68.0
1BO 51.5 +18.4 -30.4 -62.1
200 62 +42.5 -16.2 -54.4

*A, TD = 160 0 C taken as the basis; B, TD = 1200 C taken as the basis; C, TO = BOoC
taken as the basis.

7.3.S. Interferenae filter

Monitoring of the emission without using a filter yields a greater response


than that obtained when a filter is used 42 - 44 (Fig. 7.6). Under these conditions,
the selectivity for hydrocarbons is retained even though it is smaller, but the
selective resolution between phosphorus and sulphur disappears. As the noise
143

10 6

10 5

--
0 10 4

CII

.-
111
0 10 3
c

-
0

Ci 10
2 C'4 H30 ( 52 5 )
c C'2H26 (393 )
Ol

III
10 '

-1
10

Fig. 7.6. Relationships between the signal-to-noise ratios and the amounts
injected for filter and filterless modes. Closed symbols, 393 or 525 nm inter-
ference filters used; open symbols, no filters used. (From ref. 42.)

also increases, the minimum detectable mass rate is only slightly decreased 42 •
If the dimensions of the slit of the black disc placed in the detector instead
of the interference filter are optimized. the minimum detectable mass rate is
about half that obtained when a filter is used 45 •

7.3.6. Photomultiplier

The detector sensitivity and detector noise depend on the type of photo-
multiplier and on the voltage applied to the latter. The detector response in-
creases with increasing voltage applied to the photomultiplier. However, the
noise increases at the same time, so that the minimum detectable mass rate re-
mains approximately constant 24 •25 . The detector noise and. consequently, the
minimum detectable mass rate increase with increasing temperature of the photo-
multiplier tUbe 46 .
144

7.4. SELECTIVITY OF RESPONSE

The selectivity of the detector response for phosphorus (interference filter


of about 394 nm) is about four to five orders of magnitude relative to hydro-
carbons. With a detector in which the emission space is separated from the com-
partment in which combustion takes place, a selectivity as high as seven orders
of magnitude is reported 25 ,47. Filters with slit widths as narrow as possible
should be used in order to obtain minimum interference between sulphur and
phosphorus emissions. As the FPD response d~pends on the amount of the compound
present in the detector, the selectivity for sulphur increases with increasing
concentration of the compound. Hence the selectivity for sulphur compounds has
to be expressed for the minimum detectable mass rate. Owing to the different de-
pendences of the responses to phosphorus and sulphur compounds on concentration,
the S:P response ratio is also a function of the concentration of the substances
when a single filter is used. When using an S filter the S:P response ratio is
100:1 to 1000:1 and when using a P filter the S:P response ratio is 4:1 for
amounts of 100-200 ng 29 ,30. For a detector with separated combustion and emission
compartments, Joonson and Loog47 described the possibility of compensating elec-
trically for the interfering sulphur response in the P channel; part of the
signal from the S channel, corresponding to the interfering sulphur signal in
the P channel, is fed to the output of the P channel amplifier (Fig. 7.7). Hence
the response to the phosphorus compounds only is recorded.
As the detector response to hydrocarbons is approximately constant at various
temperatures of the detector, the selectivity varies with the detector tempera-
ture in the same way that the peak height of the respective compound varies de-
pending on the temperature of the detector 39 • The P:S response ratio then in-
creases with increasing detector temperature. The selectivity of the FPD response
for sulphur and phosphorus compouns is also a function of the gas flow-rate.

WP

Fig. 7.7. Dual-channel FPD system with S-interference compensator. 1 = FPD;


2 = dual-channel photometric block; 3 = light conductor; 4 = light divider;
5 = interference filters; 6 = photomultipliers; 7 = electrometric amplifiers;
8 = signal divider; 9 = separating network; 10 = recorders. (From ref. 47.)
145

This is also due to the varying dependence of the hydrocarbon response and the
sulphur and/or phosphorus response on the flow-rates of the gases. For instance,
with air flow-rates in the range 35-92 ml/min, the SIC selectivity decreases by
a factor of about 100 with detectors having separated combustion and emission
compartments 39 •

7.5. TI NAND GERt1AN IUM COMPOUNDS

In a cool hydrogen-air diffusion flame, tin compounds produce a red molecular-


band emission 48 with a sharp SnH peak at 609.5 nm. With the aid of a Shimadzu
FPD with a modified quartz tube containing the flame gases (chimney) and using
a 610-nm interference filter, about 0.1 ng of tetrapropyltin could be determined
by gas chromatography44. The application of a filterless technique (see section
7.3.5) makes it possible to determine about 30 pg of a tin compound 49 ,50. A de-
tector of special design 51 has a detection limit of approximately 10 pg when
using an interference filter. Carbon dioxide and arsine give rise to response
interference due to the CO and CO+ molecular band emissions and AS 2+ molecular
band emission. The more sensitive tin luminescence produces a broad emission
ranging from 360 to 490 nm and appearing on the quartz surface (surface lumines-
cence 44 ,49,50,52). t10nitoring of this emission without using a filter provides
for the sensitive and selective detection of tin. The detection limit is about
2.10- 13 _4'10- 14 g for tetrapropyl tin; the Sn:C selectivity is 10 4:1 to 10 5:1
and the Sn:S and Sn:P selectivity is 10 2:1 to 10 3:1. The emission is considerably

GAS PHASE
(py rex tube)

i
i
-12
i
-11 -9 -8 -7 -6
LOG GRAMS INJECTED

Fig. 7.8. Response of tetrapropyltin by two different mechanisms. (Reprinted


with permission from ref. 50.)
146

Surface

phase

att :
I I I
40 80 120 • C

Fig. 7.9. Chromatogram of a standard mixture in surface- and gas-phase modes.


S, 3 ng of (tert,-C4H9)2S2; Sn, 100 pg of (n-C3H7)4Sn; C, 5 pg of n-C15H32;
Ge, 100 pg of (n-C4H9)4Ge. (Reprinted from ref. 49, with permission.)

affected by the shape of the quartz chimney in which the flame is burning 44 ,49,53.
The peaks or the compounds are tailed and peak tailing is increased by the pre-
sence of silicon 44 (the response also increases, however), sulphur 45 and hydro-
carbons 44 in the flame and increases with increasing diameter of the quartz
tube 44 • A decrease in the response was reported 54 ,55 with injections of large
amounts of organotin ~ompounds (more than 100 ng), tropolone as the extracting
solvent and organic co-extractives.
In the most advantageous design of the conventional detector a piece of quartz
wool is placed above the detector flame 50 • Fig. 7.8 compares the responses of
both types of detection, i.e., monitoring either surface luminescence or gas-phase
luminescence with tin compounds. Fig. 7.9 compares both methods of detection
for sulphur compounds, germanium compounds, and hydrocarbons (filterless mode).
With the surface mode the minimum detectable mass rate for tin compounds is ap-
proximately two orders of magnitude smaller, but the linear dynamic range is
smaller 50 ,52. The sulphide dominates the gas phase while the response is small
in the surface mode. Tin and germanium compounds show the opposite behaviour.
147

TABLE 7.4

FPD SELECTIVITY FOR GERMANIUr1 COMPOUNDS

From ref. 53.

Heteroatoms Selectivity
compared
Surface Gas phase

Ge:Sn 1:5 100:1


Ge:S 50:1 - 500:1 10:1 - 1000:1
Ge:P 17'1
Ge:C ca. 5'10 4 :1 106:1

A situation similar to that with tin is also observed with germanium. Surface
luminescence is the most sensitive emission mode for organic germanium compounds.
The minimum detectable mass rate of tetrabuty1germanium is 8.10- 15 g/sec of Ge 53 ;
the selectivity of germanium relative to other elements is obvious from Table
7.4. The chromatographic peaks are broadened. however (even more than with tin
compounds).
Monitoring of the gas-phase luminescence (650 nm) yields less sensitive detec-
tion (2'10- 13 g/sec of Ge). The peaks are not broadened, however, the selectivity
of germanium relative to tin and hydrocarbons increases (see Table 7.4) and the
linearity of response covers about three orders of magnitude 53 •

7.6. HALOGEN COMPOUNDS

Halogen compounds combusted in the presence of indium produce an intense


emission of indium ha10genides in the range 340-420 nm 56 • By positioning indium
near the flame of the detector and using a suitable interference filter (359.9 nm
for InC1, 372.7 nm for InBr and 409.9 nm for InI), a detector selective to halogen
compounds (except fluorine) is obtained 56 - 61 • A dual-flame detector was found to
be the most suitable design 60 ,62 (see Fig. 7.10). The indium source is positioned
between the two flames. The decomposition products containing C1, Br, I react
with indium giving halogenides, the emission of which is monitored in the upper
flame. The design according to Fig. 7.10 also provides for simultaneous photo-
metric detection 62 ,63 (from the lower flame) of phosphorus compounds (by means
of a 526-nm filter) and sulphur compounds (by means of a 394-nm filter). Similar-
ly to the above-mentioned modes of flame photometric detection, the response is
highly dependent on the flow-rates of the gases. The minimum detectable amount
148

with the dual-flame detector is 1.10- 11 g/sec for chlorine compounds, and the
linearity covers four orders of magnitude. Compared with the single-flame de-
tector, the sensitivity of this detector type is increased 58- 61 , owing to the
separated electrical heating of the indium source. The detector selectivity is
1000:1 to 10 000:1 relative to hydrocarbons 60 ,61 and about 100:1 relative to
sulphur and phosphorus compounds 60 ,62.
The system of the FPD sensing the sodium emission is based on the principle
of the alkali flame-ionization detector (AFID). A constant amount of an alkali
metal salt is introduced into the flame. The Na emission (589 nm) is either in-
creased or decreased in the presence of halogen compounds (except for fluorine).
The character and the level of the detector response, as with the AFID, are
basically affected by the hydrogen flow-rate (the flame temperature)64-67.
Starting from a certain hydrogen flow-rate, the basic Na emission seems to de-
crease, even though several workers reported solely a decrease in basic emis-
sion 64 ,68 in the range of hydrogen flow-rates examined, or, on the contrary,
solely an increase in basic emission 67 • The differing results can be ascribed
to the fact that the detector design plays an important role, as with the AFID,
and, particularly, the way in which the source of the alkali metal salt is posi-
tioned in the detector. The molar response decreases 64- 66 in the order I > Br > Cl

filter 360 nm.CI-channel quartz tube

indium
__ heater
photomultiplier -t:t=::.,
r upper flame air inlet
h 0 u sin g =:::~~ 1:;""",__-

flame shielding
glass window

filter, 394 nm. 5- channel

filter, 526 nm,P-channel

mix ed
hydrogen-air inlet

column

Fig. 7.10. Dual-flame photometric detector for chloro compounds. (Reproduced


from ref. 62, with permission.)
149

compounds, even though Bowman et al. 66 reported the order Br > Cl > I compounds.
The minimum detectable mass rates of iodine compounds vary in the range from 66
1'10-9 t0 65 2'10- 11 g/sec and the minimum detectable mass rate is about 1'10- 10
g/sec for Br and Cl compounds. Depending on the experimental conditions, the
detection selectivity relative to hydrocarbons ranges 65 ,67 from 5000:1 to
10 000:1 and is 100:1 relative to phosphorus compounds 65 •
The responses of the sodium FPD and the AFID display many similar features.
As mentioned in section 3.9, one of the theories dealing with the mechanism of
the function of the AFID ascribes the response of the latter to the increased
volatility of the sodium and/or an other alkali metal salt in contact with the
flame 68 • The response of both detectors depends strongly on the hydrogen flow-
rate (flame temperature) and on the detector design. With both detectors. the
response to halogen compounds can be positive (increased ionization current or
emission) or negative (decreased current or emission), being a function of the
experimental conditions, particularly the hydrogen flow_rate 65 •67 ,69-72. The
hydrogen flow-rate at which the positive response begins to decrease depends on
the cation applied 66 ,69,71 and on the heteroatom-in the molecule of the com-
pou~d67,69,71,72. However, a direct relationship between the decrease or increase
in the emission in the FPD and the decrease or increase in the ionization current
in the AFID cannot be derived from these results. For instance, the Na emission
is decreased (the response is negative) when the AFID response is positive, but
it remains negative also when the positive AFID response is decreasing (with the
hydrogen flow-rate)64 or when it is negative already70.
The Beilstein test for halogens is utilized with the detector in which the
flame emission (526 nm) is sensed in the presence of copper 73 •74 • The detection
limit is about 1'10- 8 g for halogen compounds 74 •
For the detection of fluorine compounds an FPD in which calcium in argon is
introduced into an acetylene-oxygen flame has been described 75 . The CaF emission
is sensed.

7.7. OTHER DETECTION POSSIBILITIES

Selenium compounds display a dominant Se 2 emission spectrum between 450 and


500 nm. The minimum detectable mass rate for selenium compounds is 2.10-12 g/sec
of Se with a 484-nm interference filter, The selectivity relative to C varies 40
between 1000 and 10 owing to the exponential dependence of the selenium compound
response on concentration.
Of other heteroatoms, arsenic compounds (3.10- 11 g)26,42, boron compounds
(3'10- 10 g)76 and Cr 77 and Fe 42 ,78 compounds (1'10-9 g) can be detected by the
FPD in amounts below 1 ng. The remaining heteroatoms can be detected in amounts
larger than 1 ng (Sb, Pb, Bi, Ni, Hg, Cu, Ti, Zr, Rh, W, Al)6,79-81.
150

Fig. 7.11. Flame-ionization detector modified for simultaneous FPO detection.


1 = Photomultiplier tube; 2 = interference filter; 3 = quartz window; 4 = de-
tector jet tip assembly; 5 = GC column; 6 = collector electrode. (From ref. 61.)

The positioning of two electrodes in the FPO system makes it possible to


monitor the ionization current of the detector 30 ,82 and to obtain another non-
selective record of the compound. On the other hand, a commercial flame-ioniza-
tion detector (FlO) can simultaneously be operated as an FP061 ,78,83, as can be
seen from Fig. 7.11. As the optimum operating conditions differ for the two de-
tectors, an FPO operated under FlO conditions yields a lower response.

7.8. LINEARITY OF RESPONSE

As noted above, the dependence of the FPO response on the amount of the phos-
phorus compound is linear, whereas this dependence for sulphur. selenium and
tellurium compounds is exponential. If a constant amount of a sulphur compound
is fed into the flame to create a background, the detector response becomes
linear also for sulphur, selenium and tellurium compounds 23 ,27,41 within a
certain concentration range. The effect of the amount of the sulphur compound
supplied on the linearity range is evident from Fig. 7.12. The linearity range
covers about two orders of magnitude (the sulphur background should be suffi-
ciently high in comparison with the peak height).
151

.!1
'c
::J
>.
<-
a<- A ODED
SULPHUR
.0 BACKGROUND
<-
a .4
t-
-.
3

-
I
<!l 2

-
UJ
I
::.::
,
oCt
UJ
D- n- C1s H38
O
<!l --. 3
0 2
...J

o o•
- 11 -10 -9 -S -7 -6 -5 - 4
LOG (GRAMS INJECTED)

Fig. 7.12. Calibration graphs on different sulphur backgrounds. No interference


filters used. Octadecane peaks inverted. (From ref. 41.)

The non-linearity of the response to sulphur compounds as a function of the


amount of the compound can be compensated for with the aid of an electronic
linearization devide 19 that converts the response to a linear function of the
sulphur concentration rather than the normal relation R = k[S]n. Commercial
linearization devices are commonly based on the quadratic dependence of the de-
tector response on the amount of the compound (n = 2). As mentioned in section
7.3, n does not always equal 2 and, in addition, it varies as a function of the
flow-rates of the gases, the structure of the compound and the detector tem-
perature. The errors due to this device that arise during quantitative analysis
increase with increasing concentration of the sulphur compounds and with de-
creasing n, varying between 40 to 200% at the tens of nanograms leve1 37 . For
this reason, the linearization device should provide a variable exponential
functi on 19.
152

7.9. SULPHUR BACKGROUND

Owing to the exponential dependence of the FPD response to sulphur compounds


on the amount of the compound, the response per unit amount of solute is the
lowest for the smallest amounts of the compound. This is disadvantageous from
the point of view of trace analysis. The supply of a certain constant amount of
a sulphur compound to the flame (commonly sulphur dioxide) increases the back-
ground of the sulphur emission. Therefore, the detector response is higher at
this increased background 12 ,27,84. The relationship between the response level
and the amount of the sulphur compound supplied (the emission background) is
apparent from Fig. 7.12. The noise also increases with increasing detector re-
sponse and, therefore, the minimum detectability increases by a factor of only
5-10. Of course, the ratio of the responses without and with the sulphur emis-
sion background is higher for small amounts of sulphur compounds. For the surface
mode (see section 7.5), carbon disulphide doping also results in an increased
response to sulphur, selenium and tellurium compounds. The tin and germanium
responses are unaffected, however, except that peak tailing worsens. A similar
increase in response can also be obtained by doping the flame with selenium or
tellurium compounds 49 •
It should be realized that in trace analysis the unwanted presence of a
sulphur compound in the flame (detector contamination, column bleeding) can lead
to erroneous results owing to the effect mentioned above.
As mentioned later (section 7.10), sulphur-free organic compounds decrease
the detector response to sulphur compounds. Hence, if a sulphur compound is
supplied constantly into the flame, the signal decreases during hydrocarbon
elution 12 ,13,27,46,47,84. When using this mode, the minimum detectable mass rate
of hydrocarbons is about 10- 7 g/sec 45 •

7.10. RESPONSE QUENCHING

The presence of a volatile sulphur-free compound simultaneously with sulphur


compounds in the cool flame results in a decrease in the FPD response 3 ,23,38 and
in the FPD response 7 ,10,18,25,32,40,44,46,84-86. The decrease in the response
occurs regardless of the type of interfering compound (hydrocarbons, alcohols,
aldehydes, ketones, esters, acids), but the extent of the decrease depends on
this compound (Fig. 7.13). With an equal amount of the interfering compound.
however, the extent of the quenching effect is independent of the amount of the
sulphur compound 46 ,87. The decrease in response is attributed to the inactiva-
tion of the excited S2* species owing to their combination or to their collision
with organic compounds or the degradation products of the latter:
153

100

50

~--------1
~------2
o L-________ ~ ________ ~ ________ ~ ________ ~ _____

o 5 10 15 20
Organic substance (10-6mol/min)

Fig. 7.13. S2 emission quenching of benzo[b]thiophene. Interfering substances:


1 = ethanol; 2 = methanol; 3 = acetone; 4 = cyclohexane. ¢ = (response of FPD
with interfering substance present)/(response of FPD without interfering substan-
ce). (From ref. 7.)

k
--+ (7.13)

where X is the sulphur-free compound. The S2* concentration actually emitting


radiation is given by the equation

(7.14)

where X » S2 *, tis the mean 1ife of the excited S2 speci es and [S2 *] 0 is the
concentration of the excited S2 species in the absence of X7
The response decreases approximately exponentially with increasing concentra-
tion of sulphur-free compounds 7 ,18,85; this effect arises only starting from a
certain concentration of the compound in the carrier gas 18 ,25,46,86-88. The
critical mass rate is 5.10- 8 to 5·10-6 g/sec and depends on the type of detector
and the experimental conditions applied 7 ,25,40,44,46,86,87. The extent of the
quenching effect depends on the detector design; it is smaller with type A than
type B89. With type A the sample first diffuses from the oxygen-rich zone into
the reduction zone of the S2 emission through the part of the flame with the
highest temperature. The organic compounds are oxidized to carbon dioxide prior
to S2 formation; carbon dioxide has a low quenching effect 10 ,32,90. With type B,
the sample first diffuses through the region of lower temperature and lower
partial oxygen pressure, giving rise to products that cause considerable quench-
ing 89 • The quenching range depends on the 02:H2 ratio; the effect is less pro-
a

8
10.0 pi
7
16' 10-8 A
6

z
3 o
:cI-
i'i:z
2 1-0
..J_
>-:I:
:I: I-
I-IJJ
IJJ
1 ~

o
! I

o 2 4 o 2 4 6
MINUTES
Fig. 7.14. Chromatograms of samples containing 5 ppm each of five pesticides. (a) Single flame; (b) dual flame. Glass
column (2 m x 2 mm 1.0.) packed with 5% OV-IOI on Chrom W; temperature, 60 0 C. (Reprinted with permission from ref. 91.)
155

nounced at higher values of this rati0 89 • The detector temperature also influences
the extent of the quenching effect 86 • In the detector temperature range 100-190 0C,
the thiophene peak height varies within 5.7-10% of the initial peak height (with-
out hydrocarbons) during coelution with always the same amount of cyclohexane.
Under certain conditions the background due to stationary phase bleeding can also
induce quenching of the response 86 • Quenching occurs if the column temperature
is so high that the mass flow-rate of the stationary phase molecules through the
detector exceeds the threshold value causing the quenching effect.
Water does not cause response quenching during coelution with sulphur or phos-
phorus compounds 86 • In some instances, at lower concentrations of water in the
carrier gas, the response even increases irregularly by about 20%.
The quenching effect has also been observed with phosphorus, selenium and tin
compounds 40 ,44. The largest decrease was noted with sulphur compounds and the
smallest with selenium compounds 40 •
This phenomenon markedly affects the results of quantitative analysis if the
compounds to be detected elute immediately after a large zone of solvents or if
complex mixtures are analysed. In the former instance, the quenching effect is
caused by tailing of the solvent zone and in the latter it is due to incomplete
separation of the individual compounds on the chromatographic column and, con-
sequently, coelution of sulphur compounds and hydrocarbons. Fig. 7.14a shows this
phenomenon for the analysis of 10 ~l of a mixture of five thiophosphate pesti-
cides. Compared with the injection of 1 ~l of the same mixture, the first three
compounds show decreased responses (injection volume 10 times larger, recorder
range 100 times wider; the peaks should be of equal heights assuming the validity
of the quadratic dependence of the response on concentration).
Fig. 7.14b shows the chromatograms of the same mixture as in Fig. 7.14a, but
obtained with a dual-flame FPD. The decrease in response does not occur in this
instance. The response is not influenced with small concentrations of coeluting
substances.
The design of a dual-flame FPD is shown in Fig. 7.15. The compound to be de-
tected is first combusted in the lower flame. The decomposition products and the
uncombusted hydrogen of the lower flame proceed to the other burner, in which
the upper flame is generated with air from the second source. Hence the lower
flame serves to decompose the compounds leaving the chromatographic column into
relatively simple molecules. The radiation of the $2* and HPO* species is emitted
in the upper flame. The response of the dual-flame detector is greater than that
of the single-flame detector 36 ,91. With phosphorus compounds, the response of the
dual-flame detector is higher for phosphates and phosphites 36 (Table 7.2). The
conversion of various types of phosphorus compounds to PO requires different
amounts of energy. Phosphate, with a P=O bond energy of 140 kcal/mole, will be
156

FLAME 2 --r/7 WINDOW

FLAME TIP2 -~.,L,+--.f'..

FLAME 1 ---+'h4-~1-01

FLAME TIP1 -~~~H..I

Fig. 7.15. Schematic diagram of the dual-flame detector. (Reprinted with per-
mission from ref. 92.)

the most difficult to decompose. In the single-flame detector where PO formation


and detection as HPO species occur almost simultaneously, phosphate consequently
gives a low response. The response is greater if the compound is first dissociated
in one flame and then detected as HPO in another flame. Phosphine, with a P-C bond
energy of only 63 kcallmole, yields the same response in both detectors. The mini-
mum detectable mass rate is 5.10- 13 glsec of P and 5'10- 11 glsec of S. The fol-
lowing advantages of the dual-flame FPD have been reported 92 : (I) zero quenching
effect due to small concentrations; (2) reduced influence of the solvent; (3)
increased selectivity, PIC> 5'10 5 g Clg P, SIC from 10 3 to 106 g Clg S; (4) a
response independent of the structure of the compound; (5) a quadratic dependence
of the response on the concentration of the sulphur compounds; and (6) no quench-
ing of the flame. The small volume of this detector {O.17 mll and the short re-
sidence time of the effluent in the detector at the usual flow-rates (O.7 msec)
make it possible to use the detector with capillary cOlumns 93 •
157

7.11. FLAME STABILITY

The injection of large sample volumes extinguishes the detector flame. To


avoid this, a valve has to be placed between the detector and the chromatographic
column in order to displace the solvent from the effluent. If the column ef-
fluent is mixed with hydrogen prior to entering the flame, the flame is not
extinguished owing to changing large liquid samples, either when hydrogen mixed
with the carrier gas enters the burner (oxygen around the burner 13 ) or when both
gases are supplied around the burner (oxygen into the burner 27 ). No flame extinc-
tion has been noted with the dual-flame detector either92 • Another type of design
that does not suffer from flame extinction is an FPD having its decomposition
compartment separated from the emission compartment 47 ,88. The design of this de-
tector, which exhibits further advantageous features (see sections 7.3 and 7.4).
is shown in Fig. 7.16.

Fig. 7.16. FPD with separate combustion chamber. 1 = Combustion chamber; 2


oxidant and gas inlet orifices; 3 = emission chamber. (From ref. 47.)

7.12. OTHER IDENTIFICATION POSSIBILITIES

The fundamental identification information provided by the FPD is the selec-


tive response to sulphur and phosphorus compounds and/or to other heteroelements.
as described earlier in this chapter. However. two additional auxiliary techni-
ques allowing sulphur compounds to be differentiated from phosphorus compounds.
if necessary. can be inferred from the above analyses of the features of this
detector.
The response to sulphur compounds increases exponentially with increasing
amount of these compounds whereas the response to phosphorus compounds increases
158

linearly. By means of repetitive analyses of the same sample. but injecting dif-
ferent volumes, the two types of compounds can be distinguished by virtue of the
different increases in the peak heights. What is true for sulphur compounds is
also true for selenium compounds, whereas compounds of the other elements behave
similarly to phosphorus compounds.
The dependence of the FPD response on the detector temperature tends to de-
crease with sulphur compounds. whereas it tends to increase with phosphorus com-
pounds. Duplicate analyses. this time· at two different detector temperatures,
enables one' to distinguish the two types of compounds. because a smaller peak
height will be obtained for the sulphur compound at an elevated detector tem-
perature compared with a larger peak height for the phosphorus compound.

REFERENCES

1 G. Salet, Ann. PhYB •• 137 (1869) 171.


2 H. Draeger and B. Draeger. Ger. Pat •• No. 113918. 1962.
3 R.t1. Dagnall, K.C. Thompson and T.S. West, Analyst (London), 92 (1967) 506.
4 A. Syty and J.A. Dean, Appl. Opt., 7 (1968) 1331.
55.5. Brody and J.E. Chaney, J. Gas Chromatogr., 4 (1966) 42.
6 R.S. Juvet and R.P. Durbin, J. GaB Chromatogr., 1 (1963) 14.
7 T. Sugiyama, Y. Suzuki and T. Takeuchi. J. Chromatogr •• 80 (1973) 61.
8 M,C. Bowman and t1. Beroza, Anal:. Chem •• 40 (1968) 1448.
9 R. Pigliucci, W. Averill. J.E. Purcell and L.S. Ettre, Chromatographia. 8
(1975) 165.
10 W.E. Ruprecht and T.R. Phillips. Anal. Chim. Aata. 47 (1969) 439.
11 H.A. 11oye. Anal. Chem •• 41 (1969) 1717.
12 \~.L. Crider and R.~'i. Slater, Jr •• Anal. Chem •• 41 (l969) 531.
13 C.A. Burgett and L.E. Green. J. Chromatogr. Sci., 12 (1974) 356.
14 S.O, Farwell and R.A. Rasmussen. J. Chromatogr. Sci •• 14 (1976) 224.
15 S,O. Farwell, n.R. Gage and R.A. Kagel. J. Ch.romatogr. Sci., 19 (1981) 358.
16 T. Sugiyama. Y. Suzuki and T. Takeuchi. J. Chromatogr., 85 (1973) 45.
17 T. Sugiyama, Y. Suzuki and T. Takeuchi. J. Chromatogr •• 77 (1973) 309.
18 M. Maruyama and M. Kakemoto, J. Chromatogr. Sci •• 16 (1978) 1.
19 J.G, Eckhardt. M.B. Denton and J.L. 11oyers, J. Chromatogr. Sci •• 13 (l975)
133.
20 R. Greenhalgh and M.A. Wilson, J. Chromatogr •• 128 (1976) 157.
21 C.H. Burnett, D.F. Adams and S.D. Farwell, J. Chromatogr. sci •• 16 (1978)
68.
22 A.D. 11izany, J. Chromatogr. Sci., 8 (1970) 151.
23 T.J. Cardwell and P.J. t1arriott. J. Chromatog'J'. Sai •• 20 (1982) 83.
24 R.E. Pecsar and C.H. Hartmann, J. Ch'J'omatogr. Sci •• 11 (1973) 492.
25 M. Dressler, Chem. Liety. 78 (1984) 645.
26 S. Kapila and C.R. Vogt, J. Chrornatog'J'. Sci •• 17 (1979) 327.
27 A. R. L. r1oss. Scan~ 4 (1974) 5.
28 li.P. Cochrane and R. Greenhalgh. Ch'J'omatographia. 9 (1976) 255.
29 C.H. Hartmann, Anal:. Chem., 43 {1971} 113A.
30 H.W. Grice. M.L. Yates and D.J. David, J. Chromatogr. Sci., 8 (1970) 90.
31 J.F. licGaughey and S.K. Gangwai. AnaZ. Chern •• 52 (1980) 2079.
32 B.J. Ehrlich. R.C. Hall. R.J. Anderson and H.G. Cox. J. Chromatogr. Sai ••
19 (l981) 245.
33 D.G, Greer and T.J. Bydalek. Environ. Sai. Teahnoi •• 7 (1973) 153.
34 T. Sugiyama. Y. Suzuki and T. Takeuchi. J. Ch~omatogr. Soi •• 11 (1973) 639.
159

35 S. Sass and G.A. Parker. J. Chromatogr •• 189 (1980) 331.


36 C.R. Vogt and S. Kapila. J. Chromatogr. Sci., 17 (1979) 546.
37 C.H. Burnett, D.F. Adams and s.n. Farwell, J. Chromatogr. Sai, 15 (1977)
230.
38 R.K. Stevens, J.D. Mulik, A.E. O'Keefle and K.J. Krost, Anal. Chem., 43
(1971) 827.
39 M. Dressler, J. Chromatogr., 262 (1983) 77.
40 C.G. Flinn and \~.A. Aue, J. Chromatogr •• 153 (1978) 49.
41 W.A. Aue and C.G. Flinn, J. Chromatogr •• 158 (1978) 161.
42 W.A. Aue and C.R. Hastings. J. Chromatogr •• 87 (1973) 232.
43 J. Sev~'k and N.P. Thao, Chromatographia. 8 (1975) 559.
44 I~.A. Aue and C.G. Flinn. J. Chromatogr •• 142 (1977) 145.
45 C.R. Hastings, D.R. Younker and W.A. Aue, Environ. Health, 8 (1974) 265.
46 L. Blomberg, J. Chromatogr., 125 (1976) 389.
47 V.A. Joonson and E.P. Loog, J. Chromatogr., 120 (1976) 285.
48 R.M. Dagnall, K. C. Thompson and T.5. I'lest, Analyst (London), 93 (1968) 518.
49 C.G. Flinn and W.A. Aue, J. Chromatogr. Soi., 18 (1980) 136.
50 W.A. Aue and C.G. Flinn, Anal. Chem., 52 (1980) 1537.
51 R.S. Braman and M.A. Tompkins, Anal. Chem •• 51 (1979) 12.
52 S. Kapila and C.R. Vogt, J. Chromatogr. Soi., 18 (1980) 144.
53 C.G. Flinn and I~.A. Aue, J. Chromatogr •• 186 (1979) 299.
54 R.J. Maguire and H. Huneault, J. Chromatogr., 209 (1981) 458.
55 R.J. Maguire and R.J. Tkacz, J. Chromatogr •• 268 (1983) 99.
56 C.V. Overfield and J.D. Winefordner, J. Chromatogr. Sci., 8 (1970) 233.
57 B. Gutsche and R. Hermann. Fpesenius Z. Anal. Chem., 245 (1969) 274.
58 B. Gutsche and R. Hermann. Fresenius Z. Anal. Chern., 249 (1970) 168.
59 B. Gutsche and R. Hermann, FreseniuB Z. AnaZ. Chern., 253 (1971) 257.
60 M.C. Bowman, M. Beroza and G. Nickless, J. Chromatogr. Sci., 9 (1971) 44.
61 R.F. 110seman and W.A. Aue, J. Chromatogr., 63 (1971) 229.
62 B. Versino and H. Vissers, Chromatographia, 8 (1975) 5.
63 B. Vers;no and G. Rossi, Chromatographia, 4 (1971) 331.
64 W.A. Aue. C.W. Gehrke, R.C. Tindle, C.D. Ruyle, D.L. Stalling and S.R.
Koirtyohann, Characteristios of the Alkali Flame Detector, 5th National
Meeting for Applied Speotroscopy, Chioago, IL, June 1966.
65 A.V. Nowak and H.V. Malmstadt, Anal. Chern •• 40 (1968) 1108.
66 M.C. Bowman, ri. Beroza and K.R. Hill, J. Chmmatogr. Soi., 9 (1971) 162.
67 W.A. Aue and R.F. Moseman, J. Chromatogr., 61 (1971) 35.
68 A. Karmen. Anal. Chem •• 36 (1964) 1416.
69 M. Dressler and J. Jan&k, Colleot. Czeoh. Chern. Commun., 33 (1968) 3970 ..
70 M. Dressler, AZkali Flame Ionizaaon Detector, Thesis, Institute of Analytical
Chemistry, Brno, 1969.
71 M. Dressler and J. Janak, J. Chmmatogr., 44 (1969) 40.
72 K.O. Gerhardt and W.A. Aue, J. Chromatogr., 52 (1970) 47.
73 J.L. Monkman and L. Dubois, in H.J. Noebels, R.F. Wall and N. Brenner
(Editors) Gas Chromatography, Academic Press, London, 1961, p. 333.
74 tt C. Bowman and M. Beroza, J. ChT'omatogr. sci., 7 (1969) 484.
75 R. Gutsche and R. Hermann, Fresenius Z. Anal. Chern •• 259 (1972) 126.
76 E.J. Sowinski and I.H. Suffet, J. Chromatogr. Soi., 9 (1971) 632.
77 R. Ross and T. Shafik, J. Chromatogr. Sci., 11 (1973) 46.
78 W.A. Aue and H.H. Hill, Jr., AnaZ. Chern., 45 (1973) 729.
79 F.M. Zado and R.S. Juvet. Jr .• Anal. Chern •• 38 (1966) 569,
80 R.S. Juvet, Jr. and R.P. Durbin, Anal. Chern., 38 (1966) 565.
81 H.H. Hill, J~ and W.A. Aue, J. Chromatogr •• 74 (1972) 311.
82 J. Sevctk, Chromatographia,'4 (1971) 195.
83 W.A. Aue and H.H. Hill, Jr., J. Chromatogr., 70 (1972) 158.
84 J.M. Zehner and R.A. S1monanaitis, J. Chromatogr. Sd., 14 (1976) 348.
85 S.G. Perry and F.W.G. Carter, in R. Stock (Editor), 8th International Gas
Chromatography Symposium, Dublin, 1970, Prooeedings, Institute of Petroleum,
London, 1971, p. 381.
~o

86 M. Dressler. J. Chromatogr •• 270 (1983) 145.


87 D.A. Ferguson and L.A. Luke, Chromatographia. 12 (1979) 197.
88 S. Hasinski. J. Chromatogr •• 119 (1976) 207.
89 S.A. Fredriksson and A. Cedergren. Anal. Chern •• 53 (1981) 614.
90 S.A. Fredriksson and A. Cedergren. Anal. Chim. Acta. 100 (1978) 429.
91 P.L. Patterson. AnaL. Chern •• 50 (1978) 345.
92 P.L. Patterson. R.L. Howe and A. Abu-Shumays. Anal. Chern., 50 (1978) 339.
93 F.J. Yang and S.P. Cram, J. High Reaolut. Chromatogr. Chromatogr. Commun.,
2 (1979) 487.
161

Chapter 8

CHEMILUMINESCENCE DETECTORS

CONTENTS

8.1. Introduction • . . . . • • . • . 161


8.2. Detector for N-nitroso compounds 161
8.2. 1. Response . . . . • • • . 163
8.2.2. Selectivity of response . . . 167
8.3. Detector for nitroaromatic compounds . • . 169
8.4. Detector for nitrogen-containing compounds . . ....
8.5. Ozone chemiluminescence detector for compounds not containing
170
ni trogen . . . . . . . • . . . • • • . • . . · . 174
8.SA.Redox chemiluminescence detector . • • • . . 174/312
8.6. Chemiluminescence detector with sodium metal 174
8.7. Fluorine-induced detector • . 177
References . . . • . . . • • . . . . . . . • . 179

8.1. INTRODUCTION

Chemiluminescence detectors (CLDs) are based on emission spectroscopy. This


chapter deals with detectors utilizing (1) the reaction of ozone with nitrogen
oxide, the reaction of sodium metal vapour with nitrous oxide and halogenated
compounds, and the reaction of fluorine with sulphur compounds; and (2) the
emission of radiation due to the transition of excited species formed by these
reactions to their ground states.

8.2. DETECTOR FOR N-NITROSO Cor~POUNDS

The selective CLD designed to detect N-nitroso compounds utilizes the tech-
nique developed for the analysis of nitrogen oxides 1 ,2 based on the reaction 3
of nitrogen oxide with ozone combined with the preceding pyrolysis of nitroso
compounds.
Fig. 8.1 shows a schematic diagram of a CLD. The effluent from the chromato-
graphic column enters the pyrolyser where the selective catalytic decomposition
of N-nitroso compounds takes place, giving rise to a nitrosyl radical and an
organic radical. The N-NO bond is the weakest in these compounds 4:

(8.1)
162

3
6

8
9
Fig. B.1. Schematic diagram of a CLD. 1 = Sample injection point; 2 = chromato-
graph; 3 = pyrolysis chamber; 4 = cold trap; 5 = filter; 6 = vacuum; 7 = photo-
multiplier; B = electronics; 9 = recorder. (Reproduced with permission from
ref. 5.)

where R1 and R2 are organic radicals. The pyrolyser effluent expands into the
evacuated reaction chamber in which the nitrogen oxide (the nitrosyl radical)
reacts with ozone, giving excited nitrogen dioxide:

(B.2)

The excited nitrogen dioxide rapidly decays back to its ground state, emitting
light in the near-infrared region of the spectrum:

k
N0 2* -L NO 2 + h\! (B.3)

The emitted radiation is detected by a photomultiplier through a red optical


fi lter.
The decomposition of substances with a catalyst at lower temperatures (about
300 0 C) is more advantageous than pyrolysis at elevated temperatures, because the
decomposition of the N-NO bond is more selective, even though the detector
response to nitrosamines is diminished at lower temperatures 6 • The response
level depends on the temperature of the pyrolysis chamber also in decompositions
without a catalyst (Fig. B.2). The maximum response was observed 7 at temperatures
in the range 300-400 0 C. A mixture of W0 3 and W20 05B was found to be the most
163

100

80

60

• 1
40 Cl 2
o 3
.4
20 '" 5
A 6

0
200 300 400 500 600'C

Fig. B.2. The N-nitroso compound response as a function of pyrolytic chamber


temperature. 1 = N-Nitrosodibutylamine; 2,= N-nitrosodiethylamine; 3 = N-nitroso-
dimethyl amine; 4 = N-nitrosodipropylamine; 5 = N-nitrosomorpholine; 6 = N-nitro-
sopyrrolidine. R = Response. (Reprinted with permission from ref. 7.)

suitable catalyst 4• Gough B recommended W0 3 adsorbed on the walls of a porous


ceramic tube as a catalyst, because this arrangement results in an increase in
the catalyst lifetime.
The thermal energy analYSer9 ,10 (TEA), manufactured by Thermo Electron Corp.,
U.S.A., connected to a gas chromatograph, is a commercial detector. However, it
is much more expensive than conventional detectors. A luminescence detector
laboratory-made from available parts was described by Gough et al. B•

B.2.1. Response

The fundamental reaction of the nitrosyl radical in the detector is reaction


B.2. However, nitrogen oxide also reacts with ozone according t0 4

(B.4)

The NO fraction converted to the excited state, p*, is given by .


164

1
(8.5)

and depends on the temperature (about 7% at 20 oC, about 13% at 100 0 C). The ex-
cited nitrogen dioxide can lose its energy not only by reaction 8.3, but also
due to collisions with other molecules in the system (M):

(8.6)

N0 2" + t·1 + M ...... N0 2 + M + M (8.7)

The contribution of reaction 8.7 to the deactivation of N0 2" is negligible


under vacuum. Then, the NO fraction decaying back to the ground state with the
emission of radiation, F, is given by

F(NO) = F" • • [M] • [t1] (8.8)

x
-4
10

0.51.0 5 10 50 100 500 1000


mmHg

Fig. 8.3. Effect of pressure on the fraction X of nitrosyl radicals emitting


light. (From ref. 4.)
165

The change of F with pressure (at T = 20 0 C) is shown in Fig. 8.3. At a pres-


sure of 2 mmHg (the range within which the detector is operated). the NO frac-
tion that contributes to light emission is 3.10- 4. The number of efficient
light quanta is also reduced by approximately an order of magnitude owing to
the detector geometry and photomultiplier characteristics 4•

R
DEN DMN

DPN

- min

Fig. 8.4. Chromatogram of a 100-~1 solution containing 5 ng/ml of N-nitroso


compounds. SARCOSN = N-nitrososarcosinate; PYRN = N-nitrosopyrrolidine;
NIP = N-nitrosopiperidine; DBN = N-nitrosodibutylamine; DPN = N-nitrosodi-
propyl amine; DEN = N-nitrosodiethylamine; DMN = N-nitrosodimethylamine.
R = response. Column. 6.5 m x 2 mm 1.0 •• gacked with 15% free fatty acid phase
on Chromosorb WAI~ DMCS; temperature, 185 C. (From ref. 12.)

The detection limit for nitrosodimethylamine is about 5'10- 11 g 6,8,11


(Fig. 8.4). The response is linear over a concentration range of five 12 to
six 5 orders of magnitude.
The detector response expressed per nitrosyl group in the molecule of the
compound is not constant 13 , apparently depending slightly on the structure of
the compound (see Table 8.1).
......
0'1
TABLE 8.1 0'1

TEA RESPONSE FACTORS FOR DIFFERENT N-NITROSO COMPOUNDS

Reprinted with permission from ref. 13.

N-Nitroso compound Mol. wt. Concentration Measured Response Relative


(llg/ml) response per response
(i ntegrated nitrosyl (nitrosyl
units) group mole basis)

N-Nitrosodimethylamine 74 0.964 235 18.1 1.00


N-Nitrosodi ethyl amine 102 1.07 204 19.4 1.07
N-Nitrosodipropylamine 130 0.84 128 19.8 1.09
N-Ni trosodi phenyl amine 198 1.86 189 20.1 1.11
N-Nitroso-N-ethylaniline 150 1.17 145 18.6 1.03
9-Nitrosocarbazole 196 1.99 167 17.6 1.03
N-Nitroso-N-methyl urethane 132 0.52 74 18.8 1.04
N-Ni troso-N-phenyl benzyl amine 212 2.10 166 16.7 0.92
Ethyl N-nitrososarcosinate 146 2.15 258 17.5 0.97
N-Methyl-N-nitroso-N-nitroguanidine 147 4.71 590 18.4 1.02
N-Nitrosopiperidine 114 0.93 172 21.2 1.17
Dinitrosopiperazine 144 1.99 434 15.7 0.87
167

8.2.2. Selectivity of response


The detection selectivity for N-nitroso compounds with a CLD is based on the
selective decomposition of the compound at relatively low temperatures and on
the generation of excited nitrogen dioxide molecules by the reaction of nitrogen
oxide with ozone. Some other substance, e.g., carbon monoxide and ethylene,
also react with ozone to produce luminescence, but the wavelength of this radia-
tion is in the blue and visible light regions l2 • A filter that does not transmit
radiation up to 0.6 ~m eliminates these inferferences.
In order to provide high detection selectivity, the part of the detector
situated between the pyrolyser and the reaction chamber is maintained at tem-
peratures below _150 0 C13 ,14. The NO vapour pressure exceeds 1 atm at this tem-
perature, whereas the vapour pressure of most organic compounds is much lower
than 1 atm. Therefore, compounds that could react with ozone giving luminescence
in the IR region are collected just ahead of the reaction chamber 13 . In addi-
tion, a short column packed with the porous polymer Tenax is sometimes placed
ahead of the reaction chamber6 ,7.
Under certain conditions, the nitro compounds can also yield a nitrosyl
radica1 4 :

(8.9)

-N0 2 + C ~ CO + NO (8.10)

At equilibrium under ideal conditions, the conversion range according to reac-


tion 8.9 varies from 10% (300 0 C) to 94% (700 0 C). In practice, however, the
reaction kinetics are so slow that no conversion can be observed during detec-
tion even at 400 0 C4. When using metallic catalysts (gold, molybdenum), the re-
duction of nitrogen dioxide to nitrogen oxide (reaction 8.10) proceeds rapidly
at 200 0 C. For this reason, active forms of carbon and possible contact with
metallic catalysts should be eliminated.
When studying how the temperature of the pyrolysis· chamber affects the res-
ponse of compounds other than N-nitroso compounds (Fig. 8.5), it has been
found 7 that the detector also responds to nitrohexane, nitrotoluene, hexyl
nitrite and dipropylnitramine. The course of this dependence differs markedly
for the individual compounds. Hexyl nitrite pyrolyses to give nitrogen oxide
at relatively low temperatures of 200-300 0 C; the response decreases at temper-
atures above 300 0 C. Nitrotoluene and nitrohexane respond only at higher tem-
peratures. The molar response of dipropylnitramine at 400 0 C is very high,
amounting to 50-96%, according to the experimental conditions, of the molar
response to N-nitrosodialkylamine 7 ,15. Table 8.2 lists the relative molar
168

100
R
a/a
80

60

40

20

Fig. 8.5. Detector response as a function of pyrolytic chamber temperature.


1 = Nitrohexane; 2 = nitrotoluene; 3 = hexyl nitrite; 4 = dipropylnitramine.
R = response. (Reprinted with permission from ref. 7.)

TABLE 8.2

tl0LAR RESPONSES OF NITRAtUNES RELATIVE TO THE CORRESPONDING NITRASAMINES

From ref. 16.

Nitrami ne Molar response relative to nitrosamine

N-Nitrodimethylamine 0.87
N-Nitrodiethylamine 0.82
N-Nitrodipropylamine 0.78
N-Nitromethylpentylamine 0.81
N-Nitrodibutylamine 0.75
N-Nitropiperidine 0.80
N-Nitropyrrolidine 0.73

responses of other nitramines with regard to the corresponding nitrosamines.


These values approach 8~% for a pyrolyser temperature of 500 0 C. For this reason,
misleading interpretations could be made when analysing a mixture of nitro and
nitroso compounds 16 (see also section 8.3). N-Dimethylbenzylamine 17 , several
C-nitroso compounds 18 and nitrites do also respond. However, nitrites decompose
in the hot zone of the injection chamber of the chromatograph (200 0 C) giving
nitrogen oxide, which elutes in the dead vOlume 12 •
The initial statement that the CLD responds only to N-nitroso compounds, so
that no preliminary purification of the samples (extracts) is necessary14,
should be treated with caution.
169

N-Nitrosodimethylamine (NOMA) dissolved in a mixture of hexane and ethanol


gives a higher response than its normal response. This anomalous response
increases with decreasing concentration of ethanol in the mixture; it is ap-
proximately an order of magnitude higher at a 10% ethanol concentration. The
alcohol or the contaminants contained therein act as catalysts of the inter-
action of NOMA (or of its pyrolysis products) with hexane or its contaminants,
giving a highly volatile product that either responds itself or improves the
efficiency of the reaction of nitrogen oxide with ozone 19 •
Amines have been found to reduce the detector response to N-nitroso com-
pounds 20 • The nitrogen oxide formed by cleavage of the N-NO bond can react with
amine molecules at the chamber outlet, giving rise to another nitrosamine which
therefore cannot react with ozone. The initial response to the N-nitroso com-
pound is reduced or even zero, depending on the degree of recombination. How-
ever, Parees and Prescott6 found that the response for N-nitrosodimethylamine
in the presence of excess of diethylamine (in tailing the amine peak) equalled
the response of this compound in the presence of excess of dichloromethane. The
same is also true for N-nitrosomorpholine in morpholine.

8.3. DETECTOR FOR NITROAROMATIC COMPOUNDS

The response of the TEA detector to nitroaromatic compounds increases rapidly


with increasing temperature of the pyrolyser5 (see Fig. 8.6). The fact that
nitrogen oxide is also generated by the pyrolysis of nitro compounds (reaction
8.9) was employed by Lafleur and Mills 5 to determine nitroaromatic compounds
with the aid of a CLD. The pyrolysis temperature was 800-900 0 C (the response de-
creased drastically by several orders of magnitude at lower temperatures), the
other experimental conditions being the same as for the nitroso compound de-
tector. The detection limit of this detector is 0.6 ng of nitroaromatics and
the dynamic linear range covers four orders of magnitude. When using a capillary
column inserted deeply into the ceramic pyrolysis tube, the detection limit was
found to be tens of picograms 21 or even picograms 22 • This arrangement of the
column is of great importance, because it prevents losses of polar substances
due to adsorption. If the column is inserted only to the transfer line or into
the relatively cold zone of the pyrolysis tube, the detection limit is only
nanograms of nitroaromatics 21 • The Thermo Electron TEA Model 543 analyser is
designed for both nitroso and nitro group detection 23 •
170

log R
~ ___ ~1

·C

Fig. 8.6. Response of nitroaromatic compounds as a function of pyrolyser tem-


perature. 1 = 2,4,6-Trinitrotoluene; 2 = 2,3-dinitrotoluene; 3 = 2,6-dinitro-
toluene; 4 = 3-nitrotoluene; 5 = 2-nitrotoluene.
(Reprinted with permission from ref. 5.)

8.4. DETECTOR FOR NITROGEN-CONTAINING COMPOUNDS

The chemiluminescence analyser for nitrogen-containing compounds is based


on the catalytic decomposition of these compounds in the presence of oxygen at
temperatures of 900-IOOO oC and on the subsequent ozone oxidation of the nitrogen
oxide formed 24 ,25, as with nitroso compounds (see eqns. 8.2 and 8.3):

(8.11)

The detector consists of a pyrolyser with an oxygen supply and an NO ana-


lyser 26 - 29 ; the analyser manufactured by Antek 28 and/or the Thermo Electron
TEA Model 610 29 are commercial detectors of this type. The only difference
from the nitroso detector is in the oxidative decomposition of the substance.
The detection limit is in the nanogram region. Also with this detector, the
response level depends on the pyrolyser temperature, which increases to 900 oC.
The detection sensitivity is given by the efficiency of the conversion of
nitrogen-containing compounds to nitrogen oxide. The molar responses of am-
monia, acetonitrile and nitromethane are about 1.5 times greater than those of
amines 26 ,27,30. Ethanol, benzene and acetone in I-~l amounts yield signals
equivalent to several nanograms of trimethylamine 26 ,27, and this also enables
one to determine nitrogen-containing compounds eluting in the peaks of these
solvents (Fig. 8.7).
171

r(c)--:
I I
I I
I I
I
I
I
I
TMA
DMA

I
I
I

i-PA !,
I
,,
,,
,,
,
\ i-SA

24 28 min

Fig. 8.7. Chromatogram of mixture of amines. TMA = trimethylamine; DMA = di-


methylamine; i-PA = isopropylamine; DEA = diethylamine; i-SA = isobutyl amine;
BA = butylamine. (a) Ethanolic solution of the sample; (b) ethanol only;
(c) ethanol only, flame-ionization detection. Glass column (2 m x 3 mm I.D.)
packed with 5% squalane plus 2% potassium hydroxide on Chromosorb 104. Tem-
perature, 130 oC. (From ref. 27.)

A modified 31 commercial TEA can also be used as a gas chromatographic de-


tector (/1ode 1 610). The effl uent from the col umn is oxi di zed at 690 0 C by a
metal oxide to give the nitrosyl radical. The metal oxide is continuously re-
generated by the oxygen passing through the pyrolyser. The linearity of response
covers four orders of magnitude and the detection limit is about 3'10- 13 mole
i.e., 2'10- 11 g, of pyridine. Hence the detection sensitivity is commensurate
with that of the original CLD for .N-nitroso compounds (the TEA). The molar
responses of individual compounds are listed in Table 8.3. The response de-
pends on the number of nitrogen atoms in the molecule of the compound; the
response to N-nitrosamines is twice that of pyridine.

(8.12)

Neither carbon dioxide nor water interfere; 1 ~l of an organic solvent gives


only a small negative response caused by the passage of a large volume of
organic vapours through the detector (Fig. 8.8).
172

TABLE 8.3

r~OLAR RESPONSE OF THE CHEMI LUMI NESCENCE DETECTOR

From ref. 31.

Compound Relative molar response


N mode Nitro mode

pyri di ne 1.00 None


Ammonia 1.04 None
Trimethylamine 1.03 None
Triethylamine 1.00 None
Diethylamine 1.04 None
Morpholine 0.97 None
3-Methylpiperidine 1.05 None
n-Propylamine 0.95 None
Isopropylamine 0.92 None
sec.-Butylamine 1.05 None
Anil i ne 1.00 None
Acetonitrile 1.00 None
N-Nitrosodimethylamine 2.01 1.03
N-Nitrosodiethylamine 2.08 1.00
N-Nitrosopropyl ami ne 1.96 0.98

(A) (8 ) (C) (0)

cD 100 100 100 100


~

Z
l-
I-
«

k
UJ
VI
z
a

r-V
a.
VI
UJ hr-- ~
.:.e
a:

0 2 4 0 2 4 0 2 4 0 2 4
TIME (MINUTES)

Fig. 8.8. Chromatograms of 1 ~l of orgdnic solvents. TEA, N mode. (A) Acetone;


(B) toluene; (C) benzene; (D) methanol. Glass column (l.7 m x 2 mm 1.0.) packed
with 20% potassium hydroxide on Chromosorb W. Temperature, programmed from
90 nC (2 min) to 140 0 C at 12 oC/min. (From ref. 31.)
173

DETECTOR 690'C

)I------<GC

N MODE

DETECTOR 600'C

)i----<GC

NITRO MODE

Fig. B.9. Schematic diagram of the GC-TEA interface for the N mode and the
nitro mode (From ref. 31.)

100 2 NITRO

R
0/0

50

o 2 4 6 o 2 4 6
min

Fig. B.10. Comparison of the N mode and nitro mode chromatograms. 1 = Pyridine;
2 = nitrosodimethylamine; 3 = N-nitrosodiethylamine; 4 = N-nitrosodi-n-propyl-
amine. Glass column (1.6 m x 2 mm 1.0.) packed with 20% Carbowax 20M. Temperature,
program~~d from 160 0 C (1 min) to 190 0C at 100C/min. (From ref. 31.)
174

The positioning of a four-way valve in the oxygen flow upstream of the


pyrolyser enables the detector to be operated in the N mode or in the nitro
mode 31 (Fig. 8.9). In the N mode, oxygen is supplied to the pyrolyser and the
detector responds to all nitrogen-containing compounds. No oxygen is supplied
to the detector in the nitro mode and, at a pyrolyser temperature of 600 0 C,
the detector responds only to the compounds that produce an NO radical, such
as N-nitroso compounds, and nitroaromatic compounds (see also Table 8.3). The
change from the N mode to the nitro mode is accomplished by switching over the
valve. An example of an analysis performed with both modes is shown in Fig. 8.10.

8.5. OZONE CHEMILUMINESCENCE DETECTOR FOR COMPOUNDS NOT CONTAINING NITROGEN

The reaction of a solute with ozone, which gives rise to chemiluminescence,


is also used for the selective detection of hydrocarbons and some sulphur com-
pounds 32 ,33:
solute + 03 ~ A* + further products (8.13)

A* ~ A + hv (8.14)

In this instance, the reaction occurs at normal pressure. The response selec-
tivity is provided by variations in the reactivity of different classes of
compounds towards ozone and can be varied by changing the detector temperature
(Table 8.4). The response increases with detector temperature up to about 300 0 C,
after which it rapidly decreases owing to ozone decomposition. The response
selectivity is the greatest at temperatures up to 150 0 C. The minimum detectable
mass rate also depends on the detector temperature and is fairly high (the noise
also increases with temperature); of the compounds listed in Table 8.4, it is
the lowest for propadiene at 250 0 C, being 8.2'10- 12 mole/sec 32 .

8.5A. REDOX CHEMILUMINESCENCE DETECTOR (see p.312)

8.6. CHEMILUMINESCENCE DETECTOR WITH SODIUM METAL

Reactions producing chemiluminescence proceed between sodium metal vapour and


nitrous oXide 34 :

(8.15)

(8.16)

(8.17)

Na* ~ Na + hv (8.18)
175

TABLE 8.4

DETECTOR SENSITIVITY FOR DIFFERENT TYPES OF HYDROCARBONS

Calculated as peak area/mole (counts/mole). (From ref. 32. )

Temperature Propane Benzene Ethylene Acetylene Propadiene


(0C)

50 - 9 2.9.10 10 4.3.10 10 1.3.1011


100 - 6 1.0.10 1.4·lOH 1.7·lOH 6.9·lOB
150 2.2.10 9 10
1.2.1010 2.4.10 5.7.10 11 1.2.10
175 3.8.10 10 4.1.10 11
3.4.10 11 9.0'1012 12
1.6.10 12
200 3.8.10 1.2.10 11 6.2.10 12 1. 8.10 12 2.2.10
225 3.0.10 11 3.3.10 11 1.3.10 2.9.10 12 3.4.10 12
250 1.2·10 12 7.4·10 11 2.3·10 12 3.9·10 5.0,10 12

TABLE 8.5

RELATIVE RESPONSES

Reprinted from ref. 35 with permission.

Compound Concentration Response rati 0


(ppm) against N20 of the
same concentration

NO 50 1:400
N02 85 1:200
CO 50 1:200
CO 2 1000 0
0 50 1:200
Sb2 1000 0
H2S 1000 0

Araki et al. 35 used these reactions for the gas chromatographic detection of
nitrous oxide. The sodium metal vapour generated in the vaporization cell of the
detector at 310 0 C is carried into the reaction cell by nitrogen. A pressure of
several Torr is maintained in the detector. Chemiluminescence appears at the tip
of the nozzle through which the effluent from the chromatographic column is sup-
plied to the detector. The minimum detectability is 1.9.10- 12 mole/sec and the
response selectivity relative to other gases covers about two orders of magnitude
(Table 8.5). The response is a function of the vaporization cell temperature
(maximum at 320 oC). reaction cell temperature (maximum at 260 oC). and reaction
176

cell pressure (maximum 4-8 Torr), and is also affected by the flow-rate of the
sodium metal vapour carrier gas and the flow-rate of the carrier gas from the
chromatographic column.
Sodium vapour also reacts with volatile aliphatic hydrocarbons containing
more than two halogen atoms in the molecule 36 ,37.

RX 2 + Na ~ ·RX + NaX (8.19)

·RX + Na ~ (R) + NaX* (8.20)

NaX * + Na ~ NaX + Na * (8.21)

Chemiluminescence then again arises from the transition of Na* to the ground
state (eqn. 8.18).
A simplified and improved metallic version 38 of the above detector is shown
in Fig. 8.11. Argon was used as the carrier gas. The sensitivity of this type
is better than that of the original Pyrex model. The chloroethane molar response
(Table 8.6) varies with the position of the chlorine atoms in the molecule,
which can be explained by the following reaction mechanism 37 :

Pump

5 cm

Union Tee
( 112 in.'
~ Stainless - steel tube

Effluent
t from GC
\
~
Quartz window
t
Reducer
(1116 in.)

Fig. 8.11. Stainless-steel CLD with sodium metal. (From ref. 38.)
177
TABLE 8.6
DETECTION LIMITS AND RELATIVE MOLAR RESPONSES (RMR) OF CLD WITH Na
From ref. 38

Polyhalogenated Detection limit (ng) RMR**


hydrocarbon (SIN = 3)*

CH~C12 0.3 1
CH 13 0.2 3
CC14 0.003 230
CF3C1 10 0.04
CF~C12 0.5 1
CF 13 0.1 5
CHBr3 0.2 4
CHBr2Cl 0.3 3
C1CH2CH~Cl 0.0009 430
C1CH~CH 12 0.0009 540
C12C CHC12 0.03 25
C1CH2CC13 0.02 26
CH3CHC12 0.9 0.4
CH3CC1~ 0.8 0.7
C1CF2C Cl~ 0.2 4
eis-C1CH= HCl 0.2 2
tl'ans-C1CH=CHCl 0.2 2
C1CH=CC1 2 0.002 230
C12C=CC12 0.001 590

*S/N = signal-to-noise ratio.


**Normalized with respect to the value for CH 2C1 2•

'CH 2-CH 2Cl + Na + NaCl + CH 2=CH 2 (8.22)

CH 3-CHCl + Na + NaCl + CH 2=CH 2 (8.23)

The (R) in reaction 8.20 is the unsaturated molecule formed by the closure of a
double bond (reaction 8.22) or by the migration of a hydrogen atom (reaction
8.23). Reaction 8.23 has a much smaller light yield (e.g., for CH 3CHC1 2, CH 3CC1 3)
than reaction 8.22 (e.g., for C1CH 2CH 2Cl, C1CH 2CHC1 2 ). The other chloroethanes
investigated show medium sensitivity, with both reactions being responsible for
the chemiluminescence reaction. Fluorine atoms are inactive in the chemilumi-
nescence reaction. The linearity of response is 10 4-10 6•
The atomic sodium vapour reacts to produce chemiluminescence also with mono-
halogenated compounds and with several types of nitrogen and oxygen-containing
compounds. The light yields are low, however (Table 8.7).

8.7. FLUORINE-INDUCED DETECTOR

This detector monitors the chemiluminescence resulting from the reaction of


the gas chromatographic effluent with molecular fluorine at reduced pressure 39 •
178

TABLE 8.7

RELATIVE MOLAR RESPONSES OF OTHER COMPOUNDS RELATIVE TO DICHLOROMETHANE (= 1)

From ref. 38.

Compound RMR Compound RMR

n-C 3H I 0.02 CH~COCH3 0


6 0.009 C ~CHO 0
~R~~~C~CH3 0.007 t
( ~ B)2 NH 0
)t
( CH 30
n-Ct7 r
0.007
0.007
C 5 C2 H5
t
( RAe)2 S
0
0
CH~ N 0.001 C6 5 1 0
n- 3H~Cl 0.0009 C6 H6 0
C2 H50 0.0002

TABLE 8.8

DETECTION LIMITS AND RELATIVE RESPONSE FACTORS FOR VARIOUS SULPHUR-CONTAINING


COMPOUNDS

Reprinted with permission from ref. 45.

Compound Detection Relative


limit (pg) response factor

Allyl sulphide 24 8.3


Ethanethi 01 24 4.6
Ethyl sulphide 46 6.0
1-Butanethi 01 35 4.6
n-Butyl disulphide 73 4.3
1-Hexanethiol 84 2.5
n-Butyl sulphide 180 1.4
Isopentyl disulphide 200 1.8
1-0ctanethiol 257 1.0

The emission spectrum of sulphur compounds has been identified as vibrational


overtone bands of the ground electronic state of HF 40 • Whereas many organic
compounds react with fluorine to give excited-state HF products 39 ,41-44, only
sulphur compounds were found to react to yield HF in vibrational levels as
high as ~ = 5 and 6.
Emission is detected with a red-sensitive photomultiplier tube, and an optical
filter passing wavelengths between 660 and 740 nm is used. The detector is op-
179

TABLE 8.9

SELECTIVITY RATIOS OF n-BUTYL DISULPHIDE OVER VARIOUS ORGANIC COMPOUNDS

Reprinted with permission from ref. 45.

Compound Selectivity Compound Selectivity


ratio ratio

Anil i ne Methylene chloride


Carbon disulphide I-Dctanol
tpans-Cinnamaldehyde Propanol
Diethylamine Propionaldehyde
Hexane Tetrahydrofuran
I-Hexene Toluene
l-Iodohexane m-Xylene

erated at pressures below 2 Torr and the reagent gas is a mixture of 5% fluorine
in helium. If the flow-rate of the carrier gas (helium) is kept constant at
30 ml/min, the detector response is the highest for a reagent gas flow-rate of
15 ml/min 45 •
Detection limits and relative responses are listed in Table 8.8 and selec-
tinity ratios in Table 8.9. The fluorine-induced chemiluminescence detector does
not respond to sulphur gases such as S02' COS, H2S, CS 2. The linearity of
response for sulphur compounds covers three orders of magnitude 45 •
A iodine-selective detector based on fluorine-induced chemiluminescence and
with a detection limit of 1 ~g has also been described 46 • In this instance,
chemiluminescence is produced by reaction of the gas chromatographic effluent
with the decomposition products from a microwave discharge of SF6 in helium.

REFERENCES

1 R.K. Stevens and J.A. Hodgeson, Anal. Chem., 45 (1973) 443A.


2 A. Fontijn, A.J. Sabadell and R.J. Ronco, Anal. Chem., 42 (1970) 575.
3 P.N. Clough and B.A. Thrush, Tpans. Papaday soc., 63 (1967) 915.
4 D.H. Fine, D. Lieb and F. Rufeh, J. Chpomatogp., 107 (1975) 351.
5 A.L. Lafleur and K.ll. Mills, Anal. Chem., 53 (1981) 1202.
6 D.ll. Parees and S.R. Prescott, J. Chpomatogp., 205 (1981) 429.
7 T.J. Hansen, 11.C. Archer and S.R. Tannenbaum, Anal. Chem., 51 (1979) 1526.
8 I.A. Gough, K.S. Webb and R.F. Eaton, J. Chpomatogp., 137 (1977) 293.
9 D.H. Fine, F. Rufeh and B. Gunther, AnaZ. Lett., 6 (1973) 731.
10 TEA Uodel 502A Analyzep, Thermo Electron Corp., Waltham, MA, May 1980
11 K.S. Webb, I.A. Gough, A. Carrick and D. Hazelby, Anal. Chem., 51 (1979) 989.
12 D.H. Fine and D.P. Rounbehler, J. Chpomatogp., 109 (1975) 271.
13 D.H. Fine, F. Rifeh, D. Lieb and D.P. Rounbehler, AnaZ. Chem., 47 (1985) 1188.
180

14 D.H. Fine. D.P. Rounbehler and P.E. Oettinger. Anal. Chim. Acta. 78 (1975)
383.
15 J.H. Hotchkiss. J.F. Barbour, L.M. Libbey and R.A. Scanlan. J. Agr. Food
Chern., 26, (1978) 884.
16 E.A. Walker and M. Castegnaro. J. Chromatogr., 187 (1980) 229.
17 T.A. Gough and K.S. Webb, J. Chromatogr., 154 (1978) 234.
18 R.W. Stephany and P.L. Schuller, in B.J. Tinbergen and B. Krol (Editors).
Proceedings of 2nd International symposium on Nitrite in Meat Products,
Zeist, September 1976. Pudoc. Wageningen. 1977. p. 249.
19 G.V. Alliston, K.S. Webb and T.A. Gough, J. Chromatogr •• 175 (1979) 194.
20 K.S. Webb and T.A. Gough, J. Chromatogr •• 177 (1977) 349.
21 J.M. Douse. J. Chromatogr •• 256 (1983) 359.
22 D.H. Fine, W.C. Yu. U. Goff. E. Bender and D. Reutter, J. Forensic Sci.,
28 (1983) 29.
23 TEA Model 543 AnaZyzer. Thermo Electron Corp •• Waltham, MA. July 1980.
24 R.E. Parks, presented at 27th Pittsburgh Conference on Analytical Chemistry
and Applied Spectroscopy. Cleveland. OH. March 1976.
25 H.V. Drushel, Anal. Chern., 49 (1977) 932.
26 N. Kashihira, K. Kirita and Y. Watanabe, Bunseki Kagaku (Jap. AnaZ.). 29
(1980) 35.
27 N. Kashihira. K. Makino, K. Kirita and Y. Watanabe, J. Chromatogr., 239
(1982) 617.
28 The Antek Nitrogen AnaZyzers, Antek Instruments, Houston, TX, 1983.
29 TEA Model 610 Nitrogen Analyzer, Thermo Electron Corp •• Waltham, MA.
30 N. Kashihira, K. Makino, K. Kirita and Y. Watanabe, Bunseki Kagaku (Jap.
Anal.). 31 (1982) E13.
31 D.P. Rounbehler. S.J. Bradley. B.C. Challis, D.H. Fine and E.A. Walker.
Chromatographia. 16 (1982) 354.
32 W.B. Bruening and F.J.M. Concha. J. Chromatogr •• 112 (1975) 253.
33 W.B. Bruening and F.J.M. Concha, J. Chromatogr., 142 (1977) 191.
34 C.E.H. Bawn and A.G. Evans, Trans. Faraday Soc., 33 (1937) 1571.
35 S. Araki, S. Suzuki, M. Yamada, H. Suzuki and T. Hobo, J. Chromatogr. Sei.,
16 (1978) 249.
36 C.E.H. Bawn and R.F. Hunter. Trans. Faraday Soc., 34 (1938) 608.
37 C.E.H. Bawn and W.J. Dunning. Trans. Faraday Soa., 35 (1939) 185.
38 M. Yamada, A. Ishiwada, T. Hobo, S. Suzuki and S. Araki. J. Chromatogr.,
238 (1982) 347.
39 W.H. Duewe and D.W. Setser, J. Chern. Phys., 58 (1973) 2310.
40 D.E. Mann. B.A. Thrush, D.R. Lide. J.J. Ball and N. Acquista, J. Chern.
PhY8., 34 (1961) 420.
41 K.C. Kim and D.W. Setser, J. Phys. Chern •• 77 (1973) 2493.
42 D.J. Bogan and D.W. Setser, J. Chern. Phys., 64 (1976) 586.
43 D.J. Bogan, D.W. Setser and J.P. Sung, J. Phys. Chem., 81 (1977) 888.
44 D.J. Smith. D.W. Setser. K.D. Kim and D.J. Bogan, J. PhY8. Chem., 81 (1977)
898.
45 J.K. Nelson, R.H. Getty and J.W. Birks. AnaZ. Chern., 55 (1983) 1767.
46 R.H. Getty and J.W. Birks, Anal. Lett •• 12 (1979) 469.
181

Chapter 9

ELECTROLYTIC CONDUCTIVITY DETECTOR

CONTENTS

9.1. Detector construction. 181


9.2. Selectivity of response 1~
9.3. Response • 189
9.4. Solvent •• 196
9.5. Gases ••• 201
9.6. Temperature 203
References • • • 206

9.1. DETECTOR CONSTRUCTION

In 1962, Piringer and Pascalau 1 described an electrolytic conductivity de-


tector (ELCD) designed to determine organic compounds. In a quartz capillary
containing copper(II) oxide, the compounds eluted from the chromatographic.
column are decomposed, giving rise to carbon dioxide. The carrier gas with the
carbon dioxide produced contacts flowing deionized water in a long glass capil-
lary (about 15 cm). and the conductivity of the aqueous solution of carbon di-
oxide is measured. The minimum detectability with this detector was found 2 to
be 6'10- 11 g/ml of methane. For the determination of hydrogen and its iso-
?
topes, a modification of this detector was described" in which a PdC1 2-con-
taining reactor (80 0 C) was placed ahead of the contact space of the carrier gas
with water. Hydrogen chloride was produced; the detection limit was 4'10- 12 g
of the hydrogen isotope.
Sternberg et al. 4 in 1962 described an ELCD complementing the flame-ioniza-
tion detector (FlO). A particulate material with water flowing down on its
surface was positioned above the burner of the FlO. The decomposition products
from the FlO were trapped in the thin liquid film and the conductivity changes
of the liquid were measured. The detector is selective (Cl:C ~ 105; S:C ~ 10 4 );
the minimum detectable mass rate is 2.5.10- 10 g/sec of dichloromethane and
1.5.10- 9 g/sec of carbon disulphide. The selectivity relative to hydrocarbons
is given by the relatively low absorption capacity of carbon dioxide in the
rapidly flowing water and also by the low ionization of the carbon dioxide ab-
sorbed 5•
182

PYROLYSIS ZONE

INLET CATALYST
"
HELIUM -

~=~-LIQUID
VENT <;EPARATOR

PLATINUM
ELECTRODES

LIQUID
COLUMN

Fig. 9.1. Coulson electrolytic conductivity detector. (From ref. 6.)

The above selective detector was modified by Coulson 5 for routine work (see
Fig. 9.1). Under oxidizing or reducing conditions, the solute is decomposed in
the reaction space and the decomposition products are supplied to the space
where the gas phase comes into contact with the liquid (deionized water). The
decomposition products are absorbed in water, giving an electrolyte. The gas
phase is separated from the liquid phase in a separator. The change in the
water conductivity due to the electrolytes is sensed by platinum electrodes.
The distance between the electrodes of the glass detection cell is about 1 cm
and the cell geometry is optimized so as to display minimum polarization.
Coulson's detector was commercially produced until about 1974. Later, this de-
tector was modified by several workers. Under reducing conditions, the use of
a PTFE tube for the reactor outlet reduces the sorption losses of the ammonia
produced and increases the sensitivity of detection when using hydrogen as the
carrier gas and a nickel wire as the catalyst 7• The use of a water-jacket and
modified electrodes (larger electrodes, shorter distances between the elec-
trodes)8 also results in a higher detection sensitivity than with Coulson's
original design. A similar effect is obtained by reducing the amount of water
that enters the water-gas contact space 9 by inserting a wire into the capillary
feeding the water to the gas-water contact space (Fig. 9.2). The main disad-
vantage of Coulson's detector is its size: the detector has to be mounted on
the chromatograph as a separate unit, which requires heating of the conduit
leading to the gas chromatograph.
A microdetector (the Hall ELCD) (Tracor Instruments, Model 310) was de-
scribed by Hall 10 in 1974. This is substantially smaller and displays a higher
sensitivity than the Coulson detector (Fig. 9.3). The conductivity cell also
183

FURNACE HYDROSTATIC PRESSURE


EFFLUENT COLUMN

MIXING
CHAMBER TYGON
TUBING
DETECTOR ~~~~~~=~f--~WATER FROM ION-
CELL
If EXCHANGE RESIN

STAINLESS-STEEL
WIRE

RESERVOIR WATER
LEVEL

Fig. 9.2. Partial diagram of the Coulson detector with wire in place.
(Reprinted with permission from ref. 9.)

Fig. 9.3. t1icroelectrolytic conductivity detector cell assembly. 1 = Gas-liquid


contactor; 2 = PTFE solvent delivery tube; 3 = PTFE reaction products delivery
tube; 4 = stainless-steel detector block; 5 = solvent vent; 6 = PTFE insulator
sleeve; 7 = gas-liquid exit tube and center electrode.
(Reprinted with permission from ref. 11. Y
184

comprises the space for the separation of the gas and liquid phases. The contact
of the liquid and the gas from the reactor takes place in a small section of the
PTFE tube. The heterogeneous gas-liquid mixture produced enters the interior of
the conductivity cell where the phases separate on contacting the inner stainless-
steel detector wall. The liquid phase flows down on the cell surface, forming a
sheath with the gas phase as the core. In this manner, the liquid passes between
the inside wall of the outer electrode (cell wall) and the outside wall of the
inner electrode (gas outlet). Then the liquid again contacts the gas through the
port in the inner gas-outlet tube (E), where it mixes with the gas and flows out
through the bottom of the tube 12 • The electrolytic conductivity is measured by
means of a conductivity bridge with synchronous detection. The temperature of
the reaction chamber is adjustable, and the liquid is a non-aqueous solvent. The
adhesion of the liquid phase to the surface of the detector is critical with
regard to the function of the detector and, therefore, the material used for the
construction of the detector is an important factor. A stainless-steel surface is
more suitable than a glass surface and glass, in turn, is more suitable than
plastics' for the separation of gases from organic solvents. On the other hand,
a clean glass surface is slightly superior.to metal for the separation of a gas
from aqueous mixtures 10 •
A further improvement in the sensitivity of the ELCD is achieved by a bipolar
pulse differential detector 13 in which the conductivity of the solvent is measured
in the first detector cell and that of the solvent with the decomposition products
in the second cell. In this way, the factors particularly affecting the detector
noise (temperature variations, solvent purity) are common for both cells and be-
come reduced to the minimum by differential sensing. The classical determination
of electrolytic conductivity is performed by means of direct current or alternat-
ing current techniques. These methods, however, are limited because of electrode
polarization, capacitance effects and cell heating. The Hall 700A electrolytic
conductivity detector (Tracor Instruments, Austin, TX, U.S.A.) is based on bi-
polar pulse programmed to reach full amplitude for only one out of every ten
pulses. The output of the differential amplifier is caused only by the change in
conductivity produced by the gaseous reaction products formed in the microreactor.
A cross-section of the detector is shown in Fig. 9.4. Deactivated nickel reaction
tubes have been used for most applications; the microreactor also accepts, how-
ever, quartz reaction tubes. The conductivity solvent enters through the solvent
inlet and flows through the reference conductivity cell. The latter consists of
the top and outer electrode assemblies. Subsequently, the solvent flows into the
gas-liquid contactor where it is mixed with the gaseous reaction products enter-
ing through the gas inlet. The mixture is then separated in the gas-liquid sepa-
rator. The gas leaves through the hollow bottom electrode. The liquid flows be-
185

Top Electrode

Solvent Inlet Insulator

Gas Inlet Outer Electrode

Gas-Liq u id
Contactor
Gas-liquid
Separator
Solvent Exit Hole

Insulator
Bottom Electrode
Exit Line
Heating
Element

Reactor
Assembly

Reaction Gasc:f7W7W7¥IT1

Thermocouple

Column

Fig. 9.4. Cross-section of microreactor and differential cell. (Reprinted from


ref. 13, with permission.)

tween the outer wall of the bottom electrode and the inner wall of the gas-liquid
separator. The liquid passes into the hollow bottom electrode through a small
hole in the wall of this el~ctrode. At this point, the gas and liquid phases are
recombined and returned to the solvent reservoir.
This Hall 700A detector offers nitrogen, halogen, sulphur and nitrosamine
modes.
Recently, a new design of an ELCD was described 14 that is especially suitable
for high-resolution capillary gas chromatography. The phase separator is elimi-
nated and the gas-liquid mixture passes directly through the conductivity cell.
1~

9.2. SELECTIVITY OF RESPONSE

The ELCD can be used as a selective detector for halogen-, sulphur- and
nitrogen-containing compounds. The selectivity is determined by the products
formed in the detector reactor, the scrubber used and the chemical properties
of the electrolyte. The reaction furnace is operated under one of the following
conditions: reductive (hydrogen reaction gas), oxidative (oxygen or air reac-
tion gas) or pyrolytic (inert reaction gas). The post-furnace scrubber removes
acidic or basic products. The following cond~tions are essential for a selective
response: (I) the compounds of interest should give a decomposition product that
is soluble and ionized in the conductivity solvent; (2) the interfering com-
pound decompose to products that are either insoluble or non-ionized in the
solvent or can be chemically subtracted without interfering with the product of
interest; and (3) the dissolution and ionization processes should display discri-
mination against interferences that have not been removed previously.
Under reducing conditions (800-900 0 C) and when using a nickel catalyst, am-
monia is produced by the decomposition of organic nitrogen compounds:

Ni ,H 2
R-CN ----+ NH3 + lower alkanes (9.1)

Under these conditions, the organic halogen-, sulphur-, oxygen- and phosphorus-
containing compounds are converted into HX, H2S, H20, PH 3 and lower alkanes. The
water gives a low or no response, because it is already present in the aqueous
solvent. The lower alkanes (mainly methane) are poorly soluble in the solvent
under the given conditions and, in addition, they exhibit a low ionization.
Hydrogen sulphide gives a low response owing to the low ionization constant in
water 15 (see Table 9.1). Ammonia is the only base formed during the reductive
catalytic decomposition of the organic compounds. For this reason, the acidic
products HX and H2S can be removed with a basic scrubber containing, for instance,
strontium hydroxide, placed between the reaction chamber and the conductivity
cell (see Fig. 9.1). Under these conditions, the N:C response selectivity7,13,
15,17,18 is 10 4-10 6 (nitrogen mode).
In the reductive version in the absence of a catalyst, the organic nitrogen
compounds are converted into ammonia to only a small extent. Therefore, the de-
tector becomes selective to halogen compounds l9 • If a slightly acidic electro-
lyte, such as I-propanol or n-butanol, is used the response of the weak acid,
hydrogen sulphide, and that of the weak base, ammonia, is levelled l3 (halogen
mode).
187

TABLE 9.1

DISSOCIATION CONSTANTS OF SOME ACIDS AND AMMONIA

From ref. 16.

Species Dissociated Ka pK a Temperature


from (0C)

H2S0 3 HSOj 1.54.10- 2 1.81 18


HSOj S02- 1.02.10- 7 6.91
3
H2SO 4 HS0 4 1.20.10- 2
25
HS04
HN0 2
S02-
4
4.6.10- 4
1. 92
N0 2 4.3.10- 7
3.37 12.5
H2C0 3 HCOj 6.37 25
HCOj C02-
3 5.61.10- 11 10.25
H2S HS 9.1.10- 8 7.04 18
HS- S2- 1.1.10- 12 11.96
HCl Cl

NH 40H NH+ 1.97.10- 5* 4.75** 25


4

Pyrolysis with an inert gas without a catalyst (quartz reaction tube) gives
a selective response to certain types of nitrogen compounds. In the temperature
range 400-600 oC, ammonia is formed from the amines and nitrosamines, whereas
other types of organic nitrogen compounds produce little ammonia, if any20. The
selectivity is 21 1:10 7 and the selectivity relative to pyrazine exceeds 10 5•
Fig. 9.5 gives a comparison of the detector responses under (A) pyrolytic condi-
tions and (B) reducing conditions. The nitrosamine mode of the Hall 700A detector
operates under reducing conditions, however, in the absence of a catalyst (gold
reaction tube)22. At about 700 oC. the nitrosamine is reduced as follows:

(9.2)
188

LU
Vl
Z
A o
a.. 2
3 Vl
LU
a:
I
LU
Vl
Z
o
a..
Vl
LU
45
a:
"""'
Z
II z"""'

r I I
a 5 10 min a 5 '-
10 min

Fig. 9.5. Comparison of detector response in (A) pyrolytic mode (550°C) and (B)
reductive mode (B20 0 C). 1 = N-Nitrosodimethylamine; 2 = N-nitrosodiethylamine;
3 = N-nitrosodi-n-propylamine; 4 = N-nitrosodi-n-butylamine; 5 = N-nitroso-
piperidine. Stainless-steel column (3 m x 1/8 in. 0.0.) packed with 10% earbowax
20M-terephthalic acid on Gas-Chrom Q; temperature, programmed from 100 to 2I0 0 e
at 10 0 C/min. (From ref. 21.)

When using a 'fJater-n-propanol (1:1) electrolyte, the selectivity over nitrogenous


compo~nds ranges from 200 to 500. At temperatures above lOOoe and under non-
catalytic reducing conditions, the detector (Fig. 9.6) with a hydrogen flow-
rate of 5 ml/min gives a selective response to barbiturates 15 ,18, allowing
barbiturates to be distinguished from other drugs 23 in biological extracts.
Under oxidizing conditions S02 and S03' HX, N0 2 , CO, CO 2 and H20 are formed
by organic sulphur-, halogen- and nitrogen-containing compounds. Water and carbon
monoxide give a low or no response; the responses to carbon dioxide and nitrogen
dioxide are also low (short contact time with the solvent or use of a non-aqueous
solvent). The detector can be made selective to sulphur (sulphur mode) or halogen
compounds by using a scrubber displacing either HX (silver salt at lOOoC) or
189
50
h

40

~ 30
z
::>
>
cr:
<I
cr:
~ 20
cr:
<I

10

700 600 900 1000 ·C


Fig. 9.6. ELCD furnace temperature profiles for barbiturates (0- and S-barbs.)
and azobenzene in the reductive mode. 1 = Butabarbital; 2 = amobarbital;
3 = pentobarbital; 4 = secobarbital; 5 = thiopental; 6 = thiamylal; 7 = azo-
benzene. (From ref. 15.)

sulphur oxides (calcium oxide at 800 0 C; the oxygen is saturated with water vapour
before entering the reaction tube in order to hydrolyse the possibly produced
calcium chloride)24. When using ethanol as a solvent, the selectivity of Cl:C 10
is about 10 5 , that of S:C 10 ,13,25 is about 10 5 and that of S:C1 26 is about
3
5'10 .
Depending on the temperature of the decomposing unit (see section 9.6),
selective differentiation can sometimes be achieved between different types of
compounds containing the same heteroatom.
The detector can also be converted into the non-selective carbon mOde 13 •
Carbon is detected with a nickel reaction tube containing a platinum catalyst,
air as reaction gas, a potassium hydrogen carbonate scrubber and an aqueous
conductivity solvent. The detection limit of heptadecane is 6.10- 9 g.

9.3. RESPONSE

The sensitivity of the ELCD depends on the amount of the element investigated
(see eqn. 9.4) and, for this reason, it should be independent of the structure
of the compound. A direct proportionality between the response and the content
of the element in the molecule was indeed found for compounds decomposed under
oxidizing conditions 5 and for nitrogen compounds decomposed under catalytic
190

TABLE 9.2

RELATIVE RESPONSES OF N-NITROSAMINES

x = mean peak area (mm2); s = standard deviation; V = coefficient of variation;


P = 100' (peak areo. of given compound)/(peak area of NOMA); M = 100· (mol. wt.
of given compound)/(mol. wt. of NDMA). (From ref. 21.)

N-Nitroso derivative
-X S V P M

Dimethylamine (NDMA) 120 7.3 6.1 100 100


Diethylamine 106 6.7 6.3 87 73
Di-n-propyl ami ne 100 4.9 4.9 82 57
Di-n-butylamine 64 2.0 3.1 53 48
Piperi di ne 32 1.0 3.2 26 65
Pyrrolidine 17 1.3 7.7 14 74

TABLE 9.3

RESPONSE QUENCHING COMPARISON OF FLAME PHOTOMETRIC DETECTOR (FPD) AND ELCD

Reprinted from ref. 25 with permission.

Percentage of quenching FPD response (%) ELCD response (%)


compound in N2 matrix
H2S COS H2S COS

o (reference) 100.0 100.0 100.0 100.0


10% Carbon dioxide 95.7 100.0 89.6 97.2
10% Methane 74.3 72.1 92.4 100.0
50% Methane 66.5 66.8 90.0 100.0
1% Ethylene 83.8 93.3 99.0 95.0
10% Ethyl ene 70.2 75.4 94.4 96.0
50% Ethylene 38.2 70.4 77 .4 98.0

reduci ng conditi ons 5 ,27. However, a 1arger vari ance in the mol ar response was
found by Greenhalgh and Cochrane 28 • A chlorine atom in the aniline molecule
affects the response level in the N mode. Derivatives substituted in the para-
position give a lower response. No decrease in detector response was found as
more nitro groups were placed on the aromatic ring in chloronitroanilines 29 •
A nitrogen atom in the molecule of 'a sulphur compound decreases the response
in the S mode. whereas that of chlorine increases the response 30 • The structure
of the compound affects the response with barbiturates 15 • The response of thio-
barbiturates ;s lower than that of their oxygen analogues (Fig. 9.6). In the
191

h
em
0
° 1

p~
20

• : 3
2

10

/./: : A A

• 5
t.

'"
700 800 900 1000 °c
Fig. 9.7. ELCD furnace temperature profiles of phenothiazines in the N-selective
catalytic reducing mode. 1 = Trimeprazine tartrate; 2 = methdilazine hydro-
chloride; 3 = thioridazine hydrochloride; 4 = phenothiazine; 5 = 2-chloropheno-
thiazine. (From ref. 15.)

30

25

1/1
U
C

~ 20
::>
o
.c
c
1/1
C
::>
15
o
u
II
1/1
C

~ 10
1/1
II
a::

400 500 600 700 800 900 1000 1060·C

Fig. 9.B. Effect of furnace temperature on the response of three sulphur and
three chlorine compounds. BHC = 1.2.3.4.5.6-hexachlorocyclohexane. (From ref. 33.)
192

h
.-._k
30
N'
<>

20
A

10

500 600 700 800 900 'C

Fig. 9.9. ELCD temperature profiles of lmlpramine in the N-selective cata1 tic
reducing mode, illustrating differences in catalytic activity. (A) 500-9006C
sequence; (8) 550-900 oC sequence; (C) 900-550 oC sequence. (From ref. 15.)

pyrolysis version, the response of nitrosamines also depends on their struc-


ture 21 ,31, but there is no clear relationship between the relative nitrogen
content and the relative sensitivity for each nitrosamine 21 (Table 9.2). The
structure of the molecule also affects the sensitivity of detection in the Cl
mode. If the chlorine atom is in the para-position in mono-, di- and trichloro-
anilines, the response of this compound is always lower than that of the other
derivatives. The introduction of a nitro group into the chloroaniline molecule
appears to decrease the detector response 29 . The differences in the sensitiv-
ities seem to be given by the degree of conversion of the compounds in the re-
actor chamber. The extent of conversi on depends on the temperature in the de-
composition unit, and this dependence is affected by the structure of the com-
pound 15 ,2o,32-34 (see Figs 9.7-9.11).
If another compound is simultaneously present in the detector (hydrocarbons,
carbon dioxide), the sensitivity of the ELCD is decreased. It follows from
Table 9.3 that this decrease depends on the type of interfering compound (the
influence of carbon dioxide is less than that of hydrocarbons) and on the amount
of the in:erfering compound (the influence increasing amount). This decrease is
always lower than with the flame photometric detector.
193

100

80
~
8
CXl
C 60
:c01
'Qj
J:
~
cQJ 40
a..
~
0

20

0L-L-6~00~~~~7~00~~~~8~0~0~
Furnace Temperature (·C)

Fig. 9.10. Relative peak heights versus furnace temperature for PCBs, chlorinated
pesticides and reference compounds. 1 = Chlorocyclohexane; 2 = average for
pesticides (lindane, heptachlor epoxide, dieldrin, DDT, o-chlordane and
o,p'-DDT); 3 = Aroclor 1254; 4 = pentachloroanisole; 5 = a,3,4-trichloro-
toluene; 6 = average of eleven isomers of polychlorobenzene; 0, common to
4,5,6. (Reprinted with permission from ref. 32.)

Table 9.4 gives the detection limits for individual types of ELCD. The response
may decrease gradually during the operation of the detector, owing to contamina-
tion of the reaction tube, solvent impurities and the ion-exchange resin 10 ,21,32.
Silicon dioxide from the stationary phase is a source of contamination in the
oxidizing mode and can be washed out with 10% hydrogen fluoride. In the reducing
mode, contamination is mainly due to the condensation of non-volatile substances
in the cold section of the reaction tube and in the PTFE connecting tube (bleed-
ing from the septum, column, etc.). Carbon contamination is related to the inner
diameter of the tube. In the oxidizing mode, the injection of 3 ~l of hexane
severely contaminates the 4 mm 1.0. reaction tube, whereas no contamination was
observed with 0.5-1 mm 1.0. tubes even after 300 injections 10 •
The linear dynamic range is about four orders of magnitude 5 ,10 and >10 5 for
Pi ri nger and \~o lff IS detector 14 •
The electrolytic conductivity (c) of the solution is directly proportional
to the specific conductance (Csp ) of the solution and inversely proportional
to the cell constant 10 • The specific conductance depends on the kind of ion in
the solution and on the concentration and temperature of the latter. The cell
TABLE 9.4
-
..,.
\0

DETECTION LIMITS FOR ELCDs

Detector type Heteroatom Detection limit or Mode Solvent Ref.


minimum detectability

Coulson Cl 0.5 ng of lindane Oxidative Water 5,6


S 1 ng of Systox Oxidative Water 5
N 1 ng of azobenzene Reductive Water 35
Modified Coulson N 0.4 ng of nitrotoluene Reductive Water 7
2 pg//sec of N
Modified Coulson N 0.1 ng of atrazine Reductive Water 8
Hall 350 Cl 20 pg of lindane Oxidative Ethanol 10
20 pg of lindane Reductive Ethanol 10
Hall 700A S 1 pg/sec of S Oxidative Methanol 25
N 10 pg of nitrosamine Reductive Propanol-water (1: 1) 36
Hall 310 N 50 pg of dialkylnitrosamine Pyrolytic Water 21
Hall 700A S 0.4 pg/sec of S Oxidative Methanol 26
Hall 700A Cl 0.5 pg/sec of Cl Reductive 13
Pi ri nger and Wolff Cl 1. 5 pg/sec of Cl Reductive Water 14
195

70

60

50

f-
:J:
(.!) 40
w
:::c
::.::
i;5 30
n..

20

10

0
300 400 500 600 700 800 900
°c
Fig. 9.11. Effect of furnace temperature on the response of 3 ppm of ethyl
mercaptan (1).3 ppm of hydrogen sulphide (2), 3 ppm of dimethyl sulphide (3)
and 100% methane (4). (Reproduced from ref. 34 with permission.}

constant depends on the electrode area (A) and the distance between the elec-
trodes (d):

c = Csp /k (9.3)

k = d/A (9.4)

The detector response depends on the amount of the compound entering the de-
tector per unit of time (M), the weight percentage of the element monitored (w),
the conversion efficiency (E) of the sample, the solvent flow-rate (f), the cell
temperature (T c )' the voltage (V) and the cell constant:

(9.5)
196

where K is a constant and a is the coefficient of temperature dependence. The


cell with the largest electrode area per unit volume, i.e., with concentric
cylindrical electrodes, will give the greatest response. The theoretical de-
pendence of the response on the radius of the inner electrode and the inter-
electrode distance can be seen from Figs. 9.12 and 9.13 (the Hall detector).

o
o
~
N
~

• 0
~ 0
~
N

r 2r 3r 4r Sr 6r
2
Fig. 9.12. Calculated detector response versus radius of inner electrode (r).
Inter-electrode distance fixed at 0.0139 cm. (Reprinted from ref. 10 with
permission.)

The response of the Coulson detector increases by a factor of about 12 in the


voltage range 5-50 V. However, peak tailing and detector noise also increase at
the same time. The signal-to-noise ratio remains virtually constant 32 •

9.4. SOLVENT

Water undergoes autoprotolysis 19 :

(9.6)
197

~ d 2d 3d 4d 5d 6d
2

Fig. 9.13. Calculated detector response versus inter-electrode distance (d).


Radius of inner electrode fixed at 0.0826 cm. (Reprinted from ref. 10 with
permission.)

If an acid or a base is added to water, the autoprotolysis is suppressed. Hence


the increase in conductivity is smaller than would correspond to the addition of
an electrolyte alone:

(9.7)

(9.8)

The hydronium or hydroxyl ions formed shifts the equilibrium according to


eqn. 9.6 to the left. As a result of this phenomenon, a non-linear relationship
exists between the equivalent conductance and the electrolyte concentration at
low concentrations (Fig. 9.14). This problem can be essentially eliminated by
maintaining the water slightly basic by means of an ion exchanger 7• which con-
siderably reduces the effect of autoprotolysis. This problem can also be elim-
inated by using a non-aqueous solvent (ethanol, methanol, butanol)10,15.25.34.
198

10- 6 • 10-5 10- 4


MOLARITY OF ELECTROLYTE

Fig. 9.14. Conductivity (A, ~-1 cm- 1) of several electrolytes in water. (From
ref. 19.)

When using water, the CO~ response is about three orders of magnitude lower
than that of sulphur di/trioxide and hydrogen chloride. The use of absolute
ethanol almost eliminates the CO 2 response 10 • increasing the selectivity to 10 5
In the reducing mode, the use of ethanol increases the selectivity to halogen
response, because the H2S response is reduced. However, the absolute response
in water is about four times greater than that in absolute ethanol. Mixtures
of isopropanol and water 37 and n-propanol and water 36 have also been used as
solvents for the Hall detector. Hall recommends a hydrogen chloride-ethanol
mixture for the highly sensitive detection of nitrogen compounds 10 • Methano1 26
has been used for the oxidative decomposition of sulphur compounds (S mode).
l~hen using dilute hydrochloric aCid 38 , the conductivity of the solvent is
decreased:

(9.9)

The ionic conductance of the NH4+ ion is about five times smaller than that of
the H30+ ion ll . The response to nitrogen compounds in dilute hydrochloric acid
(5 ppm) is about four times greater than that in water. Hence the neutraliza-
tion reaction 9.9 produces greater changes in conductance than simple dissolu-
199

TABLE 9.5

COMPARISON OF DETECTION LIMITS

From ref. 38.

Compound Detection limit (ng)

Nitrobenzene 20
Chlorobenzene 1
Chloronaphthalene 0.5
Hexachlorobutadiene 0.05

tion and partial ionization of ammonia (reaction 9.8). When using dilute hydro-
chloric acid, the response to chlorine compounds is one to two orders of magni-
tude greater than that to nitrogen compounds (Table 9.5). Fig 9.15 shows a
chromatogram of nitrogen and chlorine compounds after reductive decomposition
using hydrochloric acid (1 ppm). All nitrogen compounds give negative peaks;
the conductance is reduced, whereas the responses to all the chlorine compounds
are positive, i.e., the conductance is increased. The peaks of compounds con-
taining both nitrogen and chlorine in the same molecule are doubled (this fact
was explained by the different rates of desorption of ammonia and hydrogen
chloride from the catalyst)38.

~
~ ~ c ~
c c ~ c
80 ~
~
~
~
c ~
~
.~ ~ ~
£ ~

~ 70 ~
0~ ~
~
~
c
~
0
~ 0 0 c ~
0
~ ~ ~ 0~
60 u ~
~
~
u 0 u
C I I
~ I
~
Q 0 U N
U
~
50
~
c
0
~
40
~
~ ~
~
30 c
~ ~
C N
c
~
~ ~
20 N 3 D
C ~I ~
~ 0 Z
Fig. 9.15. Chromatogram of nitrogen and chlorine compounds using 1 ppm of
hydrogen chloride. Stainless-steel column (2 m x 0.2 cm 1.0.) packed with
3% Carbowax 20M on Celite; temperature, programmed from 75 to 170 0 C at
7.5 0 C/min. (From ref. 38.)
200

In a dilute iodine solution, rapid oxidation of sulphur occurs, glvlng two


highly conducting hydrogen ions per molecule of oxidized hydrogen sulphide:

(9.10)

The response to sulphur compounds in a solution composed of 0.04% iodine,


2% ethanol and 1 ppm hydrogen chloride is about an order of magnitude higher
than that in a 1 ppm hydrogen chloride solution 38 • Hydrocarbons are not ionized.
However, the stability of the baseline is low and strong drifting occurs (the
solvent is rapidly contaminated and ion exchangers cannot be used). For this
reason, this technique is unsuitable in practic~34.
The response of the Hall detector decreases with increasing flow-rate of the
solvent. In the flow-rate range 1-2 ml/min, the atrazine response remains ap-
proximately constant. The response increases drastically if the flow-rate is
reduced below 1 ml/min. There is an approximately 10-fold increase in response 37
for a flow-rate of 0.2 ml/min. However, the detector noise increases in the same
proportion at low flow-rates. With regard to the noise, the optimum flow-rate is
0.7 ml/min 21 ,37. The flow-rate of the solvent determines the ion concentration
and, theoretically, the detector response should be inversely proportional to
the flow-rate of the solvent lO (eqn. 9.5). Hall 11 explained the non-linearity
of this relationship by the evaporation of the solvent and the limited capacity
of the gas-liquid contactor for exposing the entire solvent entering the cell to
the stream of the reaction products. When using a solvent consisting of iso-
propanol-water (50:50)11, the solvent composition changes owing to the partial
evaporation of the alcohol. The amount of solvent actually flowing through the
cell is also reduced. Hence the solvent polarity increases and the actual amount
of the solvent decreases if the flow-rate is decreased. The invariability of
the response at flow-rates exceeding 1 ml/min can be explained 11 by the assump-
tion that large solvent drops, rather than a fine mist, are formed at these flow-
rates. The solvent drops are expelled through the hole in the gas exit tube
without touching the electrode working surfaces. There is also a non-linear de-
crease in response with increasing flow-rate of the solvent (methanol) in the
oxidative determinations of sulphur compounds. The minimum detectability shows
a maximum at a flow-rate of 1.9 ml/min 26 •
The daily variations in the sensitivity were found to be due to the decrease
in the concentration and temperature variation of the conductivity solvent. When
using a mixed solvent, the isopropanol concentration and minimum detectability
decrease with time and the short-term noise increases. The use of an undiluted
alcohol as the conductivity solvent in conjunction with a temperature-controlled
water-jacket gives the most reproducible detector response from day to day39.
201

The solvent conductivity and the noise of the Coulson detector increase with
increasing temperature of the detection cell. The response of this detector
decreases by tens of percent 40 in the temperature range 8-45 0 C, whereas the
response of the Coulson detector modified by Lawrence and Moore 8 remains the
same.

9.5. GASES

Under reducing catalytic conditions, the response of the Hall detector to


atrazine increases drastically to a maximum at about 25 ml/min with increasing
hydrogen flow-rate (carrier gas helium) and then remains essentially constant
up to a flow-rate of 100 ml/min 37 • The influence of the hydrogen flow-rate is
low when using hydrogen as the carrier gas (Table 9.6)11. In the pyrolysis
version (carrier gas nitrogen), the response to barbiturates reveals a tem-
perature-dependent maximum between hydrogen flow-rates of 0 and 5 ml/minl1.

TABLE 9.6

INFLUENCE OF HYDROGEN REACTION GAS FLOW-RATE ON RESPONSE TO NITROGEN-CONTAINING


COMPOUNDS

Furnace temperature, BOOoC. Reprinted with permission from ref. 11.

Flow-rate (cm 3/min)* Peak height (mm)


Chlorpropham Atrazine Simazine

10 27 89 78
20 32 109 95
40 37 121 109
60 32 108 95
80 38 131 116
100 31 108 96

*Total flow also contained 40 cm 3/min of H2 carrier.

The Coulson detector shows an increase in the atrazine response (maximum


about 40 ml/min) similar to that of the Hall detector, but with a decrease in
response at hydrogen flow-rates exceeding 40 ml/min41.
In the oxidative version, oxygen flow-rates within the range 20-40 ml/min
were generally required in order to obtain the maximum response of the Coulson
detector to chlorine and sulphur compounds. The response profiles are dependent
202

on the structure of the compound 33 (Fig. 9.16). The detector noise increases
with increasing flow-rate of the air26.

30

25

Promctryne

,,
,,
Aldrin'-------___
, /' ,,/-
",'"
/.;1'
.......... --...... '......
~/
/'" -, /

I ,", ", ....~~BHC


/ I
II
1/
~
, Diethyl S-phenyl phosphorodithioate
,\

10 ':
"\ ",.,..--- ---------
I /'
I I Heptachlor
I I

," /
/

5 :

510 20 30 40 60 80 100 160ml/min

Fig. 9.16. Variation of the detector response with oxygen flow-rate at a furnace
temperature of 850 0 C. BHC = 1,2,3,4,5,6-hexachlorocyclohexane. (From ref. 33.)

At high separation speeds, such as occur in capillary gas chromatography,


especially at low solute concentrations, peak tailing was opserved as a result
of the low desorption rate of HCl from the surface of the nickel catalyst. The
doping of the auxiliary gas with vinyl chloride completely eliminated this
effect 14 (Fig. 9.17).
203

5 4

2
3
4

a b

Fig. 9.17. Effect of doping on the peak shape. a, Un doped gas; b, auxiliary gas
doped with 20 ppm of vinyl chloride. 1 = 4-Chlorobiphenyl; 2 3,4-dichlorobi-
phenyl; 3 = 3,5-dichlorobiphenyl; 4 = 4,4'-dichlorobiphenyl; 5 = 2,4,5-tri-
chlorobiphenyl. (From ref. 14.)

9.6. TEMPERATURE

With increasing temperature of the reaction cell, the response of the elec-
trolytic conductivity detector to nitrogen compounds (the Hall detector, Model
310, catalytic reducing mode) increases to the maximum l4 ,37. The temperature at
which the maximum response it attained depends on the compound to some extent
(see Fig. 9.7, for example) and varies within the range 700-900 oC. The course
of the temperature dependence of the response in the region that follows after
attaining the maximum (BOO-I000 oC) also depends on the structure of the compound.
For substituted phenotiazines, diphenylmethanes and tricyclic antidepressants,
the response remains constant 15 reaching the maximum value. The response to
atrazine decreases slightly in the range 900-100 oC36 • The response to barbitu-
rates, in the reducing mode without catalyst (poisoning of the catalyst occurs
if Ni is used) also decreases with increasing temperature after attaining the
maximum 15 (Fig. 9.6). The maximum response to N-nitrosamines is attained at
600 oC21 .
204

TABLE 9.7

INFLUENCE OF REACTION CONDITIONS ON RESPONSE OF COULSON DETECTOR TO MODEL


CQt·1POUNDS

Reprinted with permission form ref. 32.

Compound Relative response *


800 0 C 700 0C
H ** No H2*** H ** No H2***
2 2

Chlorobenzene 1.000 0.701 0.078 0.038


1.2-Dichlorobenzene 2.520 0.982 0.166 0.028
1.3-Dichlorobenzene 1.642 0.607 0.082 0.017
1.4-Dichlorobenzene 2.180 0.794 0.132 0.021
1.2.3-Trichlorobenzene 3.080 0.904 0.115 0.006
1.3.5-Trichlorobenzene 1.842 0.537 0.073 0.002
1.2.3.4-Tetrachlorobenzene 2.861 0.738 0.062 0.013
1.2.4.5-Tetrachlorobenzene 2.082 0.572 0.052 0.009
Pentachlorobenzene 1.495 0.861 0.026 0.012
Hexachlorobenzene 0.296 0.398 0.009 0.009
Chlorocyclohexane 22.168 27.815 21.806 26.449
1.2.3.4.5.6-Hexachlorocyclohexane 16.074 17.188 10.420 17.750
1-Chlorotetradecane 17.168 27.516 7.824 21.798
Pentachloroanisole 14.392 14.963 1.221 4.514
a.3.4-Trichlorotoluene 5.907 8.631 0.344 0.372

*Response expressed relative to chlorobenzene.


**80 ml/min of hydrogen reaction gas.
***80 ml/min of helium substituted to maintain constant residence time.

Similar relationships are also valid for the response of the Coulson de-
tector 18 •41 ; compared with the Hall detector, the difference consists in a
lower increase in the response in the range 500-900 0C (by a factor of 2-3)
and in the higher temperature required to obtain the maximum response (about
900 0C). Similar relationships can also be observed in the response to sulphur
and chlorine compounds in the pyrolysis version 31 (Fig. 9.8). However, the
response again increases at temperatures above 10000C.
Fig. 9.9 shows the temperature profile of the response to imipramine as a
function of the sequence in which the furnace temperature was varied. As can
be seen, the temperature profile of the response changes substantially, de-
pending on the starting temperature and on whether the temperature was grad-
ually increased or decreased. Pape et al. 15 ascribed this dependence to the
deactivation of the catalyst at low temperatures.
205

At a given temperature, the efficiency of the reactions giving hydrogen


chloride from polychlorinated biphenyls (PCBs) and chlorinated pesticides de-
pends on the dissociation energy of the C-Cl bond 32 • The dissociation energy
of the aromatic C-Cl bonds (PCBs) is 4-13 kcal/mole higher than that of th~
aliphatic C-Cl bonds (chlorinated pesticides). The dependence of the responses
on the temperature of the reductive decomposition differs for the two groups
of substances (Fig. 9.10). The response to pesticides in the temperature range
600-700 0 C is greater than that to PCBs, attaining the maximum at 800 0 C. With
increasing temperature, the response of PCBs continues to increase up to 900 0 C.
Table 9.7 gives the relative responses of several chlorine compounds at various
temperatures under reducing and/or pyrolysis conditions. It is obvious that the
elimination of hydrogen from the system considerably reduces the production of
hydrogen chloride from chlorobenzene, whereas it is increased with chlorinated
aliphatics. Hence a selective response to chlorinated pesticides relative to
PCBs can be attained. The selectivity and sensitivity are shown in Table 9.8.

TABLE 9.8

SELECTIVITY OF CHLORINATED HYDROCARBON PESTICIDES OVER PCBs AS A FUNCTION OF


REACTOR TEMPERATURE

Reprinted with permission from ref. 32.

Temperature Detecti on 1i mi t Selectivity


(OC)
Chlorinated PCBs
pesticides

610 5-10 ng >50 119


710 1-5 ng >5 119
810 0.2-0.5 n9 10-20 n9

Fig. 9.11 shows the temperature dependence of the sulphur response in the
oxidative mode. A selective response to mercaptans 34 can be obtained at various
temperatures.
The detector response depends on the temperature of the column if chromato-
graphic columns with a stationary phase containing nitrogen (e.g., the cyano-
propylsilico,ne polymer OV-225) are used. The resistance of the solvent either
decreases (water) or increases (dilute hydrochloric acid) with temperature owing
to an increase in the rate of dissolution of ammonia in the liquid (influence
of stationary phase bleeding)42. Low loadings should be applied with these
206

phases. The use of phases containing nitrogen or a halogen should be avoided,


as they can give rise to acidic or basic products that would "poison" the
electrolyte and drastically reduce the sensitivity43.44. Columns deactivated
with phosphoric acid should also be avoided, especially at temperatures of
I80 0C or higher. The phosphoric acid bleed has a great adverse effect on the
nickel reaction tube and on the alkaline scrubber.

REFERENCES

1 O. Piringer and M. Pascalau, J. Chromatogr., 8 {1962} 410.


2 O. Piringer. E. Tataru and M. Pascalau. J. Gas Chromatogr •• 2 (1964) 104.
3 M. l1ohnke, O. Piringer and E. Tataru, J. Gas Chromatogr., 6 (1968) 117.
4 J.C. Sternberg, D.T.J. Jones and R.A. Morris, A Method for SeLective Halogen
and Detection in Gas Chromatography, presented at 1.3th Pi t tsburgh
Confepence on Analytical Chemist~J and Applied Spectroscopy, March 1962.
5 D.M. Coulson, J. Gas Chromatogr •• 3 (1965) 134.
6 D.M, Coulson. The DeTektor. 1, No.1 (1968). f1ikro Tek Instrument. Austin,
TX.
7 G.G. Patchett, J. Chromatogr. Sai •• 8 (1970) 155.
8 J.F. Lawrence and A.H. tloore, Anal. Chem •• 46 (1974) 755.
9 J.F. Lawrence and N.P. Sen, Anal. Chern., 47 (1975) 367.
10 R,C. Hall, J. Chromatogr. Sci •• 12 (1974) 152.
11 R.C. Hall. Crit. Rev. Anat. Chern., 8 (1978) 323.
12 R.C. Hall, U.S. Pat., No.3 934 193 (1976).
13 R.J. Anderson and R.C. Hall, Amer. Lab., 12 (1980) 108.
14 O. Piringer and E. Wolff. J. • 284 (1984) 373.
15 B.E! Pape. D.H. Rodgers and T.C. Flynn. J. Chromatogl"., 134 (1977) 1.
16 J. Sevcik, Deteatol"S in Gas Chromatography, Elsevier, Amsterdam, 1976,
p. 182.
17 S.L Pape and t4.A. Ribick, J. Chr·omatogl" •• 136 (1977) 127.
18 R.C. Han and C.A. Risk, J. Sai., 13 (1975) 519.
19 a.M. Coulson, lh:-tmgen, and Cal"bon Detection Electroehemical
Methods, presented at Eastern Symposium, Net.) YOl?k, 1968.
20 J.II. Rhoades and D.L Johnson, J. Sci., 8 (1970) 616.
21 E. von Rappard, G. Eisenbrand and R. Preussmann. J. Chl"omatogr •• 124 (1976)
247.
22 R.J. Anderson, Semi-specific Detepmination of llitl"osamines
Eleotl"olytic Cond:>J.ctivity Deteato'l'~ presented at on
AnaZytioal Chemist~ and Applied'Speatl"oscoPY, 1980, paper 412.
23 S.E. Pape, C~in. Chern., 22 (1976) 739.
24 D.H. Coul son, Advan. Chromatogl"., 3 (1966) 197.
25 B.J. Ehrlich. R.C. Hall, R.J. Anderson and H.G. Cox, J. Chl"Omatogl". Sci ••
19 (1981) 245.
26 S. Gluck, J. Ch~omatog~. Sci., 20 (1982) 103.
27 J.F. Palframan, J. Macnab and N.T. Crosby. J. Chromatogr •• 76 (1973) 307.
28 R. Greenhalgh and W.P. Cochrane, J. Chromatogr., 70 (1972) 37.
29 V. Lopez-Avila and R. Northcutt. J. Reso2ut. Chromatogr, Chrornatogr>.
Commun •• 5 (1982) 67.
30 W.P. Co~hrane and R. Greenhalgh. Int. J. Environ. Anal. Chern •• 3 (1974) 199.
31 p. Issenberg and S.R. Tannenbaum. presented at JARG Meeting on Analysis and
Formation of (Ii tY'osamines ~ Heide lber>g. 1971.
32 J.W. Dolan and R.C. Hall, Anat. Chern •• 45 (1973) 2198.
33 W.P. Cochrane, B.P. \-I11son and R. Greenhalgh, J. Chl'omatogr., 75 (1973) 207.
34 R.G. Schiller and R.B. Bronsky, J. Chromatogr. Sci •• 15 (1977) 541.
207

35 D.t4. Coulson, J. Gas ChY'Omatogl'., 4 (1966) 285.


36 R.J. Anderson, Tr'acor' Chl'omatogl'aphy. Appl. 79-3. N Selective Detection in
Gas Chl'omatogl'aphy, Tracor Instruments, Austin, TX, 1979.
37 B.P. I~ilson and W.P. Cochrane, J. ChY'Omatogl'., 106 (1975) 174.
38 P. Jones and G. Nickless, J. Chl'omatogl'., 73 (1972) 19.
39 R.K.S. Goo, H. Kanai, V. Inouye and H. Wakatsuki, Anal. Chem., 52 (1980)
1003.
40 G. Winnett and \~.L. Illingsworth, J. ChY'Omatogr·. Sci., 14 (1976) 255.
41 J.F. Lawrence, J. Chl'omatogl'., 87 (1973) 333.
42 D.tt Hailey, A.G. Howard and G. Nickless, J. ChY'Omatogr>., 100 (1974) 49.
43 ttA. Luke, J.E. Froberg, G.t1. Doose and H.T. Masumoto, J. Ass. Offic. Anal.
Chem., 64 (1981) 1187.
44 Tl'acol' Chl'omatogl'aphy, Appl. 78-5, Selective Detection in Gas Chl'omatogl'aphy,
Tracor Instruments, Austin, TX, 1978.
This Page Intentionally Left Blank
209

Chapter 10

COULOMETRIC DETECTOR

CONTENTS

10.1. Introduction • . . . • 209


10.2. Response • • . • • • • 210
10.2.1. Oxidative mode 210
10.2.2. Reductive mode 210
10.2.3. Nitrogen mode 211
10.3. Quantitative results 212
References • • • • • • • . . 215

10.1. INTRODUCTION

The column effluent is mixed with oxygen in a combustion tube where organic
compounds are converted into CO 2 , H20, HX, S02 and oxides of nitrogen. H2S, HX,
and PH 3 conversion products are obtained in the reductive mode, when using
hydrogen as the reaction gas. The resulting products then enter a titration
cell where they are absorbed in an appropriate solution and titrated automa-
tically with coulometrically generated ions. Thus, the coulometric detector.
(CD) consists of three parts: the combustion tube, the titration cell and the
coulometer (Fig. 10.1). The titration cell consists of four electrodes that
function as a sensor-reference pair and an anode-cathode generator pair. The
input signal from the sensing electrode is the difference between the sensor

Gas Combustion
Chromatograph

Recorder Coulometer

Fig. 10.1. Schematic diagram of the coulometric detector.


210

and reference electrodes and it is biased so as to give a zero signal across


the input of the amplifier for a given concentration of the titrant ion in the
titration cell. As this titrant concentration is decreased by the conversion
products, additional titrant is generated to maintain a balance. For sulphur
compounds, which yield sulphur dioxide in the combustion tube, the titrant is
generated in a iodide solution at a platinum anode. For halides, the titrant
is generated at a silver anode.

10.2. RESPONSE

10.2.1. Oxidative mode

In the oxidative mode, oxygen reactant gas is fed into the combustion tube of
the detector. Oxidative degradation is most commonly followed by iodimetric
titration of sulphur dioxide with coulometrically generated iodine l - 8 :

(10.1)

Bromine can also be used as a titrant 9- 11 . Large samples of compounds that


produce strong oxidizing agents during combustion, e.g., compounds containing
nitrogen, chlorine and bromine (with iodine titrants), give negative responses
because they are stronger oxidizers than the titrant 2 ,6,9. The detection limit
is 10-8 g for sulphur and the linear dynamic range covers three orders of
magnitude 2 •
For halides, internally generated silver ions serve as the titrant 4 ,8,12-18.
Organic bromides are converted into bromine and the latter yields half the
response of the hydrogen halide owing to the hydrolysis of Br 2 to HOBr and
HBr19; only the latter precipitates silver ions 15 . The detection limit is
approximately 1.10- 9 g of chlorine 15 ,17. Sulphur compounds can also be de-
termined without combustion. However, it is necessary to standardize the titra-
tion cell against each type of compound 20
Sevcik 21 described a combination of a CO and a flame-ionization detector
(FlO). The FlO served as a burning space for the substances eluted. It is
possible to determine compounds containing sulphur and chlorine in this way,
the minimum detectability being 8.10- 9 and 1.10-8 g/sec, respectively.

10.2.2. Reductive mode

In the reductive mode, sulphur compounds are reduced 20 to hydrogen sulphide.


The sample is pyrolysed6 in hydrogen at about 11500 C over a catalyst (e.g.,
211

10% platinum on Alundum). The hydrogen sulphide formed by the pyrolysis is


automatically titrated with silver ions in a microcoulometric cell:

(10.2)

Ag + Ag + + e- (10.3)

The sulphur reduction method suffers from nitrogen interference, the extent of
which is proportional to the content of hydrogen cyanide formed by pyrolysis.
The reductive method can also be used for phosphorus compounds 22 ,23. In
the oxidative method the phosphate moiety is probably converted to P4010 in
the combustion tube, but it fails to leave the tube. Hence, the column effluent
is reduced with hydrogen at 950 0 C with the conversion of phosphates to phosphine.
Organically bound sulphur and chlorine are converted to hydrogen sulphide and
hydrogen chloride, respectively. These three gases precipitate silver ions.
They are measured with relative sensitivities of 2:2:1. However, a scrubber
containing aluminium oxide quantitatively subtracts hydrogen sulphide and
-chloride, whereas phosphine passes through the packing unchanged. The tempera-
ture range for the optimum yield of phosphine is 925-1000 0 C. Sulphur bonded
directly to phosphorus can be measured directly without interference from
phosphine in the reductive mode at a lower temperature. Compounds that contain
only phosphorus do not yield any response at 70U oC. When sulphur is bonded to
carbon, the yield of hydrogen sulphide is low at this temperature.
Mercaptans can be determined directly in the silver cell without combus-
tion 24 - 26 .

10.2.3. Nitrogen mode

Nitrogen compounds are converted into ammonia in a stream of hydrogen over


a catalyst. The ammonia is automatically titrated in the titration cell to a
constant pH in a sodium sulphate solution with hydrogen ions:

(10.4)

(10.5)

The decrease in hydrogen ion concentration is sensed by the sensor-reference


electrode pair. The second electrode pair serves for hydrogen ion generation.
The titrant ion thus restores the original titrant ion concentration:
212

+ - ( 10.6)
H2 ~ 2 H + 2 e

(10.7)

Nickel deposited on magnesium oxide of the type described by Ter Meulen 27 was
used as catalyst 28 ,29. The temperature limit for this catalyst is 440-450 oC;
at higher temperatures nickel turnings or granules are used 29 - 31 . As the
detector response is given by the concentration change of hydrogen ions, any
substances that can change this concentration will produce a response. Acidic
compounds, such as hydrogen chloride or hydrogen sulphide (from sulphur- and
halogen-containing compounds) are removed by absorption methods. A hot
(350-450 oC, to allow the ammonia formed to pass quantitatively) alkaline
scrubber is used 30 ,31 with nickel catalysts. With the Ter Meulen catalyst a
separate scrubber is not needed because of the alkalinity of the magnesium
oxide catalyst support.
The detection limit is approximately 3 ng for nitrogen and the linear dynamic
range covers three orders of magnitude 32 • The selectivity relative to hydro-
carbons is 106 , and relative to other elements such as halogens and sulphur it
is at least 10 4 .

10.3. QUANTITATIVE RESULTS

The CD is theoretically a quantitative detector based on the amount of


electricity required for the internal generation of the titrant. Coulometric
titration proceeds according to Faraday's law, and calculation of the results

TABLE 10.1

CONVERSION OF SULPHUR COMPOUNDS TO SULPHUR DIOXIDE AT DIFFERENT TEf4PERATURES

Reprinted with permission from ref. 2.

Combustion tube Conversion


temperature (OC) to sulphur dioxide (%)

550 70
660 80
650 91
700 93
750 89
850 74
950 63
TABLE 10.2

MICROCOULOMETRIC DETECTION OF HALOGEN COMPOUNDS

From ref. 18.

Reactant Pyrolysis tube Recovery ± S.D. (%) (n = 3)


gas temperature (oC)* CHC1 3 C1CH 2CH{1 CHBr 3 CH Br 2
2
(66 ng Cl-)** (53 ng Cl-)** (137 n9 Br-)** (138 ng Br-)**
O2 820 42 ± 1.5 41 ± 2.1 82 ± 2.6 77 ± 1.7
920 47 ± 1.2 73 ± 1.5 55 ± 3.6 53 ± 4.2
1020 47 ± 1. 0 64 ± 2.5 39 ± 4.7 43 ± 2.1
CO 2 820 40 ± 2.0 45 ± 5.0 68 ± 3.8 79 ± 4.5
920 48 ± 2.3 76 ± 1.5 82 ± 2.3 91 ± 2.3
1020 66 ± 3.0 92 ± 3.5 81 ± 2.9 87 ± 0.6

*
**Deviation from specified temperatures ±10%.
Amount equivalent to complete conversion of organic halogen compound to titratable halides.

N
W
TABLE 10.3

MICROCOULOMETRIC DETECTION OF HALOGENATED BENZENES WITH OXYGEN AS THE REACTANT GAS

From ref. 18.

Pyrolysis tube Recovery ± S.D. (%) (n = 3)


temperature (oC) Chlorobenzene Bromobenzene 1,2,4-Trichlorobenzene 1,2,4-Tribromobenzene
(87 ng Cl-)** (186 ng Br-)** (36 ng Cl-)** (175 ng Br-)**
820 5 5 5 5
870 14 ± 0.6 16 ± 1.2 3 ± 1.2 10 ± 2.5
920 28 ± 3.6 47 ± 5.7 18 ± 3.2 40 ± 2.9
1020 26 ± 2.1 50 ± 1.0 26 ± 1. 7 47 ± 4.4

*
**Deviation from specified temperatures ±10%.
Amount equivalent to complete conversion of organic halogen compound to titratable halides.
215

is carried out in terms of coulombs required for the titration divided by the
Faraday constant, i.e., 96 500 coulombs/equivalent 13 . Hence the amount of
titratable material in an eluted peak is given by

(10.8)

where I is the titrant generator current in amperes and t the time in seconds 17 .
The results are dependent, of course, on the level of conversion in the
combustion tube. The conversion of sulphur compounds into sulphur dioxide
depends on the temperature 2 ,4,9 (Table 10.1). 'High temperature favours the
formation of sulphur dioxide rather than sulphur trioxide. It is not necessary
for the conversion to sulphur dioxide to be quantitative, it is only necessary
that the 50 2/50 3 ratio be kept constant 9 . The pyrolysis efficiency of halogen
compounds changes with temperature, reactant gas and the type of compound 4 ,18
(Table 10.2) and gas flow-rate 4. For CHC1 3 and C1CH 2CH 2Cl with oxygen as the
pyrolysis gas, the recovery increases with increasing pyrolysis tube temperature.
The opposite effect occurs for CHBr 3 and CH 2Br 2 • When carbon dioxide is used as
the pyrolysis gas the temperature dependence is stronger for both types of
compounds 18 . For both chlorinated and brominated benzenes the recoveries in-
crease with increasing pyrolysis tube temperature; the recovery of bromobenzenes
is better (Table 10.3).
The microcoulometric detector suffers from a number of disadvantages. The CD
system is relatively complex and difficult to operate when attached to a gas
chromatograph. The peaks are broad and tailing and the time constant is higher
than with other detectors2,3,21 ,32. The catalytic properties of the catalyst
change with time, as the catalyst becomes poisoned by condensed aromatics and
sulphur compounds 28 •

REFERENCES

D.M. Coulson, L.A. Cavanagh, J.E. DeVries and B. Walther, Agr. Food Chern., 8
(1960) 399.
2 R.L. Martin and J.A. Grant, Anal. Chern., 37 (1965) 644.
3 H.V. Drushel, Anal. Chern., 41 (1969) 569.
4 L. Giuffrida and N.F. Ives, J. Ass. Offie. Anal. Chern., 52 (1969) 541.
5 5.1. Kricmar and V.E. Stepanenko, Zh. Anal. Khim., 24 (1969) 1874.
6 L.D. Wallace, D.W. Kohlenberger, R.J. Joyce, R.T. Moore, M.E. Riddle and
J.A. McNulty, Anal. Chem., 42 (1970) 387.
7 V.E. Stepanenko and 5.1. Kricmar, Zh. Anal. Khim., 26 (1971) 147.
8 D.M. Coulson, J. Forensie Sei., 17 (1972) 678.
9 P.J. Klass, Anal. Chern., 33 (1961) 1851.
10 D.F. Adams and R.K. Koppe, J. Air Pollut. Control. Ass., 17 (1967) 161.
11 R.J. Robertus and M.J. Schaer, Environ. Sei. Teehnol., 7 (1973) 849.
12 D.M. Coulson and L.A. Cavanagh, Anal. Chern., 32 (1960) 1245.
216

13 D.M. Coulson and L.A. Cavanagh, Theory and Equipment for Microcoulometric
Gas Chromatography, presented at 140 Meeting of the American Chemical Society,
Division of Analytical Chemistry, Chicago, IL, September 1961.
14 J. Burke and W. Holswade, J. Ass. Offic. Anal. Chem., 47 (1964) 845.
15 H.P. Burchfield and R.J. Wheeler, J. Ass. Offic. Anal. Chem., 49 (1966) 651.
16 H.P. Burchfield, J.W. Rhoades and R.J. Wheeler, in L.R. Mattick and H.A.
Szymanski, Lectures on Gas Chromatography 1964, Plenum Press, New York, 1965,
p. 59.
17 D.M. Coulson, Nitrogen, Halide, Sulfur and Carbon Detection by Electrochemical
Methods, presented at Eastern Analytical Symposium, New York, November 1968.
18 J.A. Sweetman and E.A. Boettner, J. Chromatogr., 212 (1981) 115.
19 E.E. Storrs and H.P. Burchfield, Contrib. Boyce Thompson Inst., 21 (1962) 423.
20 D.F. Adams, G.A. Jensen, J.P. Steadman, R.K. Koppe and T.J. Robertson, Anal.
Chem., 38 (1966) 1094.
21 J. Sevcfk, Chromatographia, 4 (1971) 102.
22 H.P. Burchfield, J.W. Rhoades and R.J. Wheeler, Agr. Food Chem., 13 (1965)
511.
23 H.P. Burchfield, D.E. Johnson, J.W. Rhoades and R.J. Wheeler, J. Gas
Chromatogr., 3 (1965) 28.
24 A. Liberti, Anal. Chim. Acta, 17 (1957) 247.
25 E.M. Fredericks and G.A. Harlow, Anal. Chem., 36 (1964) 263.
26 V.T. Brand and D.A. Keyworth, Anal. Chem., 37 (1965) 1424.
27 H. ter Meulen, Recl. Trav. Chim. Pays-Bas, 43 (1924) 1248.
28 R.L. Martin, Anal. Chem., 38 (1966) 1209.
29 D.K. Albert, Anal. Chem., 39 (1967) 1113.
30 R.F. Cook, R.P. Stanovick and C.C. Cassil, Agr. Food Chem., 17 (1969) 277.
31 C.C. Cassil, R.P. Stanovick and R.F. Cook, Residue Rev., 26 (1969) 63.
32 R.C. Hall, CRC Rev. Anal. Chem., December (1978) 323.
217

ChapteY' 11

ELECTRON-CAPTURE DETECTOR

CONTENTS

11.1. Introduction............. 217


11.2. Design... . . . . . • . . . . . . • 218
11.3. Sources of primary electrons. . . . . 219
11.4. Methods of measuri ng detector current 224
11.4.1. Direct-current mode •. . . . . • 224
11.4.2. Pulse mode with constant frequency 225
11.4.3. Pulse mode with constant current 227
11.4.4. Other modes .• . . . . . . 229
11.5. Response theory ••• • • • • . • . 230
11.5.1. Recombination theory. . . . 230
11.5.2. Positive space-charge model 233
11.5.3. Negative space-charge theory 234
11 .6. Response.............. 235
11.6.1. Effect of compound structure. 235
11.6.2. Effect of detector temperature •• 239
11.6.3. Derivatization for electron-capture detection 243
11.6.4. Effect of impurities. . . . . • . . . . 248
11.7. Linearity of response • . . . . . . . • . . . . 249
11.8. Selective electron-capture sensitization. . . . 251
11.8.1. Nitrous oxide doping of the carrier gas 252
11.8.2. Oxygen doping of the carrier gas. . . . 257
11.8.3. Sensitization of aromatic hydrocarbons. 263
11.9. Coulometric and hypercoulometric response . . . . . . • • 263
11.10. Use of the electron-capture detector with capillary columns 266
References . . . . • • . • • . • . • • . . . . . . • • • . 269

11.1. INTRODUCTION

The electron-capture detector (ECD) is the oldest of the selective detectors.


Owing to its high sensitivity, which is the highest of all gas chromatographic
detectors, it is highly attractive for chromatographers. The number of original
papers dealing with this detector is very large, and it is included in all re-
views on detectors l - 8 . An excellent book 9 is devoted to particular aspects of
the ECD. For this reason, it is difficult to treat the topic of ECDs in an ex-
haustive manner without repeating material that is already well known.
218

11.2. DESIGN

The basic arrangement of the ECD consists of an ionization chamber containing


a source of particles (generally a radioactive source) and two polarized
electrodes. By applying a potential difference to the ECD electrodes it is pos-
sible to collect the thermal electrons.
The following detector types can be distinguished according to the position
and shape of the electrodes. The parallel-plate detector 10 - 13 (Fig. 11.1A) has
a very simple geometry and its design enables the relative configuration of the
radioactive foil and the collector electrode to be changed. The volume of this
detector is large, particularly if a 63Ni source is used, the specific activity
of which is lower as than that of tritium. Therefore, a large foil area is
required with 63Ni . This foil is more easily accommodated in the coaxial
design 15 ,16, (Fig. 11.1B) where the anode is positioned inside the cylinder

A B
63 Ni FOIL

PTFE
c
L . . - _...... Q

Fig. 11.1. Basic designs of the electron-capture detector. A, Parallel-plate


detector. 1 = Carrier gas inlet and anode; 2 = diffuser; 3 = source of ionizing
radiation; 4 = carrier gas outlet and cathode. B, coaxial detector. C, asymmetric
(pin-cup) detector. a = Anode; c = cathode; s = source. (From ref. 14.)
219

INSULA
DETECTOR
TOWER CAP

FOILCYLINDER

DETECTOR
ASSEMBLY
COLLECTOR
CYLINDER

DETECTOR - - - -
TOWER

'--NA,RRC)W CLIP

~ SWAGE FERRULE

Fig. 11.2. Asymmetric (displaced coaxial cylinder) electron-capture detector.


(From ref. 14.)

formed by the source. The overall detector volume is 2-4 ml with the higher
energy and high-temperature 63Ni source. In the asymmetric configuration 17 - 22
(Figs. 11.1C and 11.2), the cylindrical cathode, which may serve as the detec-
tor body, is separated by a glass, ceramic or PTFE insulator from a small anode.
If the system of electrodes is arranged so as to maximize their spacing, the
applied electric field is longitudinally asymmetric and minimizes the effect
of the positive space charge by concentrating the field in the vicinity of the
anode, while making the field near the cathode less intense 14 . In the displaced
coaxial cylinder design, the cell geometry makes direct collisions of particles
with the anode unlikely. Smaller diameters are possible, provided the collisions
of particles with the radioactive source itself are minimized 23 . A cell with
a total volume of 0.3 ml has been described 22 •

11. 3. SOURCES OF PRII~ARY ELECTRONS

As a rule, the ECD utilizes a radioactive emitter, generally in the form


of a metal foil, as a source of primary ionizing particles. An a-emitter, which
would produce 10 5 ion pairs per 1 cm of travel, would generate high detector
noise. The ion-pair yield per 1 cm through a gas from a very high-activity
y-emitter is extremely low and meets the requirements for an ideal source.
However, the health hazard inherent in such an emitter excludes it from use.
N
N
o

TABLE 11. 1

SOME PROPERTIES OF ELECTRON SOURCES FOR THE ECD

From ref. 23.

Properties Source
63 Ni foil 3H titanium 3H in scandium 55 Fe on nickel
foil foil alloy foil
B-Particle energy (keV) 66 18 18 5.387-5.640
B-Particle range (mm) -10 -2.5 -2.5 -0.5
Maximum activity (mCi/cm2 ) 10 170 3
Upper temperature limit (OC) 350 220 325 400
Maximum current (pA) 9 30 0.5
(15-mCi source) (500-mCi source) (5-mCi source)
Rate of electron production, Rv (sec- 1 ) 6.10 10 2.10 11 3.10 9
Noise level* (pA) 1.5 3 0.1

*Measured at ambient pressure in nitrogen at 21 0 C.


221

The best compromise was found in isotopes that emit low-energy S-particles
(minimum number of ion pairs per disintegration) at relatively high specific
activities (maximum total ion pair formation)16.
The choice of an irradiation source is governed 3 by (1) the emanation rate
of the radioactive material at elevated temperatures, (2) the energy of the
radioactive particle, (3) the availability of adequate specific activity, (4)
the radiochemical form and (5) the costs. 3H and 63Ni are the most frequently
used materials. Tritium is usually preferred owing to its lower energy S-radia-
tion (18 keV for 3H, 67 keV for 63 Ni )24,25 and the fact that foils of higher
specific activities which provide a denser radiation (9800 and 65 Ci/g for 3H
and 63Ni , respectively)26 can be manufactured. The disadvantage of tritium
sources is their low operational stability owing to a loss of activity at
elevated temperatures. The temperature limit for tritium is 200-225 0 C in the
case of a titanium- 3H foi1 24 ,27,28 and 300-325 0 C for tritium embedded in a
rare earth, e.g., scandium 16 ,20. The maximum temperature that can be used with
63Ni is as high as 400 0 C15 , however. The temperature limits quoted for the
tritium source are valid for nitrogen, helium, argon and argon plus 5% methane
carrier gases. When hydrogen is used, the tritium emanation rate is as much
as ten times higher owing to the exchange between hydrogen gas and bound
tritium 20 ,28. The use of hydrogen as the carrier gas is not recommended, there-
fore, particularly at higher detector temperatures. The application of high
temperatures in the detector has two aspects: (1) it decreases the possible
contamination of the source (for this reason 63 Ni sources are preferred in
practice) and (2) the detector sensitivity increases or decreases depending on
the nature of the capturing process (see section 11.6.2).
147 pm29 and 99 Tc 24 have also been used as ionization sources. The properties
of the ECD with promethium are similar to those of the nickel detector, but
the 147 pm foil is much less affected by the nature of the sample 30 . In comparison
with other emitters, technetium shows 24 a disadvantageous signal-to-noise ratio.
Hence the use of radioactive sources has several disadvantages 31 : (1) the
radioactive source can be contaminated by column bleed or by compounds of low
volatility eluted from the GC column, (2) the radioactive sources have an upper
temperature limit related to the thermal stability of the foil, (3) the radio-
active metal foil appears to react with the electron-capturing species, as
evidenced by the discoloration of the surface of the foil after continued use
and (4) the disadvantages arising directly from the application of the radioac-
tive emitter proper.
The Auger electron emitter with 55Fe has been employed 32 as a source that
gives low noise with an operating current lower that that with S-emitters. In
222

this case, the electrons are not produced by decay, but formed in extra-nuclear
readjustments that follow radioactive decay by orbital electron capture. Some
properties of the above-mentioned electron sources are listed in Table 11.1.
As a non-radioactive electron source for the ECD, a thermionic emitter 33
with a barium zirconate cathode protected by a guard gas has been decribed
(Fig. 11.3). A directly heated thermionic cathode supplies electrons, which
are attracted towards a mesh-like anode. The electrons diffuse through the anode
and are attracted with a small potential towards a collector. The column effluent
flows in the outer cylinder and the guard gas in the inner cylinder. The guard
gas prevents excessive penetration of the solvent into the filament chamber.
The reaction chamber proper consi sts of the .annul us between both cyl i nders. Thi s
new mode of operation is based on the phenomenon of space charge amplification.
The detection limit for lindane is 3.2.10- 16 g.
An ECD based on photoionization was described by Wentworth et al. 31. The
lamp was a laboratory-made device exploiting the Lyman a-resonance line of
hydrogen (10.2 eV) and was provided with a lithium fluoride window. In the
photoelectron-capture detector, UV lamps cannot bring about the ionization of

EXHAUST

HEATER WE

L..---+--FILAMENT
K---f---CIJLLECTOR
GRID
'-~--STRUCTURE

FLOW
"t---·STRAIGHTENER

Fig. 11.3. Schematic diagram of a non-radiative (thermionic) electron-capture


detector. (From ref. 33.)
223

the carrier gas like that with an ECD having a radioactive source, because the
ionization potential (IP) of these gases exceeds the energy from the lamp. For
this reason, a compound with an IP lower than the photon energy and at higher
concentration than any electron-capturing species should be added to the carrier
gas ahead of the detector. After the production of primary electrons

( 11.1)

the remaining electron attachment and neutralization reaction are similar to


those in the radioactive ECD (section 11.4). In the d.c. mode with triethyl-
amine (IP 7.50 eV), the detection limit for carbon tetrachloride is 50 pg
(10-150 times less than that with a tritium ECD in the pulsed mode). Fig. 11.4
shows a schematic diagram of a detector 34 using a commercial UV lamp (HNU
Systems) as the ionization source. The detector cavity contains five openings,
two of which serve as the electrode ports, two as the inlets and the last as
the outlet. When operating in the ECD mode (the detector can also work as a
photoionization detector (PID); cf., section 6.1), the column effluent is
brought to the bottom of the detector through inlet 1; easily ionizable sub-
stances (naphthalene or tri-n-propylamine) are introduced with a nitrogen stream
through the top inlet in close proximity to the magnesium fluoride window of
the UV lamp. The operation in the PID mode requires a reversal of the inlets.
The bottom electrode was polarized with a d.c. power supply (+20 V) and acted

-~~
exit

tL--_ _ inlet 1

BN Detector Cell
Top View

·,: :· · Linlet2
o·fI'~"""'''''~
~F====
Top electrode

D ~,,::::::::::..:
Linlet1
0---
Bottom electrode

Stainless-Steel Cell Housing

Fig. 11.4. Schematic diagram of combined photoionization-electron capture


detector. (From ref. 34.) BN = boron nitride.
224

as the anode. The detection limit is 1 pg for lindane and the linear dynamic
range is similar to that of the ECD in the d.c. or constant-frequency pulse
modes.

11.4. METHODS OF MEASURING DETECTOR CURRENT

The electron concentration in the ECD can be measured continuously by applying


a d.c. voltage or intermittently by applying pulses of short width, long period
and sufficient amplitude to collect all the electrons available.

11.4.1. Direat-aurrent mode

In the d.c. mode 35 , a constant potential is applied to the detector electrodes


and the detector is exposed to this potential throughout the operation. The
detector current increases with increasing potential applied to the ECD. At a
certain voltage, the saturation plateau is attained at which all the electrons
produced are collected (Fig. 11.5). The presence of an electron-capturing
compound in the detector reduces the concentration of free electrons, thus
inducing a drop of the standing current at all potentials until the field
becomes strong enough to collect both the electrons and negative ions simuta-
neously, and no effect due to electron capture is noted subsequently. The
voltage range for the ECD operation extends up to the onset of the saturation
plateau region. The optimum applied voltage occurs at the knee of the current-
voltage curve, approximately at 85% of the detector saturation current 6 ,36,37.
The exact optimum-voltage value is influenced by a number of factors, e.g., the
flow-rate of the make-up gas 38 or the pulse width 39 in the pulse method. The

I-
Z
LU
0::
0::
:::>
u
z
Q

APPLIED POTENTIAL

Fig. 11.5. Relationship between current and applied potential in a d.c. ECD.
A, Pure carrier gas; B, carrier gas containing a trace of a strongly electron-
capturing compound. (From ref. 14.)
225

maximur,l sensitivity of the ECD in the d.c. mode can be observed at low applied
voltages. However, owing to space-charge effects, contact-potential effects and
non-electron-capture ionization processes 12 , the detector may behave anomalously
in this region. Space charge is produced owing to differences in the mobilities
of positive ions and electrons in the applied field 12 ,40. A slowly moving
positive-ion drift to the cathode generates a cloud of positive ions in the
vicinity of the cathode, the potential of this cloud being inverse to the
potential applied. Changes in electron concentration occur. The secondary
electrons produced by the collisions of the B-particles with the carrier gas
molecules do not have enough time to attain thermal equilibrium and are rapidly
collected at the anode. This reduces their life times in the detector and makes
ther,l inaccessible to reaction with the solute. The negative molecular ions
formed by electron capture may also be collected at the anode, thus producing
an erroneous value for the detector current. Detectors with asymmetric geometry
are less influenced by space-charge phenomena.
The eluted solute can be adsorbed on the electrode surface, which can result
in the generation of a contact potential that is either complementary or inverse
to the potential applied 12 ,40. At a value of several volts, the contact poten-
tial may cause errors in the electron attachement process. If the potential is
inverse, the chromatographic peak area is anomalously large and this peak often
tails. If the contact potential is complementary, the response decreases and
a negative deflection at the tailing edge of the peak occurs. A solute that
generates a potential on the electrode surface, even if it does not absorb
electrons itself, may also create false responses. The use of high detector
temperatures and higher voltages (the contact potential being low in comparison
with the applied voltage) reduces the problems resulting from the generation
of contact potentials.

11.4.2. Putse mode with eonstant frequeney

I~any of the prob 1er,lS encountered wi th the d. c. ECD can be overcome by us i ng

the pulse mode 41 ,42. When a voltage pulse is applied, the electron concentra-
tion drops to zero owing to the collection of all electrons at the anode (Fig.
11.6). After each pulse, the electron concentration is restored, attaining a
constant value as a result of the irradiation of the gas. Hence the detector
is voltage free for a larger proportion of the working period, leaving enough
time for the concentration of the thermal electrons to be replenished by the
ionizing B-radiation and to attain thermal equilibrium. The amplitude and width
of the pulses should be of adequate size to provide for the complete collection
226

I-
Z
UJ
a:
Q::
;:)
U
a:
a
I-
u
UJ
I-
UJ
a

Vl
~
a
>

! !

o 100 200

TIME (MICROSECONDS)

Fig. 11.6. Effect of a pulsed voltage supply on electron concentration in an


ECD. w = pulse width in ~sec. (From ref. 3.)

of electrons (Fig. 11.7A and B), i.e., for the withdrawal of the standing cur-
rent from the detector. However, they should not be too large to cancel the
advantages of field-free operation. The relationship between the electron
concentration and the pulse period, t p' is shown in Fig. 11.7C. The optimum tp
value exceeds 1000 ~sec. In view of the low detector currents, lower values
are frequently preferred in practice. As a rule, the sensitivity of the detec-
tor increases with increasing intervals between the pulses, because enough time
is allowed for the recombination of positive and negative ions. A limit is set
by the increase in the recombination of positive ions and electrons 43 . During
voltage-free time periods, the electrons do not drift out of the plasma.
Negative ions are generated in the region where positive ions are also present
simultaneously, and the recombination of these ions is more effective for this
reason. The duration of the brief pulse is insufficient for collecting the
227

6 A

0 OS 10 1.5 2.0 2.5


PULSE WIDTH ( tw) IJ SEC
6
B
Vl 5
I-
Z 4
:J
>-
a: 3
«
a: 2
I-
m
a:
«IC])
z 0 10 20 30 40 50 60 70 80 90 100
PULSE AMPLITUDE ( VA) VOLTS
100

80
C

60

40

o 1000 2000 3000


PULSE PERIOD (t p ) IJ SEC

Fig. 11.7. Dependence of electron concentration (N e ) on (A) pulse width, (B)


pulse amplitude and (C) pulse period for a pulse sampled electron-capture de-
tector. (Reprinted with permission from ref. 15.)

negative molecular ions. In comparison with the d.c. mode, the sensitivity is
higher with the constant-frequency pulse mode, and the noise is approximately
the same.

11.4.3. FUZee mode with aonetant aurrent

At present, the most frequently used ECD mode is the constant-current


ECD 11 ,44, i.e., by changing the pulse frequency the detector current is kept
constant during the run. The base pulse frequency, fo' is low in the presence
of carrier gas alone. When an electronegative species enters the ECD, some of
the electrons are removed by the electron-capture processes to form negative
228

ECD CELL
NEGATIVE
RADiO-liONIZED
ACTIVE GAS
I L..J VOLTAGEPULSE VARIABLE f--
FOIL I FREQUENCY
PULSER

ELECTRON
COLLECTOR
II -I D
SIGNAL
0 UT
I
(I s - IDl ELECTRO-
METER
Is

REFERENCE
CURRENT

Fig. 11.8. Schematic diagram of the electronic components of a constant-current


pulse electron-capture detector. (From ref. 22.)

ions. The current drop is matched by an increase in frequency in order to


keep the cell current constant. The output signal is represented by the
frequency difference fA - fa, where fA is the frequency corresponding to the
sample concentration A. The change in pulse frequency is a measure of the
concentration of the electron-capturing solute passing through the detector.
A schematic diagram of the pulse-modulated constant-current ECO is given in
Fig. 11.8. The circuit forms a closed-loop electronic feedback network in which
an external reference current, Is' is compared with the cell current, 1 0 , so
as to maintain the relation 10 - Is = O. The base frequency, fa, giving the
optimum limit of detection for a particular solute in this mode is identical
with the frequency of operation that gives the optimum limit in the constant-
frequency mode. The optimum fa varies from solute to solute depending on its
electron-capturing ability in both pulse mOdes 23 . The main advantage of this
mode is claimed to be a larger range of linearity of the dependence of the
function fA - fa on concentration, A. The difference in the principles of the
two pulse modes is obvious from Fig. 11.9. The vertical line AS represents
the change in detector current with changing solute concentration at a constant
pulse frequency. The horizontal line CO represents the change in applied pulse
frequency with changing sample concentration at a constant detector current.
229

d.t.

__~~===="--~1~~
10.
0(

-...
5!

!z
i
it 1.0
PURE H2

FREQ=
(PI..lSE WlDTHrl

0.1 1.0 10. 100. 1000.


FREQIENCY (11Hz)

Fig. 11.9. Current versus frequency operating curves illustrating operation in


a constant-current pulse mode and in a d.c. mode (or constant-frequency pulse
mode). (From ref. 22.)

11.4.4. Other modes

By changing the applied voltage, constant-current operation is also possible


in the d.c. mode 45 ,46. The current drop due to electron capture is set to the
initial value by increasing the voltage. Fig. 10.10 shows a schematic diagram
of a d.c. constant-current system. The output current of the ECD is propor-
tionally converted to voltage and the latter is subtracted from a reference
voltage. The voltage difference is then amplified and fed back to the ECD.
The monitored detector output represents the variable polarizing voltage that
is applied to the ECD in order to maintain the current at a set level. Also
in this mode the linear dynamic range surpasses that of the d.c. mode with
constant voltage or that of the constant-frequency pulse mode and covers about
four orders of magnitude 46 .
An ECD with an a.c. input has also been described 47 , its sensitivity being
similar to that of the d.c. and pulse modes.
2~

ECD

RECORDER

Fig. 11.10. Schematic diagram of d.c. constant-current system. I/V = I/V con-
verter, KEPCO = KEPCO OPS operational power, AMP = amplifier. (From ref. 46.)

11.5. RESPONSE THEORY

11.5.1. Recombination theory

The standing current in the electron-capture detector arises from the produc-
tion of secondary electrons through non-elastic and elastic collisions between
primary electrons (a-particles) and molecules of the carrier gas (nitrogen,
argon or helium). A plasma of positive ions (p+), radicals (R) and thermal
electrons homogeneous through most of the detector cell is generated:

B + P + p+ + e- + a* + energy (11.2)

A direct process of ionization of the carrier gas is most probable with nitrogen.
Metastable atoms 3 may be formed with argon:

a + Ar + Ar+ + e- + a* + energy (11.3)

(11.4)

Ar* + X + Ar + X (11.5)
231

8* represents 8-particles with reduced energy as a result of thermal electron


production and X is a polyatomic quencher (usually methane).
The 8-particles lose their energies during their collisions with argon and
the quench gas until their energy becomes lower than that necessary for the
generation of ion pairs. Each 8-particle may generate 10 2-10 3 thermal electrons
before its kinetic energy is reduced to the thermal level, as about 30 eV are
expended for the generation of an electron-positive ion pair. The rate of produc-
tion of thermal electrons is assumed to be constant, being neither increased
nor decreased by the presence of capturing species 43 . Pure argon or helium is
not suitable for the attachn~nt of electrons to solute molecules, because the
molecules of these gases are readily converted into metastable forms that would
produce considerable ionization of the solute molecules 48 . p+ represents any of
the positive ions in the plasma, e.g., Ar+, ArH+, ArCH+, ArCH;, ArCH;, ArCH:,
CH:, CH;, CH; and R is any radical, e.g., H·, CH 3, :CH 243 ,49 The thermalization
(cooling) of electrons coming from high-energy 8-particles is essential in order
to allow or enhance the capturing process while minimizing solute ionization.
The thermalization of fast electrons is brought about by a polyatomic gas in
5-10% concentrations. Hence, the addition of a quenching gas serves two purposes:
(1) it reduces and maintains the electron energy at a constant thermal level
and (2) it removes, by deactivating cOllisions 43 , metastable argon species as
fast as they are formed. Each electron with an energy of 10 keV is cooled to
10% above the thermal energies (2.10- 2-5.10- 2 eV) in 0.076 ~sec43. Under these
conditions, the detector can work neither as an argon ionization detector nor
as an electron mobility detector.
If an electron-capturing solute enters the ECD, the thermal electrons are
captured giving negative molecular ions (non-dissociative reaction 11.6) or
fragment ions (dissociative mode, eqn. 11.7). This can be described by the
following set of reactions 43 ,49-52

e- + AB -+ AB ( 11.6)

(11.7)

( 11.8)

AB- -+ A- + B· (11.9)

e- T PT -+ neutra 1s (11.10)

PT T AB - -+ neutrals ( 11.11)

( 11.12)
232

Fig. 11.11. Potential energy diagrams for four electron-capture mechanisms.


EA = Electron affinity, E* = activation energy, ~E = overall change in the
internal energy for the process. (Reprinted from ref. 50 with permission.)

The attachment of an electron to a solute molecule is related to the electron


affinity and to the requirement for sufficient energy to cause attachment at a
given temperature. The larger the activation energy necessary for attachment,
the slower is the attachment reaction. The activated complex AB* represents an
intermediate form during which the electron is being accommodated by the solute
molecule. In order to produce this activated complex, AB has to absorb an
energy Ea'
Fig. 11.11 shows potential energy diagrams for four capture mechanisms 50
Mechanism I depicts a non-dissociative electron attachment. The overall change
in the internal energy for the process, ~, equals the electron affinity.
Aromatic hydrocarbons and carbonyl moieties serve as examples of compounds
reacting in this way. Mechanism.II depicts a single bimolecular electron-
attachment step, leading immediately to dissociation via a dissociative potential
233

energy curve (alkyl halogens except C_F)ll. In mechanism III, Ea is greater


than 6£ (aromatic halogens Cl, Br, I). This is a two-step dissociative process
via a dissociative potential energy curve. The two-step dissociative process
that first involves the formation of a molecular negative ion followed by
dissociation by means of the same potential energy curve represents mechanism
IV. Ea equals 6£. Mechanisms I and III involve a negative ion curve with a
large dissociation energy, but with mechanism III this curve is crossed by a
dissociative curve. In mechanism II the negative ion curve has a small or zero
dissociation energy, while the dissociation energy of the negative ion is
thermally accessible for mechanism IV. The kinetics of the electron-capturing
process is more complex than described here in a simplified way; detailed
studies can be found elsewhere 23 ,43,49-58.
The negative ions formed have a higher mass than the original electrons.
Hence, they display a lower drift velocity and a substantially higher rate of
recombination with positive ions 42 ,43. The conditions for operating the detector
are optimized by arranging that thermal electrons rather than negative ions be
collected. The detector response is then given by the background current drop
due to the loss of thermal electrons by attachment to solute molecules.

11.5.2. Positive space-charge modeZ

Wentworth's classical mode1 43 ,49,50 assumes the [CD to be a well mixed


reactor, where a single concentration expression describes the presence of each
species throughout the cell. The model described by Grimsrud et al. 59 for the
pulse mode differs from the previous model in that it considers the electro-
static forces between the charged particles as the dominant force 60 in deter-
mining the concentrations and locations of the charged particles within the [CD.
Thermal electrons are not evenly distributed throughout the [CD volume at all
tir,~s, but are concentrated in a local zone where charge neutrality exists

(plasma). The size of this plasma increases with time after the end of a pulse.
The plasma is separated form the cell boundaries by positive ions, forming a
sheath. This sheath decreases in size with time after the end of a pulse. All
thermal electrons are removed from the cell by the anode during the application
of each pulse, which results in a momentary excess of positive ions. The posi~
tive charge created in the cell by electron removal tends to dissipate itself
by space-charge-driven migration to all grounded surfaces of the cell during
the periods between pulses. A fraction of these positive ions strikes the anode
and causes a reduction in the time-average negative current, Ie' indicated by
the electrometer 59 ,61. Hence the [CD current measured need not necessarily be
234

dttributed to the collection of electrons alone. The observed current I is


given by I = (1-o)Ie , where 0 is the fraction of the excess positive ions that
migrate to the anode 61 • The magnitude of 0 depends on the cell geometry: it is
0.25 for the pin-cup detector61 and 0.01 for the displaced coaxial geometry62.

11.5.3. Negative speae-ahapge theopy

In the classical theory, the neutralization of electrons (response generation)


occurs via the intermediary negative ions being neutralized. In an alternative
(and/or complementary) theory63 of the response in the d.c. mode, the recombina-
tion of electrons and positive ions in the ionization region is increased owing
to the migration of negative ions outside this region.
The centre of ionization is very close to the foil (as close as 1 mm63 ,64
for 63Ni and 0.2 mm for 3H, depending on the detector shape). Electrons (and
negative ions) migrate over a relatively long distance from this region, setting
up an opposing or counter field. Owing to a space charge, the field gradient in
the ionization region decreases and, consequently, the electron-positive ion
reco,mbi nation increases. As electrons mi grate through all the vol ume, negati ve
ions can be formed in regions where no .positive ions are available for neutrali-
zation. Neutralization of these negative ions can occur only by contact with
the counter electrode or with any other conducting surface63 • Maximization of
the counter field by having the centre of ionization .situated as close as
possible to the cathode and as far away as possible from the anode (either by
moving the anode farther away from the radioactive foil or by increasing the
pressure in the detector cell) aids in maximizing the response. The response
increases with increasing electrode distance 65 ,66 (the pulse response follows
approximately the same trend 66 as the d.c. response; see Fig. 11.12) and also
with increasing pressure 65 ,67. The extent of electron capture, however, remains
constant66 • The same applies for the increasing pulse interval in the pulse'
mode 66 .
According to this mechanism, such an ECD should be able to function, even
if it were not possible for the negative ions to be neutralized by positive
ions. Such a situation may occur if the anions are generated far away from the
cations and if the two species are kept apart 68 (Fig. 11.13). In the "separated"
mode, the column effluent enters and leaves the anode chamber.
The cathode chamber is flushed with pure nitrogen only. Hence, the cations
should be located only in the cathode chamber and the anions only in the anode
chamber. Slow anions set up a sizable counter field. The cations and electrons
present in the ionization zone of the cathode chamber are slowed down. The
235

'A_ _ ~_l'00
'I. OF MAXIMUM CURRENT ",,"'''''

"'" .....MU" RE5"'",:/// " ~


ty/
/ 20
d. c. RESPONSE

PULSE RESPONSE
/'-0'
~ .......
/
I
!
Ii
i
o
o 4 8 12
ELECTRODE OISTANCE (mm)

Fig. 11.12. Variation of response with inter-electrode distance for d.c. and
pulse conditions. Scandium tritide. (From ref. 65).

second-order recombination rate increases and less electrons (and cations) reach
the electrodes: the d.c. system produces current drops typical of the ECD 68 ,69;
the same detector operated in the pulse mode 70 also produces a response.

11.6. RESPONSE

11.6.1. Effect of compound structure

The ECD response depends on the substance-specific term K, the electron-


capture (also electron-absorption) coefficient. This represents the degree to
which the compound is able to capture thermal electrons. The probability of
electron capture by different types of molecules spreads over a range of 10 6
depending on the presence of so-called electrophores (some atoms, groups and
structures) in the molecule 3 ,42,53,71,72 (see Table 11.2).
ptwires

ANODE CHAMBER

steel
....
nickel
cathode

purgegas
plus ..,
column effluent
in'conventional mode'

63 N i foil

borosilicate
space. reducer

t
col umn effluent
in 'separated mOde'

Fig. 11.13. Schematic diagram of the ECD with separated ionization and capture regions. (From ref. 68.)
237

TABLE 11. 2

RELATI VE ATTACHI4ENT COEFFICI ENTS K' FOR VARIOUS COMPOUNDS 41 ,42,73-75

From ref. 3.

Chemical class K'* Selected example


scale

Alkanes, alkenes, alkynes, Hexane

1
aliphatic ethers, esters 0.01 Benzene
and dienes Cholesterol
Benzyl alcohol
Naphthalene
0.10
Aliphatic alcohols, ketones,
aldehydes, amones, nitriles,
monofluoro and monochloro
compounds ~
]: Vinyl chloride
Ethyl acetoacetate
Chlorobenzene
1.0
Enols, oxalate esters, monobromo, cis-Stilbene
dichloro and hexafluoro
compounds
1
10.0
trans-Stilbene
Azobenzene
Acetophenone

Trichloro compounds, chloro- Allyl chloride


hydrates, acyl chlorides,
anhydrides, barbiturates,
tha 1i domi de and a1kyll eads } Benzaldehyde
Tetraethyllead
Benzyl chloride
Azulene
300
Cinnamaldehyde

J
Monoiodo, dibromo and trichloro
compounds, mononitro compounds, Nitrobenzene
lachrymators, cinnamaldehyde, Carbon disulphide
fungistatic compounds and 1,4-Androstadiene-3,11,17-triene
resticides Chloroform
1000

J
1,2-Diketones, fumarate esters, Dinitrobenzene
pyruvate esters, quinones, diiodo, Diiodobenzene
tribromo, polychloro, dinitro Dimethyl fumarate
compounds and organomercurials Tetrachloromethane
10 000

*Values for K' are relative to chlorobenzene, which is arbitrarily given a value
of 1.0.
238

The highest response is usually found with electronegative compounds con-


taining halogens or nitro groups, with organometallic compounds 76 - 78 and
compounds characterized by the presence of two or more weakly electron-capturing
groups connected by some specific bridge promoting a synergic interaction be-
tween the two groups79 (conjugated carbonyl compounds, some polycyclic aromatic
hydrocarbons and certain steroids).
The sensitivity of the ECD depends strongly on the structure of the com-
pounds 42 ,52. With halogen compounds the response decreases in the sequence 41
I>Br>Cl>F. The sensitivity of detection is affected by the position of the
electrophores and their number in the molecule of the compound 3 ,39,41,42,53,
72-75,80-84. The sensitivity increases synergistically with multiple substitu-
tion on the same carbon atom (Tables 11.3 and 11.4).

TABLE 11.3

RELATIONSHIP BETWEEN MOLECULAR STRUCTURE AND RELATIVE CAPTURE COEF-


FICIENT 41 ,42,73-75

From ref. 3.

Parameter K'*

Halogen series
I
Br
Cl
F
Substitution on carbon atom
Tertiary 10
Secondary 2
Primary 1
Frequency on carbon atom
Tetra-
Tri-
Di-
Mono-
Positional isomer (di-, tri-, etc.)
Alpha- 10
hb- 5
Delta- 1
Geometrical isomer
Trans- 4
Cis- 1

*Capture coefficients are relative to the lowest value of the series, which is
arbitrarily given a value of 1.0.
239

TABLE 11.4

EFFECT OF THE POSITION AND THE MULTIPLE SUBSTITUTION OF ELECTROPHORES ON


ELECTRON ABSORPTION

Reprinted with permission from ref. 42.

Electrophore Compound or class Absorption coefficient

Cl Vinylic 0.2
Aromatic 1
Aliphatic 0.3
Allylic 55
Benzylic 110
Benzene, 0- 42
Benzene, m- 30
Benzene, p- 11
-CHC1 2 1
Benzene, 1,2,3- 113
Benzene, 1,2,4- 75
Benzene, 1,3,5- 60
- CC1 3 500
Hexachlorobenzene 1100

The minimum detectability and detection limit of the ECD for compounds with
the highest electron-capture coefficients are the lowest of all gas chromato-
graphic detectors. Examples for some selected compounds are as follows: 3.10- 16
mole/sec for tert.-butyliodide83 , 1.10- 13 g for lindane and aldrin 22 , 1.10- 14
g/sec for chloropyrifos 85 , 1.10- 15 g for tetrachloromethane (capillary column)86,
3.7.10- 14 mole/sec for tetraethyllead 76 • 3.1.10- 15 mole/sec for benzophenone87 ,
2.2.10- 16 mole/sec for the heptafluorobutyramide derivative of S-phenylethyl-
amine 88 and 1.3.10- 16 mole/sec for chromium(III) trifluoroacetylacetonate89 .
Many other data can be found in Zlatkis and Poole's book on ECD 9.

11.6.2. Effect of detector temperature

The ECD response is highly temperature dependent. In principle, the charac-


ter of this dependence is given by the reaction mechanism of electron attach-
ment 43 ,49. Mechanism I gives a stable negutive molecular ion, the potential
energy curve of which (Fig. 11.11) l~es below that of the neutral molecule. On
electron attachment, this energy difference must be liberated either by
radiation or through collisions with other molecules by energy exchange. A
temperature increase in the detector would increase the population of higher
240

vibrational levels and, therefore, the probability of attachment would decrease.


In contrast, with mechanism II, representing dissociative attachment with the
simultaneous production of negative ions and radicals in a single step, a
temperature increase in the detector increases the probability of attachment due
to an increase in the population of excited levels. In this instance, the
potential energy curve would cross that of the neutral molecule at a level cor-
responding to the vibrational excited state. The activation energy of this
process would be the energy that is necessary to populate those states where
the dissociative surve crosses. Mechanisms III and IV also describe dissociative
attachment 90 •
The character of the detection mechanism can be inferred43 ,49,50 from the
plot of ln KT3/ 2 vs. l/T, where K is the capture coefficient and T is the
absolute detector temperature. With compounds that capture electrons in a non-
dissociative manner, the plot shows a positive slope at higher temperatures,
i.e., the response increases with decreasing detector temperature. On the other
hand, a negative slope of the plot characterizes the dissociative type of
electron capture; the response increases with increasing detector temperature
in this instance. Idealized plots related to the four basic types of electron
capture are shown in Fig. 11.14. The presence of a positive slope region for
mechanism IV is evidence of the formation of a negative molecular ion inter-
mediate. From a practical point of view, the mechanism of electron capture is
deduced from the plot of ln AT3/ 2 vs. l/T, where K is replaced by the peak area
A. The two plots have the same shape and their interpretations are similar.

N
M
I ;:; n
l- I-
:.: :.:

r
c c
..J ..J

1fT
"'---- lIT
N
M
m ;;; Ii
l- I-
:.: :.:
c c
..J ..J

V-
'" lIT II T

Fig. 11.14. Idealized plots of ln KT3/2 vs. l/T for the four basic mechanisms
of electron capture. (From ref. 91.)
241

Clearly, the nature of the electron-capture mechanism can also be assessed from
the dependence of the ECD response on the detector temperature. Generally,
aromatic compounds display a non-dissociative type of electron attachment,
whereas chlorine, bromine and iodine compounds show a dissociative type 15 ,43,44,
49,50,56. The constant-current mode can also be applied to establish the mecha-
nism of electron capture 92 •
It is evident from this analysis that the detector temperature substantially
affects the detection sensitivity of the ECD. Hence, it follows that (1) accurate
control of the detector temperature is necessary - it should vary within 44 ,SO
±0.3-0.1 0 C in order to obtain a 1% precision in the measurement of response;
(2) comparisons of the responses to various compounds or various derivatives
of the same compounds at the same temperature and/or without temperature indica-
tion (as well as quoting the detection limits or minimum detectability) are
misleading 93 ,94, as this gives no idea of the maximum possible detection sensi-
tivity, viz., the temperature dependences of the responses to different compounds
can follow entirely different courses; Table 11.S gives an idea of the extent
to which the sensitivity varies with detector temperature. This variation may
be as large as up to three orders of magnitude with a temperature difference
of 250 0 C93 , the temperature changes in the ~esponses to individual substances
are also outlined in section 11.6.3; (3) by optimizing the detector temperature
it is possible to increase the selectivity of the ECD response. Provided the
responses to individual compounds display different temperature dependences,
the responses to the compounds to be detected can be increased by increasing
the detector temperature, thereby suppressing the response to the other com-
pounds present.

TABLE 11. 5

LIMIT OF DETECTION FOR 1 ml OF SAMPLE

From ref. 93.

Compound Detection 1imit (ppb)


BOoC 227 0 C 3S0 0 C

0.01 0.01 0.01


1.0 0.10 O.OS
1000.0 40.0 B.O
1000.0 20.0 1.0
N
~
N

TABLE 11.6

RELATIVE RESPONSE OF THE ELECTRON-CAPTURE DETECTOR TO SOME HALOACYL DERIVATIVES

Reprinted with permission from ref. 96.

Derivative Compound
Amphetamine 100 Testosterone 101 Thymol 102 Diethylstilbesterol Benzylamine 105
Ref. 103 Ref. 104

Acetyl 1.0 1.0


Monofluoroacetyl 0.007
Monochloroacetyl 1.0 40 0.3 2.7 750
Chlorodifluoroacetyl 340
Dichloroacetyl 2.6
Trichloroacetyl 540 2.1
Tri fl uoroacety 1 0.1 4 1.7 1.5 200
Pentafluoropropionyl 40 50 1.3 15 5725
Heptafluorobutyryl 90 190 1.0 23 17875
Perfluorooctyl 230 600 21
Pentafluorobenzoyl 770 6.9
243

11.6.3. Derivatization for electron-capture detection

The electron-capture detector is the most sensitive detector in gas chroma-


tography as far as the compounds to which it responds are concerned. With
compounds having a low ECD response, efforts have been made to increase the
response so as to attain the highest ECD responses and lowest possible detec-
tion limits. Advantage is taken of the fact that the detector responds to certain
electrophores. Compounds with low responses are therefore converted into
electron-capturing derivatives whose responses are many times greater than those
of the parent compound 90 ,95,96.
Generally, the derivatizing agent contains a reactive group that provides
for bonding with the substrate, and an organic chain that provides for detector
sensitivity, i.e., an electrophore. Of course, this chain must meet certain
chromatographic criteria, i.e., it must render sufficient volatility, thermal
stability and chemical stability. The problems of derivatization proper, i.e.,
the methods of preparation and the choice of the reactive groups with regard
to the above chromatographic requirement, are broad and exceed the scope of
this section. Their detailed analysis can be found elsewhere 97 - 99 •
Naturally, derivatives with the highest ECD responses are most advantageous
for detection with an ECD. A comparison of the individual derivatives made
from this viewpoint is mostly based on a comparison of the responses of the
resulting products at the same detector temperatures. An example is given in
Table 11.6. As mentioned in section 11.6.2, such a comparison can be misleading
owing to different courses of the temperature dependence of the response.
Halogen, nitro, trialkylsilyl, haloacyl, pentafluorophenyl and boronic groups
are mostly introduced in derivatization for the ECD 96
The sensitivity of the ECD to chlorine, bromine and iodine compounds is
greater than that for fluorine compounds, but the larger molar masses of the
former result in less volatile derivatives, which can be undesirable in the
analysis of high-molecular-weight compounds. The presence of fluorine atoms
in alkyl and acyl compounds causes a slight increase in the boiling point
relative to hydrocarbons with the same carbon number. This makes it possible
for multiple fluorine substitution to be carried out, thus increasing the ECD
response (Table 11.7).
An increase in response by a factor as large as 10 4 can be attained 106 if
chlorine is replaced with iodine atoms through the reaction of the chlorine
compounds with sodium iodide directly ahead of the ECD.
Trimethylsilyl reagents are frequently applied derivatization agents in gas
chromatography, but the trimethylsilyl group shows no particular electron-
capture properties. By introducing a halogen atom (Cl, Br, I) into one of the
244

TABLE 11.7

RELATIVE VOLATILITY OF A SERIES OF RR 1(CH 3)Si-CHOLESTEROL ETHERS

Reprinted with permission from ref. 96.

R R1 Relative retention
time
CH 3 CH 3 1.00
CF 3(CH 2)2 CH 3 1.26
CF 3(CF 2)2(CH 2)2 CH 3 1.37
CH 2Cl CH 3 2.10
C6F5 CH 3 3.14
C6F5 CH(CH 3)2 4.57
CH 2Br CH 3 5.13
C6F5 CH 2Cl 6.26
C6F5 C(CH 3)3 6.30
CH 2I CH 3 12.82

12.0 , . . - - - - - - - - - - - - - ,

In AT¥2
11.5

11.0
c

10.5

10.0 L--'----.l._~.........__'_""'____..............J
1.5 1.6 1.7 1.B 1.9 2.0 2.1 2.2
3
1.10
T

Fig. 11.15. Temperature dependence of the ECO response to some octanol deriva-
tives. (a) Pentafluorophenyl dimethylsilyl ether; (b) pentafluorophenyl iso-
propylmethylsilyl ether, (c) tert.-butylpentafluorophenyl methylsilyl ether;
(d) chloromethylpentafluorophenyl methylsilyl ether. (From ref. 109.)
245

30

29

28

27
~
I-
:.: 26
c
...J

25

24

1.5 20
3
1fT .10

Fig. 11.16. Plots of ln KT3/2 vs. liT for the (1) pentafluoropropionate, (2)
heptafluorobutyrate, (3) chloroacetate, (4) chlorodifluoroacetate, (5) penta-
fluorobenzoate, (6) pentafluorophenacetate and (7) pentafluorophenoxyacetate
derivatives of n-hexanol. (From ref. 91.)

methyl groups or by rep 1ac i ng the methyl group wi th a pentafl uorophenyl group,
the detection sensitivity is increased107-113. In Fig. 11.15 the response levels
of the individual silyl derivatives of octanol and their temperature dependences
are shown 109 .
The haloacyl anhydrides rank among the most frequently studied reagents 91 ,
100-105, 114-116. The temperature dependence of a series of haloacyl derivatives
of n-hexanol is shown in Fig. 11.16; a comparison of the responses at non-
optimized temperatures of the detector is given in Table 11.6. Generally, the
Illonochloroacetyl and chlorodifluoroacetyl derivatives are more sensitive than
the trifluoroacetyl derivatives. An increase in the fluorocarbon chain length
of fluorocarbonacyl derivatives brings about an increa?e in their ECD response
without inconveniently increasing their retention times. In accordance with
the above properties, heptafluorobutyryl derivatives have found most universal
use.
A comparison of the properties of electron-capturing boronic acids as
derivatizing reagents at optimum detector temperatures is given in Table 11.8.
3,5-Bis(trifluoromethyl)benzeneboronic acid, 2,4-dichlorobenzeneboronic acid,
4-bromobenzeneboronic acid, and 4-iodobutaneboronic acid seem to be the best
reagents 117 ,118. The temperature dependences are given in Fig. 11.17.
TABLE 11.8

COMPARISON OF THE VOLATILITY AND ECD SENSITIVITY OF THE ELECTRON-CAPTURING BORONIC ACIDS

Reprinted with permission from ref. 96.

Boronic ester Relative Detection limit Optimal detector


vol atil ity* (x 10 -12 9 pinacol) temperature (oC)

3,5-Bis(trifluoromethyl)benzeneboronate 0.3 ± 0.05 3.0 180


Benzeneboronates 1.0 200.0 150
4-Iodobutaneboronates 1.8 ± 0.5
4-Bromobe~zeneboronates 3.9 ± 0.8 3.0 350
2,6-Dichlorobenzeneboronates 4.3 ± 2.0 18.0 380
2,4-Dichlorobenzeneboronates 4.7 ± 1.7 4.0 380
3,5-Dichlorobenzeneboronates 5.0 ± 1.1 11.0 380
2,4,6-Trichlorobenzeneboronates 6.9 ± 1.8 4.0 380
3-Nitrobenzeneboronates 11.7±3.4 4.0 300
Naphthaleneboronates 18.5 ± 4.6 2550.0 350

*Based on a comparison of the retention times for a series of bifunctional compounds compared with the benzeneboronate
derivative.
247

11.0

9.0'-------'-----'---
t50 2n0

Fig. 11.17. Plot of 1n AT3/2 versus liT for the (A) 4-bromobenzeneboronate. (B)
naphthaleneboronate, (C) 2,4-dichlorobenzeneboronate, (D) 3.5-dichlorobenzene-
boronate, (E) benzeneboronate. (F) 3-nitrobenzeneboronate derivatives of pinacol.
( From ref. 117.)

32

31

30
/i 1
5

• .2
_______4
29
);
.~:
I-
I<:
28
5
27

26
"'-----6
25
1.5 2.0
3
I/T·10

Fig. 11.18. Plots of 1n KrJ/2 vs. liT for some acetyl derivatives of phenol. For
the identification of the derivatives see Fig. 11.16. (From ref. 91.)
248

The response mechanism of the ECO is given not only by the properties of the
derivatizing agent - the same derivatives of different compounds can manifest
very different mechanisms (c.f., Figs. 11.16 and 11.18 for the pentafluoro-
propionate, heptafluorobutyrate and pentafluorophenacetate of n-hexanol and
phenol) and, consequently, would have different optimum temperatures. The plots
are similar 91 for similar types of compounds (e.g., n-hexanol, cyclohexanol or
n-hexylamine, cyclohexylamine). Therefore, the detection limits depend on both
the properties of the reagent and the structure of the parent compound and hence
are compound dependent.

11.6.4. Effect of impurities

It has been known from the early days of the ECO that it is extremely
sensitive to impurities. Impurities either contained in the carrier gas or
originting from column bleeding may contaminate the surface of the radioactive
source foil (see contact potential, section 11.4.1). A reduction in the standing
current 22 ,26,36,39,45,119 in the d.c. mode or an increase in the base frequency
in the pulse mode 22 ,120 occurs, followed by'a decrease in sensitivity. The
presence of the stationary phase molecules or their decomposition products
diminishes the probability of the capture reaction between the solute molecules
and electrons 36 . It is advisable, therefore, that the maximum column temperature
applied be lower than that with other detectors for conventional (non-cross-
linked) stationary phases. This phenomenon can be suppressed by increasing the
detector temperature 15 with some stationary phases, particularly non-polar ones.
However, this leads to changes in the response level due to the change in the
detector temperature, depending on whether the dissociative or non-dissociative
type of electron attachment is involved (see section 11.6.2). The simultaneous
elution of an electron-capturing and a non-capturing (alkane) solute can induce
changes in the detection sensitivity in both the d.c. and the pulse modes, the
extent of the changes depending on the detector shape and the carrier gas
employed 121 (see Table 11.9).
The presence of oxygen and water molecules has an appreciable effect on the
magnitude of the standing current and, consequently, on the sensitivity of the
EC0 12 ,122,123. Oxygen reduces the standing current due to electron capture,
giving rise to clustered species 124 (H20)n02' This effect is greater at lower
temperatures and is related to the water concentration. For this reason,
thorough removal of oxygen and water from the carrier gas is recommended with
the ECO. On the other hand, oxygen can increase the ECO response (see section
11.8.2) .
24')

TABLE 11.9

INTERFERENCE EFFECT (%) OF ALDRIN

Reprinted with permission from ref. 121.

Carrier gas Concentration Constant voltage Pul sed voltage


of i nterferi ng
n-C 20 alkane Concentric Parallel Concentric Parallel
tube plate tube plate

Nitrogen 0.01
0.1 + 3 - 15
1.0 - 13 - 2 - 19
10.0 + 8 + 36 + 3 + 14
Argon-methane(95:5) 0.01
0.1 + 11 - 7 - 5
1.0 + 15 + 43 - 17 - 18
10.0 + 39 + 65 + 24 + 36

11.7. LINEARITY OF RESPONSE

In the conventional operation of the ECD, Ib - Ie is measured, where Ib is the


detector current in the absence of electron-capturing species and Ie is the
detector current in the presence of electron-capturing species. Except for low
solute concentrations over a range of about two orders of magnitude, the above
difference is not linearly proportional to the solute concentration 125 ,126
Under certain conditions, the following equation for the detector response is
valid in the pulse sampled mode 50 :

= :'a ( 11.13)

where:. is the electron-capture coefficient and a is the concentration of the


species. In order to satisfy eqn. 11.13, the detection system can be made to
give an output linear with a over a range of 1.10 5 by the use of a reaponse
converter 125 ,126. The pulse periods for which this equation is valid are
10ng 126 ,127, i.e., 1000-2000 ~sec, which implies a need for extremely stringent
conditions of cleanliness to ensure an adequate electron concentration in the
detector for satisfactory operation. Hence the disadvantage of the ECD in the
d.c. and constant-frequency modes is a narrow range of the linear dependence
of the response on the solute concentration.
250

In the constant-current mode of the ECO, the measured term is the change in
pulse frequency44, fA - fO' where fO is the applied pulse frequency required
to keep the current constant when the electron-capturing species are not present
in the detector. In this instance, the linear dynamic range of the response is
much larger, amounting 22 ,44 to about 5.10 4 • The dynamic range 22 is about 7.10 5
The linear dynamic range depends on the flow-rate of the carrier gas 22 , an
additional stream of 70 ml/min of make-up gas shifts the linear range towards
higher sample weights, which represents the behaviour of a concentration-type
detector.
The problem concerning the relationship between the linear dynamic range of
the constant-current ECO and the type of carrier gas used seems to depend on
the geometry of the detector. The linear dynamic range is generally reported
to be smaller 22 ,86,128 in nitrogen than argon plus 5% of methane. However, with
the detector described by Patterson 22 , the linear dynamic ranges are commensu-
rate with both gases. The linear dynamic range is slightly decreased by
decreasing the gas flow-rate or increasing the standing current 55 with the
constant-current mode, and by changing the detector temperature or decreasing
the pulse period 126 with the constant-frequency mode.
However, the dependence of the response on the solute concentration of a
constant-current ECO can also be non_linear 23 ,55,56,129-131 for some of the
compounds for which the ECO is most sensitive, i.e., for highly capturing com-
pounds such as tetrachloromethane, trichlorofluoromethane, aldrin, lindane,
dieldrin and mirex. The course of this dependence is linear in the region of
low solute concentrations the sensitivity increasing at higher concentrations.
This course is frequently S-shaped 129 , with a linear section again in the region
of the highest concentrations (Fig. 11.19). Alkyl chlorides show 132 a non-linear
response dependence in the low concentration region, whereas the response is
linear at higher concentrations. For highly halogenated compounds, this course
has been explained as follows 130 ,131: at low solute concentrations, a significant
fraction of the solute is destroyed by the electron-capture reaction, because
the electron-capture rate constant and the average electron density are both
large. As long as the electron population remains high and relatively constant,
causing the same fraction of solute molecules to react over this limited con-
centration range, the response is linear. However, as the solute concentration
increases, a smaller population of electrons consumes a successively smaller
fraction of the solute entering the cell. Also, as the response is directly
proportional to the instantaneous concentration of the solute within the detector,
detector, the response becomes non-linear. It is possible to minimize the
problem by employing 22 ,133 a small effective detector volume (e.g.,22 in which
the response to lindane, aldrin and mirex is linear over a range of 5.5 orders
251

RESPONSE FACTOR mV/pg

0.3
ALDRIN

0.2

0.1

10 100 1000 10 000

PICOGRAMS

Fig. 11.19. Calibration graph for aldrin. Constant-current ECD. (From ref. 129.)

of magnitude. Bipolar pulsed microvolume detector extends the linear dynamic


range to higher concentrations 134 . Knighton and Grimsrud 131 suggested the func-
tion (f-fO)(H+f)/F instead of the common function f-fO in order to improve the
dependence of the linearity of the response on the solute concentration,
especially using high baseline pulse frequencies (ca. 2000 Hz)135. However, it
could be inferred from Connor's theoretical conclusions 23 that operation in the
constant frequency mode rather than in the constant current mode should give
optimum linearity of response.
A linear dynamic range covering four orders of magnitude can also be attained
in the constant-current d.c. mode 45 ,46 and the advantage of the latter over the
constant-current pulse mode is the fact that no deviation from linearity has
been found for strong electron capturers 46 .

11.8. SELECTIVE ELECTRON-CAPTURE SENSITIZATION

The sensitivity of the ECD is strongly influenced by many factors. Impurities


in the carrier gas usually reduce the detection sensitivity when using the ECD.
For a long time, it had beer. required for the ECD that the carrier gas be inert
to the electron-capture process. Later it was found, however, that the presence
of certain impurities could increase the detector response to certain compounds.
This effect has been proved for nitrous oxide and oxygen in the last few years.
252

11.8.1. Nitrous oxide doping of the carrier gas

The molecule of nitrous oxide is electron attracting and nitrous oxide itself
responds in the ECD136-13S. If nitrous oxide is used as a dopant in the carrier
gas, the following reactions take place in the ECD 139 :

e- + N20 + 0- + N2 (11.14)

0- + N0 + NO- + NO (11.15)
2
NO- + N2 + NO + N2 + e- (11.16)

The concentrations of electrons, 0- and NO depend on the rate constants in


eqns. 11.14-11.16, the concentration of nitrous oxide in the carrier gas and
the operating temperature of the detector. Any solute that can react with 0 or
NO-, giving a stable negative ion, interrupts the reaction cycle, causing a
reduction in the electron density and a concomitant detector response:

0- + A + stable negative ion ( 11.17)

The result of summing the steady-state ion chemistry outlined in the above
equations represents the electron-catalysed conversion of nitrous oxide to
nitrogen and nitrogen oxide:

( 11.1S)

The formation of nitrogen oxide in the ECD could be demonstrated by comparing


the NO levels in a carrier gas containing 20 ppm of nitrous oxide in the inlet
and outlet of the detector 142 . Whereas the NO level in front of the ECD was
~ 0.5 ppb, the level behind the detector was 5.5 ppb.

Hence, compared with the common ECD, the presence of nitrous oxide causes
response sensitization (selective electron-capture sensitization, SECS) for
compounds that do not rapidly attack the electron, but that do react with 0
to form stable negative ions and, hence, also for compounds that do not directly
capture electrons (see Table 11.10). The detection limit for compounds that rap-
idly convert electrons to stable negative ions is slightly increased by addition
of nitrous oxide, owing to the increased baseline noise 140 • Compared with the
value at zero nitrous oxide concentration, with 20 ppm of nitrous oxide the
noise level is increased 141 by a factor of 4-5. The detection limits for N20-
sensitized ECDs are in the picogram range: 1.4 pg for vinyl chloride 141 , 16 pg
for carbon monoxide 142 and 30, 16, 15, 1S, 35 and 62 pg for methane, ethane,
propane, butane, pentane and hexane, respectively143. Carbon monoxide does not
253

TABLE 11.10

ELECTRON-CAPTURE DETECTOR RESPONSE ENHANCEMENT FACTORS WHEN NITROUS OXIDE IS


ADDED TO THE CARRIER GAS

Detector temperature, 350 0C. From ref. 140.

Compound Formula Enhancement factor*

Hydrogen H2 40
Carbon dioxide CO 2 10 000
Methane CH 4 13
Ethane C2H6 27
n-Propane C3H8 31
n-Butane C4H10 34
n-Pentane C5H12 29
n-Hexane C6H14 11
Dich10rodif1uoromethane CC1l2 0.23
Dich10rotetraf1uoroethane CC1F 2CC1F 2 0.29
Trich10rof1uoromethane CC13F 0.40
1,1,1-Trich10roethane CH 3CC1 3 0.51
Trich10rotrif1uoroethane CC1lCC1F2 0.33
Carbon tetrachloride CC1 4 0.43
Dich10romethane CH 2C1 2 0.32
Chloroform CHC1 3 0.40
Tetrach10roethene C2C1 4 0.45
Carbonyl sulphide COS
Methyl chloride CH 3C1
Ch1orodif1uoromethane CHC1F 2 63
Methyl f1 uori de CHl 782
Acetaldehyde C2H4O 11
Vinyl chloride C2H3C1 760
Carbon monoxide CO 100

* detection limit with 20 ppm N20 in N2 carrier


Enhancement factor = detection limit without N 0 in N2 carrier
2
N
<.n
.:>0

TABLE 11.11

SECS ENHANCEMENT FOR POLYNUCLEAR AROMATIC HYDROCARBONS

From ref. 141.

Compound Detector temperature 250 oC* Detector temperature 350°C


Signal S/N** Signal S/N**
enhancement enhancement enhancement enhancement

Anthracene 3.2 0.8 6.3 1.6


9-Methylanthracene 1.5 0.4 5.9 1.5
Phenanthrene 9.7 2.4 8.1 2.0
Tetracene 1.4 0.4 1.6 0.4
l,2-Benzanthracene 1.3 0.3 3.2 0.8
Chrysene 5.5 1.4 6.5 1.6
Triphenylene 9.5 2.4 17 4.3
Pyrene 6.8 1.7 11 2.8
Benzo [gJ pyrene 0.6 0.2 1.7 0.4
Benzo [iiJ pyrene 3.5 0.9 34 8.5
Perylene 3.2 0.8 6 1.5
Acenaphthene 4.0 1.0 4.0 1.0
Fluorene 1.2 0.3 1.5 0.4
Dibenzofuran 17 4.3 29 7.3
Dibenzothiophene 18 4.5 25 6.3
Carbazole 3.2 0.8 8.4 2.1
Acridine 3.7 0.9 14.4 3.6

*Selective electron-capture sensitization response at a detector temperature of 350 0 C divided by the ECD response at
**250 0 C.
SIN = signal-to-noise ratio.
255

:;;DIMETHYLPHENOL

E CD

Fig. 11.20. Identification of phenols in shale oil waste water (steam distil-
late); 2-~1 split injection, temperature programmed from 50 to 200 0 C at 20 C/min.
Top, selective electron-capture sensitization; bottom, without nitrous oXide.
Capillary fused-silica column, 25 m, Carbowax 20M. (From ref. 141.)

directly react with 0- to form a stable negative ion. For this compound, the
sensitivity is explained 142 by the catalytic conversion of carbon monoxide to
carbon dioxide in the presence of nitrous oxide on the hot detector wall. The
practical effect of adding nitrous oxide (or other dopant gases) to the carrier
gas consists in converting the relatively selective ECD into a much less selec-
tive detector with no change in instrumentation.
The increase in the response of certain compounds in a complex chromatogram
due to the addition of nitrous oxide to the carrier gas in the ECD can also
be utilized to identify unknown peaks 140 ,141,143 (Fig. 11.20). Even isomers
of compounds 141 can be distinguished in this way (c.f., Table 11.11, isomers
of toluidine, benzo[gJpyrene vs. benzo[eJpyrene). In addition to the compounds
listed in Table 11.11, an enhancement of response was also observed 143 for
benzene, ethanol, methyl isobutyl ketone and other groups of compounds 141 such
as phenols, amines, polynuclear aromatic hydrocarbons (PAHs), aromatic and
heteroaromatic compounds and water. A negative response was observed for benzene
and furan 143 and for some other PAHs and aromatic hydrocarbons 141 , frequently
as a W-shaped peak.
256

The dependence of the sensitivity and minimum detectability for vinyl chlo-
ride on the concentration of nitrous oxide in the carrier gas in shown in Fig.
11.21. The increase in sensitivity with increasing nitrous oxide concentration
up to about 20 ppm results in an increase in the minimum detectability. Both
parameters stop changing 144 at higher nitrous oxide concentrations. Fig. 11.22
shows the dependence of the response to chloroethane on the detector tempera-
ture. The sensitivity decreases with decreasing temperature so that the minimum
detectability at 190 0 C is increased by a factor of about ten with respect to
the value at the maximum operating temperature of 350 0 C'40. Similar relation-
ships are also valid for other compounds 139 ,140,142-144. The optimum concentra-
tion range is from 15 to 70 ppm of nitrous oxide in nitrogen; nitrous oxide
is introduced into the carrier gas between the column outlet and the ECD 140 ,144.

f] :~: 0 : 0: ; : J
.!!!
'c ,
~ , I
I

,,
t' I
~
ii

,,,
0...
o I
-;;1(J'"

,
\/I
c:
o
0.
\/I
Q)
a: I
I

,~L-__L -__L-~L-~L-~L-~L-~

o 10 20 30 40 50 60 70
Concentration of ~o (ppm,vlv)

Fig. 11.21. Response of the sensitized ECO to vinyl chloride and the signal-to-
noise ratio as a function of nitrous oxide concentration. (Reprinted with
permission from ref. 144.)
257

>
~20
QI
11\
C
o
Co
11\
QI
0::
....
o
'0
QI

Gi10
Cl

350 300 250 200 'C


O~--,-~~----~--~----~--~
W v ~ ~ ~ ~ W
10001T ( K-1)

Fig. 11.22. Response to vinyl chloride as a function of the detector temperature


with carrier gas nitrogen containing 17 ppm of nitrous oxide. (From ref. 140.)

The degree of sensitization is significantly affected by the presence of


impurities in the ECD, the source of which can be the carrier gas transfer
lines, controllers, valves and other equipment, column bleed and leaks 140 ,144.

11.8.2. Oxygen doping of the ea~~ie~ gas

Oxygen present in the ECD carrier gas considerably reduces 12 ,122,123 the
detector sensitivity and linearity by producing O2 ions, and in a wet carrier
gas also by molecular complexes 124 (see eqn. 11.20). This removes a significant
proportion of the electrons and prevents them from performing their function
of reacting with solute molecules. However, Van de Wiel and Tommassen 123 found,
using a constant-frequency ECD, that, compared with pure nitrogen, the electron-
capture coefficient of a weakly responding compound such as n-butyl bromide
increased and that its ECD response increased at high oxygen concentrations in
the carrier gas and at high temperatures. These findings were later extended
to other compounds: halogen compounds 132 ,145-149, polycyclic aromatic hydro-
carbons 150 ,152, polycyclic aromatic amines 152,153, polycyclic aromatic
258

154
hydroxides 152 , carbon dioxide 138 and bis(chloromethyl)ether • For all these
compounds, both the sensitivity and the linearity of response were improved by
oxygen doping.
The reaction scheme is as follows 145 ,147,149:

( 11.19)

or, if the carrier gas is not dry,

( 11.20)

The overall reaction is

O2 + sample molecule ~ negative + neutral species (11.21)

2
Reaction 11.21 consumes O and shifts the position of the equilibrium of
reaction 11.19 to the right. Similarly to nitrous oxide doping, the reaction of
compounds that have very low rates of electron capture with 0; can be faster
than that with the free electron. Then, for instance, highly chlorinated
molecules show little response enhancement, because their rate of direct electron
capture is much faster.
In the constant-current mode of the ECD, the baseline frequency increases 145
at oxygen concentrations above ca. 100 ppm with increasing oxygen concentration
and decreasing detector temperature. The response also increases with in-
creasing oxygen concentration 145 ,146,149. The increase begins mainly at
concentrations of about 1% (Fig. 11.23). With decreasing detector temperature
and an oxygen concentration at a ~oo level, the response for halogen compounds
increases 145 ,146 (Fig. 11.24), but it decreases for anthracene and 1-chloro-
butane 145 . In the constant-frequency mode, the course of the dependence of the
response increases on the oxygen content displays different features (Fig.
11.25). The maximum for methyl chloride is attained 155 at relatively low dopant
concentrations. The value of the response increase is about 4 with 0.030% of
dopant, but the response decreases at higher concentrations. At the same
detector temperature, the sensitivity increase is about 200 for the constant-
current mode. Hence the constant-current mode is more advantageous for oxygen
doping.
When oxygen is added to the detector, both the sensitivity and noise increase.
Like the response, the noise becomes the greater the lower is the temperature
of the detector 146 • Hence the signal-to-noise ratio for a given compound does
not increase on oxygen doping as much as the response does 122 . For instance,
for methyl chloride at a detector temperature of 250 0 C and a 0.2% oxygen con-
259

+
/ 300'C
60
CH 3 CI

GJ
~40
.1
1/
o
0-
Il)

,4
....GJ
GJ
.~
C
~20 + 6 /

~
/o/O
6/0 CHCI 3
1:>/0 _x
o x~-
o~ __ ~==b~~~~-DO.--~~--O­
CCI4
o 123 4 5
O2 concentration (%0)

Fig. 11.23. Effect of oxygen doping on the ECD response to halogenated methanes.
(From ref. 149.)

C~CI
/
o
300
2000cj
+

GJ
~200
o
0-
Il)
GJ
....
GJ
>

GJ
a:: 100

a 1 234
O2 concentration (0/00)

Fig. 11.24. Effect of detector temperature on the oxygen-induced response


enhancements of chloromethane. (Reprinted with permission from ref. 146.)
N
m
o
TABLE 11. 12

ECD RESPONSES AND RESPONSE ENHANCEMENTS CAUSED BY 2.0% OF OXYGEN IN THE CARRIER GAS

ECD responses under the normal conditions of clean carrier gas are the first values listed under each detector temperature.
These are relative molar responses normalized with respect to the case of methyl chloride at 300 oC. Oxygen-induced
response enhancement are listed in parentheses under each detector temperature.
Reprinted with permission from ref. 147.

Compound Detector temoerature Relative oxygen-


0 0 0
caused response, *RO
250 C 300 C 350 C 2

CH 3Cl 1.6( 189) 1.0(113) 1.0(56) 1.0 ± 0.3


CH 2C1 2 3.8( 108) 4.6(32) 8.3(10.9) 1.3 ± 0.4
CHC1 3 459(4.8) 662(1.71) 815(1.42) 4.1 ± 1.6
CC1 4 9100( 1.90) 10 500 (1. 20) 11400(1.15) 18 ± 9
CH 3CH 2Cl 2.1(228) 1.5(135) 1.4(57) 1.8 ± 0.6
C1CH 2CH 2Cl 4.6(161) 5.1(69) 5.2(22) 3.1±1.0
CH 3CH 2CH 2Cl 2.0(201) 1.3(127) 1.2(57) 1.4 ± 0.5
CH 3CHCl CH 3 2.0(195) 1.4(113) 1.2(50) 1.4 ± 0.5
C1CH 2CH 2CH 2Cl 2.2(180) 3.6(81) 5.1(34) 2.6 ± 0.9
CH 3CH 2CH 2CH 2Cl 2.3(145) 1.6 (84) 1.4(34) 1.2 ± 0.4
CH 3CHC1CH 2CH 3 3.5(132) 2.2(65) 2.3(25) 1.2 ± 0.4
(CH 3)3 CCl 1. 7(95) 1.3(47) 1.8( 13) 0.53 ± 0.20
C1CH 2CH 2CH 2CH 2Cl 3.4(197) 3.2(94) 2.9(53) 2.6 ± 1.0
CH 2=CHCl 0.0068(107) 0.0068(69) 0.013(29) 0.0041 ± 0.0014
CH 2=CC1 2 197(1.78) 373(1.13) 675( 1.02) 0.3 ± 0.3
tmns-Cl CH=CHC1 1.7(17) 3.7(3.9) 8.3(1.59) 0.09 ± 0.03
cis-Cl CH=CHC1 1.2(20 ) 2.3(5.5) 4.9(2.0) 0.09 ± 0.03
C1CH=CC1 2 505(2.9) 732(1.42) 1070( 1.11) 2.6 ± 1.5
C1CH=CHCH 3 0.0036( 161) 0.0031 (91) 0.0056(33) 0.0024 ± 0.0007
CH 2=CC1CH 3 0.180(3.0) o•190 ( 1.88) 0.21(1.11) 0.0015 ± 0.0006
CH 2=CHCH 2Cl 5.7(153) 4.9(61 ) 6.3(20) 2.6 ± 0.9
trans-C1CH 2CH=CHCH 2Cl 380(8.5) 500(2.9) 580( 1.69) 8.4 ± 3.0
C6H5CH 2Cl 42(44) 67(10.3) 97(4.5) 5.5 ± 2.3
C6H5Cl 0.029( 14.7) 0.068(3.7) 0.16( 1.84) 0.0017 ± 0.0006
o-Cl-C 6H4C1 20(3.9) 43(1.52) 77 (1.26) 0.18 ± 0.07
m-Cl-C 6H4Cl 29(10.2) 60(2.4) 103( 1.39) 0.75 ± 0.26
C1 2C=CC1 2 4000(1.79) 4880( 1.16) 6160(1.02) 7±5
p-Cl-C 6H4C1 10.6(8.8) 26(2.0) 47(1.25) 0.23 ± 0.09
CH 3Br 18(55) 24(13.3) 39(5.0) 2.6 ± 0.9
CF C1 6.9(2.7) 9.3(1.11) 12.8( 1.00) 0.008 ± 0.008
3
CHF 2Cl 2.0(190) 1.35(90) 0.67(62) 1.1 ±0.4
CF 2C1 2 174(3.4) 253(1.52) 361(1.10) 1.1 ± 0.5
CFC1 3 4450(2.2) 5210 (1. 31) 5850(1.03) 14 ± 7

*The relative oxygen-caused response is the contribution to an overall response provided by 0.2% of oxygen. These values
have been normalized with respect to methyl chloride using the ECD data at 300 0 C.
N
0'1
262

Oxygen Concent rat ion (0/00)

Fig. 11.25. Response enhancement as a function of the oxygen concentration in


the carrier gas for (a) constant-frequency ECO (®) and (b) constant-current
ECO (x). (Reprinted with permission from ref. 155;)

centration, the signal-to-noise ratio increases by about 16 times, whereas the


sensitivity increases by about 200 times. If' the response increase with oxygen
doping is very low, such as that with carbon tetrachloride, the signal-to-noise
ratio even decreases.
The number of chlorine atoms per molecule of solute significantly affects
both the response enhancement and the ECO response (see Table 11.12). For
instance, with chlorinated methanes from methyl chloride to carbon tetrachloride,
the ECO response increases by almost four orders of magnitude, while the cor-
responding response enhancement decreases from 113 to 1.20. The highest response
increase has generally been observed for halogen compounds that yield a low
reaponse in the ECO, e.g., the monohalogenated alkanes. An exception is the
relatively small response increase with some monohalogenated alkanes without
any hydrogen atom on the carbon to which the chlorine atom is attached 147 (tert.-
butyl chloride, 2-chloropropene, chlorotrifluoromethane). Isomers display dif-
ferent response enhancements 147 ,151-153,156 (cf., for instance, Table 11.3,
trans-, ais- and 1,1-dichloroethylenes, 0-, m- and p-dichlorobenzenes and
also 151 ,152 1-, 2- and 9-chloroanthracenes, 2- and 7-methylanthracenes, benzo-
pyrenes, aminobiphenyls, dihydroxynaphtalenes, etc.), which suggests an
instrumentally simple means for compound verification in analysis by gas
chromatography - electron-capture detection, particularly with the simulta-
neous use 157 of two ECOs, with oxygen added to one of them. For inxtance, the
response enhancement is 3.9, 5.5, and 1.13 for trans-, ais- and 1,1-dichloro-
ethylenes, respectively. The response enhancement is independent of the solute
concentration over a range of at least two orders of magnitude 156 ,157
263

No enhancement is generally observed with compounds that do not yield a


normal ECD response (e.g., aliphatic hydrocarbons). The method of adding oxygen
provides a nreans of enhancing certain responses that may already be faintly
evident in preference to the response of the strongly responding compounds which
normally dominate the chromatogram 149

11.8.3. Sensitization of aromatic hydrocarbons

The response of a pulsed ECD to some aromatic hydrocarbons is anomalous141 ,150.


The direction of this response is opposite to that of the normal ECD response.
The peaks are often "M-shaped". In pure carrier gas, the identity of the positive
ions changes as the aromatic hydrocarbon (e.g., anthracene, acridine) passes
through the detector, with a simultaneous increase in the population of both
positive ions and electronsI58-160. Increase in the electron population is due
to the smaller rate constant for the recombination of electrons with the
protonated aromatic positive ions compared with those for the positive ions
normally presented l58 . This anomalous effect can be eliminated by adding a
compound (e.g., methylamine, trimethylamine, diethylamine) with a gas-phase to
the carrier gasI58-160.
Doping of the carrier gas with 10 ppm of ethyl chloride (or isopropyl chlo-
ride) enhances the detector response to anthracene by ca. one order of magnitude.
Alkyl chlorides react with short-lived negative anthracene ions to form Cl and
neutral species 159 (reactions 11.22 and 11.23).

(11.22)

( 11.23)

The use of both alkyl chloride and alkylamine as dopants altogether produces
useful and well behaved responses 159

11.9. COULOMETRIC AND HYPERCOULOMETRIC RESPONSE

The calibration of the detector response in a range of very high sensitivies


with corresponding small amounts of compounds is a difficult problem. In order
tosolvethis problem, Lovelock 161 suggested that the ECD itself could be used as
an absolute calibrant. This means using a detector in the coulometric mode in
which, at 100% ionization efficiency, the solute concentration can be calculated
via Faraday's law directly from the number of the electrons absorbed. With
264

intensely electron-absorbing substances, the ECD tends to become a destructive


detector in which a large proportion pf the substance entering the detector is
ionized irreversibly.
Coulometry requires the equilibrium of reaction 11.5 to be shifted almost
entirely to the right or AB- to be rapidly scavenged from the system by reac-
tions 11.7, 11.8 and 11.10. The concentration of positive ions within the
detector is about 1000 times higher than the electron concentration, and this
would favour reaction 11.10. Therefore, under certain conditions, the ioniza-
tion of most of the molecules entering the ECD is favoured and, hence, a
cou10metric response 23 ,54-56,62,161-168 is obtained. The cou10metric method of
detecting electronegative compounds enables one to decrease the detection limit,
because at 100% ionization efficiency the ECD shows its maximum sensitivity.
If an ECD is to function successfully as a gas-phase cou10meter, the follow-
ing requirements should be envisaged 167 • (1) The concentration of electrons
must be in excess of the concentration of sample molecules (under typical
conditions of operation, the average electron concentration in the pulse mode
is in the range 10- 13 mole/lor 6.10 7 e1ectrons/m1). The minimum detectable
concentration of intensely electron-absorbing substances, e.g., sulphur hexa-
fluoride, tetrach10romethane and, halogenated pesticides, is about 10 6 mo1ecu1es/
m1. (2) The electron-capture reaction must be nearly complete. (3) The
stoichiometry of the reaction of thermal electrons with the sample must be 1:1
or some other known ratio. (4) No side reactions of the sample molecules may
occur. (5) The current monitored from the ECD must reflect the absolute amount
of electrons present in the entire ECD cell, and the ECD response must abso-
lutely reflect the loss of these electrons due to electron-capture reactions.
The ventilation of the detector by the flow of the carrier gas also removes
the electron-absorbing substances from the detector. Hence, the amount of the
substance ionized decreases 161 ,162 with increasing flow-rate of the carrier
gas, and the highest cou10metric response is observed at the lowest flow-rate
of the carrier gas.
At an ionization efficiency of less than 100%, the application of several
identical detectors in series makes it possible to determine the portion ionized
in either of them and, hance, the signal corresponding to complete ioniza-
tion 161 ,162. The cou10metric mode is difficult to attain with a common ECD.
The ionization of most of the emerging molecules is provided for by a detector
of special design with a long ionization chamber 55 replacing the arrangement
of several detectors in series. This detector operates in the cou10metric mode
for dibromodif1uoromethane at flow-rates of up to about 60 ml/min, whereas the
conventional detector approaches this mode giving a cou10metric response only
265

at very low flow-rates. In practice, the detector never yields a signal that
is truly coulometric; a small portion of the electron signal is always lost
by recombination 56 • This loss is about 10% in argon-methane or argon-hydrogen
carrier gas and at a pulse period of 250 ~sec. The loss is even higher in
nitrogen, about 30%. Even with the long coulometric detector the signal effi-
ciency attained is 90-97%.
According to the "classical" theory (section 11.5.1), the [CD in the
coulometric mode of operation should yield the maximum response, and the ratio
of the peak area in Faradays to the amount injected in moles should not exceed
unity. This ratio can be 2-3 if the products of solute ionization possess
greater electron affinities than the reactants 2. Aue and Kapila 169 , however,
found far greater ratios for many compounds with commercial d.c. mode [CDs at
elevated pressure. The ratio attained a value of 20 with a detection limit of
10 fg for hexachloroethane. This hypercoulometric response increases with
increasing pressure 63 ,67; the ratio reaches a value of 50 for tecnazene at
5 atm. This effect also exists in the pulse [CD63 ,170 (Fig. 11.26) and increases
with increasing pressure. The hypercoulometric response is explained by a
mechanism based on migrating negative ions 63 (section 11.5.3).
Coulometric detection forms the basis 55 ,171-174 for solute switching using
the [CD. Two [CDs coupled in series are used. The first detector, having a much
higher density of thermal electrons available to attach compounds, operates on
the coulometric principle and serves as a solute switch. This detector switches
between the destruction and the free passage of solute by switching the applied
potential on or off. The second [CD is locked into the switching frequency of

d.c.

~
8
E
.....
Q)
6
Q) Pulse
III .J;.- - - - .....t>- - - - - 'Zl.-
C
a. 4
0 ~

III
.... ~'"
Q)
0:: ,. "
2 l:(

Cell Pressure

Fig. 11.26. Hypercoulometric response vs. cell pressure for d.c. and pulse modes.
elm = Faradays (peak areal/moles of substance injected. (From ref. 63.)
266

the first and is used to determine the sample concentration. The signal from
the detector is modulated with an a.c. component at the switching frequency.
This may be amplified and synchronously demodulated. The signal-to-noise ratio
is increased owing to noise filtering. The selectivity is also increased 173 .
With respect to the required high flow-rates (excluding the use of capillary
columns) that reduce the sensitivity of the detector, the improvement in the
signal-to-noise ratio is not yery high if the detector itself is the source
of noise. The improvement is significant if the noise is caused by interfer-
ences 174 .

11.10. USE OF THE ELECTRON-CAPTURE DETECTOR WITH CAPILLARY COLUMNS

The volume of most commercial ECDs is so large that they cannot be used
directly in connection with capillary columns. In order to prevent the disper-
sion of chromatographic zones in the detector and, as a consequence, deteriora-
tion of the separation of substances on the capillary chromatographic column,
the volume of the ECD should be as small as possible. A peak with a baseline
width of 1 sec would elute from the capillary column in a gas volume equal to
16.7 ~18 at a flow-rate of 1.0 ml/min. If the peak distortion due to the
detector volume were not to exceed 1%, the residence time of the peak in the
detector would have to be less than 0.05 times the peak width at half-height 175 .
This would require a detector with a cell volume of about 0.5 ~l for the above
peak. The make-up gas used to reduce the effective volume of the detector
dilutes the sample and decreases the sensitivity, because the ECD is a con-
centration-sensitive detector.
The limit to the ECD volume is conditioned by the source type and the
geometry of the cell. The penetration depth of the B-particles in pure carrier
gas is 3 ,8 ca. 6-10 mm for 63Ni and ca. 2 mm for 3H• In order to ensure full
deactivation of all B-particles by collisions and their reduction to thermal
energies without colliding with the anode, the source-to-anode distance should
be larger than the penetration depth. For this reason .the tritium source is
better suited for ECD miniaturization (cf., section 11.3). However, the
distance should not be so large as to make efficient electron collection
impossible when a narrow, low-voltage pulse is applied to the anode 23 . It is
evident, therefore, that the above requirement of 0.5 ~l for the beta ioniza-
tion source is impractical.
A small-volume ECD using a plane-parallel electrode configuration was
described by Devaux and Guiochon 38 . The ECD with this configuration has a larger
volume than the cylindrical design with the same source area, which causes both
the standing current and the sensitivity to be lower. The coaxial design of the
267

high-temperature scandium tritide EC0 16 results in a volume of about 0.4 ml.


Even though the volume of the 3H cell is several times smaller than that of the
63Ni cell, the standing current is higher because higher specific activity foils
can be produced with tritium. The detection limit for testosterone dihepta-
fluorobutyrate was 13 about 2 pg. The volume of the cylindrical scandium tritide
microdetector is 120-160 ~1133,176. The asymmetric geometry of the cell
(displaced coaxial cylinder) with 63Ni allows the volume to be reduced to 350
~122, 177 to 100 ~1178 (the cell volume is smaller than that with the symmetrical
cell). The volume of the non-radioactive ECO with a thermionic emitter is 50
~133 (c.f., section 11.3).
However, detectors with the above volumes are still too large in view of
the peak width quoted in the first part of this section. For this reason, make-
up gas has still been used in this instance in order to reduce the effective
detector volume.
The influence of the gas flow-rate through the detector on the initial column
efficiency and on the peak shape with the microvolume 63Ni EC0 177 (380 ~l) is
shown in Fig. 11.27. Although the minimum flow-rate through the cell should be
about 68 ml/min in order to maintain the 90% efficiency with a l-sec peak when
assuming perfect mixing within the ECO chamber 179 , the compromise flow-rate is
about 30 ml/min 177 ,180,181. Complete mixing usually does not occur in the de-
tector, and with a cell designed to attain plug-like flow 169 ,182 (Fig. 11.28),

6 25
DIBJ)RIN
k'''22.2
5 2D
M
I
III
0 III
.- 4 1.5 ~
...J
~
UJ
~
Z 3 Will

2 05

10 0
10 20 30 40 50 60 '70 80
TOTAL FLOW-RATE, ml/min

Fig. 11.27. Effect of total detector gas flow-rate on the column efficiency
(NIL, plates/m) and peak shape (~) for capillary ECO. 25 m x 0.25 mrn 1.0. column
with QV-l01, 170 0 C. k' = Capacity factor. Skewness in (m3/m2)3/2 where m3 and
m2 are third and second moment, respectively. (From ref. 177.)
268

INSERT

SAMPLE

SIDE
PORTS

ANODE

CERAMIC
COLUMN
INSULATOR

Fig. 11.28. Micro-volume electron-capture detector cell. (From ref. 182.)

lpAa-II
30 .A.
" ! \
{ \/ \ .....•.....
I " \... ....
I I It.:\ ...
lI. I " .... \. ..
\ I ! I ~. ~ ~
20 \ I ; I ./ \
1;
1\;
I ~!'
a \.
I! \

i.
It-.i "'.. \
I j> r--
10
I' If
:

10 20 30 40 50 60
mil min
DETECTOR FLOW

Fig. 11.29. Peak height as a function of total detector flow-rate at different


pulse intervals: • = 160 ~sec; 0 = 290 ~sec; ~ = 530 ~sec; • = 750 ~sec. (From
ref. 180.)
269

the real flow-rates need not be extremely hi9h 182 In this 63Ni microdetector
the active region of the cell (the part of the cell below the insert from which
electrons can be extracted by the anode) is 100 ~l. The nitrogen make-up gas
flows from the bottom concentrically around the column. The sample flows from
the middle hole of the anode into the centre of the cell. In addition, the
sample flows along the sides of the cell through the cross-drilled holes in
the anode. This results in a nearly plug-like flow through the active region
of the cell. As the column effluent is not pre-mixed with the purge gas, the
gas flowing through the side ports and up along the sides of the cell walls
contains no sample and forms a boundary layer 182
The dependence of the response on the gas flow-rate through the ECD differs
in its nature from that of the common concentration-sensitive detectors,
reaching the maximum value 180 ,181,183,184. This course also depends on the
pulse intervals 180 (Fig. 11.29). As the maximum detectability for the ECD is
proportional to the concentration of the sample molecules at the peak maximum,
the detection limit is lower for capillary columns than for packed columns and
can be as low as a few femtograms 87 ,177,182,185.

REFERENCES

1 M. Krej~f and M. Dressler, Chromatogr. Rev., 13 (1970) 1.


2 W.A. Aue and S. Kapila, J. Chromatogr. Sci., 11 (1973) 255.
3 E.D. Pellizzari, J. Chromatogr., 98 (1974) 323.
4 E.R. Adlard, CRC Crit. Rev. AnaZ. Chem., 5 (1975) 13.
5 A. Zlatkis and D.C. Fenimore, Rev. Anal. Chem., 5 (1975) 217.
6 C.F. Poole, Lab. Pract., 25 (1976) 309.
7 S.O. Farwell, D.R. Gage and R.A. Kagel, J. Chromatogr. Sci., 19 (1981) 358.
8 C.F. Poole, J. High Resolut. Chromatogr. Chromatogr. Commun., 5 (1982) 454.
9 A. Zlatkis and C.F. Poole (Editors), EZectron Capture - Theory and Practice
in Chromatography, Elsevier, Amsterdam, 1981.
10 J.E. Lovelock, Anal. Chem., 33 (1961) 162.
11 J.E. Lovelock, in C.L.A. Harbourn (Editor), Gas Chromatography 1968,
Institute of Petroleum, London, 1969, p. 95.
12 J.E. Lovelock, Anal. Chem., 35 (1963) 474.
13 E.D. Pellizzari, J. Chromatogr., 92 (1974) 299.
14 C.F. Poole and A. Zlatkis, in A. Zlatkis and C.F. Poole (Editors), EZectron
Capture - Theory and Practice in Chromatography, Elsevier, Amsterdam, 1981,
Ch. 2, p. 13.
15 P.G. Simmonds, D.C. Fenimore, B.C. Pettitt, J.E. Lovelock and A. Zlatkis,
Anal. Chem., 39 (1967) 1428.
16 D.C. Fenimore, P.R. Loy and A. Zlatkis, Anal. Chem., 43 (1971) 1972.
17 M. Scolnick, J. Chromatogr. Sci., 7 (1969) 300.
18 W.L. Yauger, L.M. Addison and R.K. Stevens, J. Ass. Offic. Anal. Chem., 49
(1966) 1053.
19 J. Lasa, T. Owsiak and D. Kostewicz, J. Chromatogr., 44 (1969) 46.
20 C.H. Hartmann, AnaZ. Chem., 45 (1973) 733.
21 J.F. Uthe and J. Solomon, J. Chromatogr., 95 (1974) 169.
22 P.L. Patterson, J. Chromatogr., 134 (1977) 25.
23 J. Connor, J. Chromatogr., 210 (1981) 193.
270

24 G.R. Shoemake, D.C. Fenimore and A. Zlatkis, J. Gas Chromatogr., 3 (1965)


285.
25 G. Friedlander and J.W. Kennedy (Editors), Nuclear and Radiation Chemistry,
Wiley, New York, 1949.
26 C.H. Hartmann, D.Oaks and T. Burroughs, Technical Bulletin 130-66, Varian
Aerograph, l-Ialnut Creek, CA, 1966.
27 C. Kahn and M.C. Goldberg, J. Gas Chromatogr., 3 (1985) 287.
28 G.R. Shoemake, J.E. Lovelock and A. Zlatkis, J. Chromatoqr., 12 (1963) 314.
29 J.A. Lubkowitz and W.C. Parker, J. Chromatogr., 62 (1971) 53.
30 J.A. Lubkowitz, D. Montoloy and W.C. Parker, J. Chromatogr., 76 (1973) 21.
31 W.E. Wentworth, A. Tishbee, C.F. Batten and A. Zlatkis, J. Chromatogr., 112
(1975) 229.
32 D.J. Dwight, E.A. Lorch and J.E. Lovelock, J. Chromatogr., 116 (1976) 257.
33 A. Neukermans, W. Kruger and D. McManigill, J. Chromatogr., 235 (1982) 1.
34 S. Kapila, D.J. Bornhop, S.E. Manahan and G.L. Nickell, J. Chromatogr., 259
(1983) 205.
35 J.E. Lovelock and S.R. Lipsky, J. Amer. Chem. Soc., 82 (1960) 431.
36 P. Devaux and G. Guiochon, J. Gas Chromatogr., 8 (1970) 502.
37 S. Kapila and W.A. Aue, J. Chromatogr., 108 (1975) 13.
38 P. Devaux and G. Guiochon, J. Gas Chromatogr., 5 (1967) 341.
39 P. Devaux and G. Guiochon, Chromatogrpahia, 2 (1969) 151.
40 C. Gosselin, G.B. Martin and A. Boudreau, J. Chromatogr., 90 (1974) 113.
41 J.E. Lovelock, Nature (London), 189 (1961) 729.
42 J.E. Lovelock and N.L. Gregory, in N. Brenner, J.E. Callen and M.D. Weiss
(Editors), 3rd International Symposium on Gas Chromatography, Academic Press,
New York, 1962, p. 219.
43 W.E. Wentworth, R.S. Becker and R. Tung. J. Phys. Chem., 70 (1966) 445.
44 R.J. Maggs, P.L. Joynes. A.J. Davies and J.E. Lovelock, Anal. Chem., 43
(1971) 1966.
45 W.A. Aue and K.W.M. Siu, Anal. Chem .• 52 (1980) 1544.
46 K.W.M. Siu, C.M. Roper, L. Ramaley and W.A. Aue. J. Chromatogr., 210 (1981)
401.
47 K.W.M. Siu and W.A. Aue, J. Chromatogr., 268 (1983) 273.
48 M.M. Shanin and S.R. Lipsky, Anal. Chem., 35 (1963) 467.
49 W.E. Wentworth, R.S. Becker and R. Tung, J. Phys. Chem., 71 (1967) 1652.
50 W.E. Wentworth and E. Chen, J. ~as Chromatogr., 5 (1967) 170.
51 W.E. Wentworth and E.C.M. Chen, J. Chromatogr., 186 (1979) 99.
52 E.C.M. Chen and W.E. Wentworth, J. Chromatogr., 217 (1981) 151.
53 W.E. Wentworth and E.C.M. Chen, in A. Zlatkis and C.F. Poole (Editors),
Electron Capture - Theory and Practice in Chromatography, Elsevier, Amsterdam,
1981, Ch. 3, p. 27.
54 J. Connor, J. Chromatogr., 200 (1980) 15.
55 J.E. Lovelock. J. Chromatogr., 99 (1974) 3.
56 J.E. Lovelock and A.J. Watson, J. Chromatogr., 158 (1978) 123.
57 W.E. Wentworth and J.C. Steel hammer. Advances in Chemistry Series, Radiation
Chemistry, American Chemical Society. Washington, DC, 1968, Ch. 4.
58 W.E. Wentworth, in 1.1. Damsky ann J.A. Perry (Editors), Recent Advances in
Gas Chromatography, Marcel Dekker, New York, 1971, Ch. 5.
59 E.P. Grimsrud, S.H. Kim and P.L. Gobby, Anal. Chem .• 51 (1979) 223.
60 M.W. Siegel and t4.C. McKeown, J. Chromatogr., 122 (1976) 397.
61 P.L. Gobby, E.P. Grimsrud and S.W. Warden, Anal. Chem., 52 (1980) 473.
62 E.P. Grimsrud and S.W. Warden, Anal. Chem., 52 (1980) 1842.
63 W.A. Aue and S. Kapila, J. Chromatogr., 188 (1980) 1.
64 LP. Grimsrud and t~.J. Connolly, J. Chromatogr., 239 (1982) 397.
65 S. Kapila, C.R. Vogt and W.A. Aue, J. Chromatogr., 195 (1980) 17.
66 S. Kapila, C.R. Vogt and W.A. Aue, J. Chromatogr., 196 (1980) 397.
67 S. Kapi1a and W.A. Aue, J. Chromatogr., 118 (1976) 233.
68 W.A. Aue and K.W.M. Siu, J. Chromatogr., 203 (1981) 237.
69 K.W.M. Siu and W.A. Aue, Microchim. Acta, I (1983) 419.
271

70 W.A. Aue and K.W.M. Siu, J. Chpomatogp., 239 (1982) 127.


71 C.A. Clemons and A.P. Altshuller, Anal. Chem., 38 (1966) 133.
72 J. Vessman, in A. Zlatkis and C.F. Poole (Editors), Electpon Captupe -
Theopy and Practice in Chpomatogpaphy, Elsevier, Amsterdam, 1981, Ch. 7,
p. 137.
73 J.E. Lovelock, P.G. Simmonds and W.J.A. VandenHeuvel, Natupe (London), 197
(1963) 249.
74 J.E. Lovelock, Phys. Processes Radiat. Biol., Proc. Int. Symp., 1963, (1964)
183.
75 A. Zlatkis and J.E. Lovelock, CZin. Chem., 11 (1965) 259.
76 H.J. Dawson, Jr., Anal. Chem., 35 (1963) 542.
77 J.E. Lovelock and A. Zlatkis, Anal. Chem., 33 (1961) 1959.
78 S. Yamaguchi and H. Matsumoto, KUPUme Med. J., 16 (1969) 33.
79 J. Vessman, J. Chpomatogp., 184 (1980) 313:
80 W.L. Zielinski, Jr., L. Fishbein and R.O. Thomas, J. Chpomatogp., 30 (1967)
77.
81 W.L. Zielinski, Jr., L. Fishbein and L. Martin, Jr., J. Gas Chpomatogp., 5
(1967) 552.
82 R.A. Landowne and S.R. Lipsky, Anal. Chem., 34 (1962) 726.
83 J.J. Sullivan, J. Chpomatogp., 87 (1973) 9.
84 J.A. Sweetman and E.A. Boettner, J. Chpomatogp., 236 (1982) 127.
85 W.P. Cochrane, R.B. Maybury and R.G. Greenhalgh, J. Envipon. Sci. Health,
B14 (1979) 197.
86 G. Ecklund, B. Josefsson and C. Ross, J. High Resolut. Chpomatogp.
Chpomatogp. Commun., 1 (1978) 34.
87 J. Wessman and P. Hartvig, Acta Phapm. Suecica, 8 (1971) 235.
88 A.C. Moffat and E.C. Horning, Anal. Lett., 3 (1970) 205.
89 O.K. Albert, Anal. Chem., 36 (1964) 2034.
90 C.F. Poole and A. Zlatkis, in A. Zlatkis and C.F. Poole (Editors), Electpon
Captupe - Theopy and Practice in Chpomatogpaphy, Elsevier, Amsterdam, 1981,
Ch. 8, p. 151.
91 A. Zlatkis and B.C. Pettit, Chpomatogpaphia, 2 (1969) 484.
92 A. Zlatkis, C.K. Lee, W.E. Wentworth and E.C.M. Chen, Anal. Chem., 55 (1983)
1596.
93 E.C. Chen and W.E. Wentworth, J. Chpomatogp., 68 (1972) 302.
94 C.F. Poole, J. Chromatogp., 118 (1976) 280.
95 K. Blau and G.S. King, in K. Blau and G.S. King (Editors), Handbook of
Depivatives fop Chpomatogpaphy, Heyden, London, 1978, p. 135.
96 C.F. Poole and A. Zlatkis, Anal. Chem., 52 (1980) 1002A.
97 K. Blau and G.S. King (Editors), Handbook of Depivatives fop Chpomatogpaphy,
Heyden, London, 1978.
98 C.D. Knapp, Handbook of Analytical Depivatization Reactions, Wiley, New
York, 1979.
99 J. Drozd, Chemical Depivatization in Gas Chpomatogpaphy, Elsevier, Amsterdam,
1981.
100 E. Anggard and A. Hankey, Acta Chem. Scand., 23 (1969) 3110.
101 L. Dehennin, A. Reiffstock and R. Scholler, J. Chpomatogp. Sci., 10 (1972)
224.
102 N.K. McCallum and R.J.X. Armstrong, J. Chpomatogp., 78 (1973) 303.
103 J.F. Lawrence, J.J. Ryan and R. Ledue, J. Chpomatogp., 147 (1978) 398.
104 J.J. Ryan and J.F. Lawrence, J. Chpomatogp., 135 (1975) 117.
105 D.O. Clarke, S. Wilk and S.E. Gitlow, J. Gas Chpomatogp., 4 (1966) 310.
106 A.J. Watson, G.L. Ball and D.H. Stedman, Anal. Chem., 53 (1981) 132.
107 C.F. Poole and A. Zlatkis, J. Chromatogp. Sci., 17 (1979) 115.
108 C.F. Poole, S. Singhawangcha, L.E. Chen Hu, W.F. Sye, R. Brazell and A.
Zlatkis, J. Chr>o"latogl'., 187 (1980) 331.
109 C.F. Poole, W.F. Sye, S. Singhawangcha, F. Hsu, A. Zlatkis, A. Arfwidsson
and J. Wesman, J. Chpomatogp., 199 (1980) 123.
110 P.M. Burkinshaw, E.D. Morgan and C.F. Poole, J. Chpomatogp., 132 (1977) 548.
272

111 D. Morgan and C.F. Poole, J. Chromatogr., 89 (1974) 225.


112 D. Morgan and C.F. Poole, J. Chromatogr., 104 (1975) 351.
113 J. Francis, LD. Morgan and C.F. Poole, J. Chromatogr., 161 (1978) 111.
114 O. Gyllenhaal and P. Hartvig, J. Chromatogr., 189 (1980) 351.
115 B.C. Pettitt, P.G. Simmonds and A. Zlatkis, J. Chromatogr. Sci., 7 (1969)
645.
116 F. Walle and H. Ehrsson, Acta Pharm. Suecica. 7 (1970) 389.
117 C.F. Poole, S. Singhawangcha and A. Zlatkis, J. Chromatogr., 158 (1978) 33.
118 C.F. Poole, S. Singhawangcha and A. Zlatkis, J. Chromatogr., 186 (1979) 307.
119 A.W. Holden and G.A. Wheatley, J. Gas Chromatogr., 5 (1967) 373.
120 M.J. Hartigan, J.E. Purcell and E.W. March, Chromatogr. Newsl., 3 (1974) 23.
121 C.H. Hartmann, D.M. Oaks and K.P. Dimick, presented at Pittsburgh Conference,
February 21-25, 1966.
122 G.G. Guilbault and C. Herrin, Anal. Chim. Acta, 36 (1966) 252.
123 H.J. van de Wiel and P. Tommassen, J. Chromatogr., 71 (1972) 1.
124 F.W. Karasek and D.M. Kane, Anal. Chem., 45 (1973) 576.
125 D.C. Fenimore, A. Zlatkis and W.E. Wentworth, Anal. Chem., 40 (1968) 1594.
126 D.C. Fenimore and C.M. Davies, J. Chromatogr. Sci., 8 (1970) 519.
127 E. Chen and J. E. Lovelock, J. Phys. Chem., 70 (1966) 445.
128 Electron Capture Detector for SP 7100 Gas Chromatograph, 195D 10/82, Spectra-
Physics, San Jose, CA, 1982.
129 J.J. Sullivan and C.A. Burgett, Chromatographia, 8 (1975) 176.
130 E.P. Grimsrud and W.B. Knighton, Anal. Chem., 54 (1982) 565.
131 W.B. Knighton and E.P. Grimsrud, AnaL Chem., 55 (1983) 713.
132 E.P. Grimsrud and D.A. Miller, J. Chromatogr., 192 (1980) 117.
133 B. BrechbUhler, L. Gay and H. Jaeger, Chromatographia, 10 (1977) 478.
134 R. Simon and G. Wells, J. Chromatogr., 302 (1984) 221.
135 W.B. Knighton and LP. Grimsrud, J. Chromatogr., 288 (1984) 237.
136 W.E. Wentworth, E. Chen and R.R. Freeman, J. Chem. Phys., 55 (1971) 2175.
137 W.E. Wentworth and R.R. Freeman, J. Chromatogr., 79 (1973) 322.
138 P.G. Simmonds, J. Chromatogr., 166 (1978) 593.
139 M.P. Phillips, R.E. Sievers, P.D. Goldan, W.C. Kuster and F.C. Fensenfeld,
AnaL Chem., 51 (1979) 1819.
140 F.C. Fehsenfeld, P.D. Goldan, M.P. Phillips and R.E. Sievers, in A. Zlatkis
and C.F. Poole (Editors), Electron Capture - Theory and Practice in
Chromatography, Elsevier, Amsterdam, 1981, Ch. 4, p. 69.
141 M.A. Wizner, S. Singhawangcha, R.M. Barkley and R.E. Sievers, J. ChroI7atogl'.,
239 (1982) 145.
142 P.D. Goldan, F.C. Fehsenfeld and M.P. Phillips, J. Chromatogr., 239 (1982)
115.
143 R.E. Sievers, M.P. Phillips, R.M. Barkley, M.A.Wizenr, M.J. Bollinger, R.S.
Hutte and F.C. Fehsenfeld, J. Chromatogr., 186 (1979) 3.
144 P.D. Goldan, F.C. Fehsenfeld, F.C. Kuster, M.P. Phillips and R.E. Sievers,
AnaL Chem., 52 (1980) 1751.
145 E.P. Grimsrud and R.G. Stebbins, J. Chromatogr., 155 (1978) 19.
146 E.P. Grimsrud and D.A. Miller, Anal. Chem., 50 (1978) 1141.
147 D.A. Miller and E.P. Grimsrud, Anal. Chem., 51 (1979) 851.
148 G. di Pasquale and T. Capaccioli, J. Chromatogr., 206 (1981) 589.
149 LP. Grimsrud, in A. Zlatkis and C.F. Poole (Editors), Electron Capture -
Theory and Practice in Chromatography, Elsevier, Amsterdam, 1981, Ch. 5,
p. 91.
150 E.P. Grimsrud, D.A. Miller, R.G. Stebbins and S.H. Kim, J. Chromatogr.,
197 (1980) 51.
151 D.A. Miller, K. Skogerboe and E.P. Grimsrud, Anal. Chem., 53 (1981) 464.
152 J.A. Campbell, E.P. Grimsrud and L.R. Hageman, Anal. Chem., 55 (1983) 1335.
153 J.A. Campbell and E.P. Grimsrud, J. Chromatogr., 284 (1984) 27.
154 C.J. Kallos, Anal. Chem., 53 (1981) 963.
155 E.P. Grimsrud, S.W. Warden and R.G. Stebbins, Anal. Chem., 53 (1981) 716.
156 J.A. Campbell and E.P. Grimsrud, J. Chromatogr., 243 (1982) 1.
273

157 J.A. Campbell and E.P. Grimsrud, J. Chromatogr., 291 (1984) 13.
158 E.P. Grimsrud, Anal. Chem., 56 (1984) 1797.
159 E.P. Grimsrud and O.A. Valkenburg, J. Chromatogr., 302 (1984) 243.
160 E.P. Grimsrud, J. Chromatogr., 312 (1984) 49.
161 J.E. Lovelock, R.J. Maggs and E.R. Adlard, Anal. Chem., 43 (1971) 1962.
162 D. Lillian and H.B. Singh, Anal. Chem., 46 (1974) 1060.
163 H.B. Singh, D. Lillian and A. Appleby, Anal. Chem., 47 (1975) 860.
164 E. Bros and F.M. Page, J. Chromatogr., 126 (1976) 271.
165 J. Rosiek, I. Sliwka and J. Lasa, J. Chromatogr., 137 (1977) 245.
166 J. Lasa, E. Bros, I. Sliwka, A. Korus and J. Rosiek, Report No. 1029/C4,
Institute of Nuclear Physics, Krakow, 1978.
167 E.P. Grimsrud and S.H. Kim, Anal. Chem., 51 (1979) 537.
168 P. Popp and J. Leonhardt, Isotopenpraxis, 19 (1983) 160.
169 W.A. Aue and S. Kapila, J. Chromatogr.{ 112 (1975) 247.
170 D.C. Legget, J. Chromatogr., 133 (1977) 83.
171 J.E. Lovelock, J. Chromatogr., 112 (1975) 29.
172 P.G. Simmonds, A.J. Lovelock and J.E. Lovelock, J. Chromatogr., 126 (1976) 3.
173 P.R. Boshoff and B.J. Hopkins, J. Chromatogr. Sci., 17 (1979) 588.
174 G. Well s, J. Chrornatogr., 285 (1984) 395.
175 J. Sevcik and J.E. Lips, Chromatographia, 12 (1979) 693.
176 H.R. Buser, Anal. Chem., 48 (1976) 1553.
177 F.J. Yang and S.P. Cram, J. High Resolut. Chromatogr. Chromatogr. Commun.,
2 (1979) 487.
178 G. Wells and R. Simon, J. High Resolut. Chromatogr. Chromatogr. Commun., 6
(1983) 427.
179 A.P.J.M. de Jong, J. High Resolut. Chromatogr. Chromatogr. Commun., 5 (1982)
213.
180 J.J. Franken and H.L. Vader, Chromatographia, 6 (1973) 22.
181 L. Rejthar and K. Tesarik, J. Chromatogr., 131 (1977) 404.
182 G. Wells, J. High Resolut. Chromatogr. Chromatogr. Commun., 6 (1983) 651.
183 P. Devaux and G. Guiochon, J. Chromatogr. Sci., 7 (1969) 561.
184 K.P. Dimick and H. Hartmann, Aerograph 1/63 W-106, Wielkens Instrument,
Walnut Creek, CA, 1963; presented at the ACS Winter Meeting, Cincinnati,
Ohio, January 1963.
185 G. Eklund, B. Josefsson and C. Ross, J. Chromatogr., 142 (1977) 575.
This Page Intentionally Left Blank
275

Chapter 12

ION MOBILITY DETECTOR

CONTENTS

12.1. Introduction . . . . . • • 275


12.2. Principle of the technique 275
12.3. Detection principles 279
12.4. Effect of background 286
References . . . . 288

12.1. INTRODUCTION

The ion mobility spectrometer was developed in 1970 by Karasek and Cohen 1- 3
for organic trace analysis. The original name of this technique, plasma chro-
matography4, relates to the application of ions (plasma) and to the analogy
between ion mobility separation and chromatography. The term plasma chromato-
graphy has been used until now.

12.2. PRINCIPLE OF THE TECHNIQUE

The principle of the method is apparent from the schematic diagram of an ion
mobility detector (IMD) Fig. 12.1. A mixture of the effluent from the chro-
matographic column and the reactant gas (nitrogen or air) enters the detector
space, passing adjacent to a radioactive 8-particle source (63Ni , about 10 mCi).
In the ionization space of the detector a weak plasma consisting of positive
and negative ions is generated. Ions such as (H20)nH+ where n = 2-4, and
(H 20) NO+ and (H 20) NH4+ ions where n = 0-2 are generated in nitrogen in the
n n 4-7
presence of a small amount of water . The value of n depends on the water
content of the gas and on temperature. The negative particles are low-energy
electrons (about 0.5 eV). (H20)n02 ions are formed additionally when air is
used as the reactant gas. When molecules of the sample enter the reaction
space, the mentioned ions and electrons react with the sample molecules, giving
product ions. The product ions formed, together with the unreacted initial
reactant ions, are released in pulses (for 0.2 msec in intervals of about
10 msec) through the shutter grid into the drift space of the detector. Under
the effect of ca. 200-300 V/cm electric field, the ions move to the collector
276

DRIFT GAS

I~
7 6

0

0

• • 5
0 0


0

0 4

2 3
PLASMACHROMATOGRAM

Fig. 12.1. Schematic diagram of the ion mobility spectrometer (plasma chromato-
graph). 1 = Polarization electrode; 2 = shutter grid; 3 = gate grid; 4 = col-
lector electrode; 5 = signal space; 6 = drift space; 7 = reaction space; 8
gas exhaust. (Reprinted from ref. 2, with permission.)

electrode. Within this drift (separation) space, the ions are separated at
normal pressure on the basis of their different mobilities, the heavier ions
moving more slowly. Nitrogen has commonly been used as the drift gas, because
it prevents additional reactions of the ions with the molecules. Positive or
negative ions can be observed by selecting the polarity of the electric field.
Various ions arrive at the gate grid at various times. If the gate grid is
open, the ions pass through it and are collected at the electrode.
The drift speed Vd (the speed of ion movement through the drift space) is
proportional to the intensity of the electric drift field E (ref. 4):

(12.1)

where"K is the linear ion mobility. The IMD measures the time required for the
ion to migrate to a fixed distance. Hence, Vd can be replaced with drift length,
d, divided by drift time, T, between the ion injected grid and the collector:

K = .E.... (12.2)
TE

The following equation applies for the ion mObi1ity8:

..l. e rl ll1/2 [21TJ 1/2 1 + Il


K = 16 • N Lm + MJ LkTJ • 2n{ 1, 1)* (12.3)
l'm"
277

1
4.8 em

1 4

10----1
5

9-=;:=13
8---1
6

Fig. 12.2. Schematic diagram of the modified ion mobility spectrometer. 1 =


Sample inlet; 2 = shutter grid; 3 = protective stainless~steel ring; 4 = gate
grid; 5 = passive grid electrically connected to the protective ring; 6 = col~
lector electrode; 7 = electrometer; 8 = PTFE insulation; 9 = drift gas inlet;
10 = glass insulation; 11 = high voltage (e.g. +3000 V plate; 12 = gas exhaust.
(Reprinted with permission from ref. 7.)

where N is the molecular number density, k is the Boltzmann constant, e is the


electronic charge, T is the absolute temperature, ~ is a small correction term
for higher approximations, Pm is the position of minimum potential for the
interaction between the ion and molecule, Q(1,1)* is the first~order collision
integral, M is the molecular mass and m is the ionic mass.
For the standard conditions T = 273 K and p = 760 mmHg, the following rela~
tion is valid:
278

12
10
11 13

15
9

14

7
4

40 ·c
Fig. 12.3. Gasoline chromatogram, obtained in non-selective reactant ion mode.
1 = Heptane; 2 = methylcyclohexane; 3 = toluene; 4 a m,p-xylene, 5 = a-xylene;
6 = trimethylbenzenes; 7 = ethyltoluene; 8 = tert.-butylbenzene, 9 = naphthalene;
10 = dodecane; 11-13 = unknowns; 14 = 2-methylnaphthalene; 16 = l-methyl-
naphthalene; 17,18 = unknowns. 15 m quartz capillary column, SE-54; temperature,
programmed from 40 to 100 0 e at 8oe/min, held for 10 min at 100oe. Ion monitoring
with drift time between 8 and 9 msec, gradient 215 V/cm; drift gas, nitrogen,
600 ml/min; detector temperature, 140 oe. R = response. (Reprinted with permis-
sion from ref. 7.)

K =!l... 76PO • 273 (12.4)


o TE T

where Ko is the reduced mobility, depending on the ionic size, charge and mass,
on the molecular size and mass, and on the drift gas composition and polar-
izability4. Values of Ko have been published for a number of compounds: n-alkyl
halides 9 , substituted aromatics 10 , isomeric halogenated nitrobenzenes 11 , tri-
nitrotoluene 12 , n-alkanes 13, phthalic acids 14,15, lysergic acid diethylamide
and ~9_tetrahydrocannibinol16, n-alkyl acetates 17 , heroin and cocaine 18 ,
n-alkanols 19 , aliphatic n-nitrosamines 20 , barbiturates 21 and polychlorinated
biphenyls22.
279

R
567

Fig. 12.4. Gasoline chromatogram, obtained with an FlO. For description and
chromatographic conditions, see Fig. 12.3. (Reprinted with permission from
ref. 7.)

12.3. DETECTION PRINCIPLES

The equipment described above has been used as a detector combined with a
gas chromatograph2,14,21 ,23-25. Commercially produced IMD (plasma chromato-
graphs), when used in gas chromatography, have shown some disadvantages, mainly
owing to the large inner volume, absorption phenomena occurring in the detector
and the influence of the stationary phase molecules on the spectra. Baim and
Hill? described a detector (Fig. 12.2) modified so as to be usable with capil-
lary columns. The modifications are as follows. (1) The gas flow through the
ionization space is reversed and, as a result, the gas flows through the entire
detector in only one direction. The drift gas enters the detector near the
collecting electrode and flows through the drift and ionization spaces towards
2S0

FID
A ~ II E-5 (I !~~KGR
~~msec E-8
(~ I
I
~~08~~'
20
x4
40msec

r~'
~
';' (io~x1c:2C~c: :t: :~lOXo4 '-2

40msec
p 20 40msec

OL-~=-~~~----~~~-- __~~~~__-----
o 2 ,4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 min

Fig. 12.5. Chromatogram and ion mobility spectra for freons (E-1 to E-10).
negative ion monitoring; drift and reaction gas, air; detector temperature,
1990 C; 6% SE-30 on Anakron ASS, 1.S3 m x 3.2 mm 1.0. column; temperature, 100 0 C;
BKGR, background spectrum. (Reprinted from ref. 24, with permission.)

the detector outlet. In earlier equipment the gases were introduced at the
opposite sides of the detector, leaving the detector near the shutter grid.
(2) The volume of the ionization cell is reduced to 1 ml, in contrast to the
original 7 ml. (3) The detector inlet is situated 'between the ionization cell
and the drift space. Hence the sample is removed from the ionization space by
means of the drift-gas flow through the ionization space. Therefore, neutral
sample molecules, neutral products or radicals cannot interact with the product
ions moving through the drift space. (4) The rings forming the wall of the drift
tube are separated from each other by rings made of borosilicate glass. This
closed drift tube increases the efficiency of removal of neutral species from
the detector. The chromatograms obtained with this detector and those obtained
with a flame-ionization detector (FID) are compared in Figs. 12.3 and 12.4.
The IMO can be used as a detector in gas chromatography in several modes.
(1) If the gate grid is closed, no detection takes place. If this grid is
opened at intervals gradually increasing compared 'with opening the shutter grid,
the ion mobility spectrum is monitored. This means that for each peak leaving
281

a-.............-...I'---
I iii i I

50 62 74 86·C

Fig. 12.6. Chromatograms of a mixture of 30 ng of iodobenzene and 180 ng of


chlorobenzene. (a) Selective record of the negative ions, 1- drift time 7.14
msec; (b) selective record of the negative ions, Cl- drift time 6.07 msec; (c)
electron monitoring. Drift time, 0.2 msec; 214 V/cm; drift gas, nitrogen;
detector temperature 140 0 C; 1.22 m x 4 mm 1.0. glass column, packed with Ultra-
Bond; temperature, programmed from 50 to 80 0 C at 60 C/min. 1 = Chlorobenzene;
2 = iodobenzene. Arrows indicate injection. (From ref. 26.)

the chromatograph the corresponding spectrum characterizing the given compound


is obtained from the spectrum, being either positive or negative according to
the polarity chosen. An example of these spectra for commercial freons is given
in Fig. 12.5.
(2) If the gate grid is opened only for certain fixed intervals subsequent
to the shutter grid, only ions of a certain mobility are monitored 7 ,26-29. Hence
the IMD becomes a selective detector responding only to compounds producing
ions that move for the chosen drift time. An example showing the application
of this mode of detection is given in Figs. 12.6 and 12.7 for the negative
product ion mode. The positive ions can be detected in a'similar manner 7 •29 •
(3) If the gate grid is opened so as to monitor the drift time corresponding
to the time of the reactant ions, the detector background current corresponds
to the detection of these ions. As the solute molecules react with these ions,
giving product ions of different mobility, the presence of organic molecules
is associated with a decrease in the background current and the response is
negative. The resulting chromatogram is either the record of the decrease in
negative ions 26 ,28 (Figs. 12.6 and 12.7) or a non-selective record of the
decrease in positive ions 7,26. When oxygen is used as the reactant gas, (H20)n02
282

NON-SELECTIVE MODE SELECTIVE MODE

iii iii
,.! I- ...:
U
4: 4:
UJ
N
..., ~
I ~
0 0
~
~

L/'I
L/'I
...!

100 150 200 HOLD 100 150 200 HOLD

TEMPERATURE,OC

Fig. 12.7. Chromat09ra~ of soil sample extracted with acetone-hexane. (a) Mon-
itoring of the (H20)n02 reactant ions; (b) selective record of 2,4-dichloro-
phenoxyacetic acid product ion. Drift time, 7.90 msec; 230 V/cm; 15 m SE-54
fused-silica capillary column. Temperature, programmed from 100 to 200 0 C at
10 0 C/min. 1 = 2,4-Dichlorophenoxyacetic acid. (From ref. 28.)

ions are formed, the monitoring of which again gives the complete recording 27 ,28_
In the negative reactant ion mode, the detector responds only to compounds that
capture an electron. The response is therefore selective again, analogous to
the electron-capture detector in this instance. An example of the non-selective
positive reactant ion mode using a capillary column is shown in Fig. 12.3. The
chromatogram resembles that obtained with the same mixture by means of an FID
(Fig. 12.4).
(4) The detector can operate in the non-selective mode also with a positive
response 7 ,29. All product ions with drift times within a certain interval are
monitored.
The IMD when used as a gas chromatographic detector in mode 3 (the non-
selective positive reactant ion mode 7) and in mode 1 (the IMD spectrum23 ,25)
is one order of magnitude more sensitive, and in mode 3 (the selective negative
reactant ion mode with the use of oxygen 27 ) up to two orders of magnitude more
sensitive than the FID. The dependence of the response on the amount of the
compound is not linear 25 ,27,28.
The utilization of a photoionization source (UV lamp) instead of a 63Ni
ionization source results in a lack of reactant ions and uncomplicated fragmenta-
283

1.0

0.5

0
0 5 10 15 20
I REACTANT ION
REGION
I PRODUCT ION
REGION

1.0

~
<{
0.5

0
0
f
5
J , t t
10 15
f
20
DRIFT T I ME, msec

Fig. 12.8. Comparison of benzene ion mobility spectra using 63Ni (top) and
photoionization (bottom) sources. (Reprinted with permission from ref. 29.)

tion patterns 29 (see Fig. 12.8). As the energy of the 10.00-eV Kr lamp is
substantially lower than the ionization potential of nitrogen (15.58 eV), no
ionization occurs and, hence, no background peaks can be observed. The advan-
tage of this fact is obvious from Fig. 12.8. The reactant ions present in the
63Ni ionization spectrum obscure a portion of the ion mobility scan from 6 to
8 msec. Product ions of any kind having mobilities similar to those of the
reactant ions cannot be observed in the form of discrete peaks separate from
the reactant ions. The first two peaks from the three product ion peaks present
in the 63Ni spectrum of benzene are not completely resolved from the (H20)nH+
reactant ion peak. When using a UV lamp, a single large production peak with
a drift time of 7.31 msec is observed. This drift time is virtually identical
with that of the (H20)nH+ reactant ion with 63 Ni ionization.
Laser multi-photon ionization (MPI)30,31 was also used as an ionization
source. The MPI process allows the direct ionization of organic compounds with
the production of only one peak, which is either the molecular ion or MH+. Multi-
wavelength-selective ionization of organic compounds in the IMD can be obtained
284

TABLE 12.1

IONIZATION AS A FUNCTION OF WAVELENGTH

IP = Ionization potential; + = high response; ± = weak response; - no response.


Reprinted with permission from ref. 31.

Compound Mol IP Expected Wavelength (nm)


wt. (eV) cut-off
(nm) 320 310 293 280 266 249 194
Anil i ne 93 7.7 322 ± + + + + + +
N-Methyl anil i ne 107 7.32 338 + + + + + + +
m-Toluidine 107 7.5 330 + + + + + + +
2,4-Lutidine 107 8.85 280 ± + + + +
p-(n-Butyl)aniline 149 7.53 329 + + + + + + +
N,N-Diethylaniline 149 6.99 354 + + + + + + +
N,N,3,5-Tetra-
methylaniline 149 7.25 342 + + + + + + +
N,N-Dimethyl anil ine 121 7.14 347 + + + + + + +
2,4-Dimethylaniline 121 7.44 333 + + + + + + +
Hexylamine . 101 ± ± + + + + +
Diisopropylamine 101 7.73 320 ;!: + + + + + +
Triethylamine 101 7.50 330 + + + + + +
tert.-Butylamine 73 8.64 287 + + + + +
sec.-Butylamine 73 8.70 285 ± + + + +
n-Butylamine 73 8.71 284 ± + + + + +
Methylamine 31 8.97 276 + + +
Formamide 45 10.25 242 +
Dimethylformamide 80 9.12 ' 271 ± + + +
Benzene 78 9.23 268 + + +
Toluene 92 8.82 281 ± + + + +
Xylene 106 8.5 291 + + + + +
Phenol 94 8.51 291 + + + +
Cresol 108 8.52 291 + + + +
Cyclohexane 84 9.8 253 + +
Hexane 86 9.45 262 ± + + +
Indene 116 8.81 281 ± + + + + + +
Pyridine 79 9.3 266 ± ± + + +
Pyrrole 67 8.2 302 + + + + +
Naphthalene 128 8.13 305 + + + + +
Azulene 128 7.42 334 ± + + + +
Anthracene 178 7.43 333 ± + + + + + +
Phenanthrene 178 7.80 317 ± + + + +
Tetracene 228 7.01 353 + + + + + + +
p-Nitrotoluene 137 9.82 252 ± +
2,6-Dinitrotoluene 182 +
Methanol 32 10.84 228 +
Ethanol 46 10.49 236 +
l-Nonanol 144 ± +
1-0ctanol 130 ± +
Acetone 58 9.98 248 + +
Benzophenone 182 9.4 263 + +
Benzaldehyde 106 9.52 260 ± + + +
p-Dioxane 88 9.13 271 ± + +
Ethyl formate 74 10.61 233 ± +
285

Drift Time (msec)


5 10 15 20 25
r--+--+--4~Ir-4------

0.5 19.1 msec


2QOmsec A

to
DriftTime (msec)
VJ 5 10 15 20 25
:;
0
>

"5
c
.21 0.5
(/)
c B
0
:.;:;
0
N to
·c 20.0 msec
.2
Drift Time (msec)
5 10 15 20 25

13.6 msec
0.5 C

to

Fig. 12.9. Ion mobility spectrum of p-xylene and N-methylamine. (A) Laser source,
266 nm; T = 220 oC; input laser energy, 0.1 mJ; laser beam size, 2 x 6 mm rect-
angle; electric field, 170.3 V/cm; drift gas flow-rate 600 cm 3/min. (B) Laser
source, 310 nm; input laser energy, 0.25 mJ; other conditions as in (A). (C) 63Ni
source. (Reprinted with permission from ref. 31.)

with MPI 31 . In resonant two-photon ionization, a molecule will ionize if the


two-photon energy is greater than the ionization potential of the molecule and
if there is a real intermediate state resonant with the first photon. Table 12.1
presents the ionization results for 44 organic compounds in an IMD based on eight
different wavelengths in the UV region. It is evident that amines, for instance,
can be distinguished from aromatic hydrocarbons by this wavelength selectivity.
This effect is illustrated in Fig. 12.9 for the N-methylaniline and p-xylene
286

TABLE 12.2

ISOMER REDUCED ION MOBILITIES (Ko) AND AVERAGE COLLISIONAL CROSS-SECTIONS (~D)

Ko is in cm 2/V-sec and nD in ~2. Reprinted with permission from ref. 32.

Compounds Ion Ortho Meta Para

Ko ~D Ko ~D Ko ~D

Fluorotoluenes + 1.939 115.9 1.951 115.2 1. 951 115.2


Dimethoxybenzenes 1*+ 1.733 127.0 1.761 125.0 1.766 124.7
2*+ 1.738 126.7 1.757 125.3 1.763 124.8
Methylphenetoles + 1.706 129.2 1.703 129.5 1. 716 128.5
Toluic acid methyl esters + 1.680 130.2 1.649 132.6 1.671 130.9
1.691 129.3 1.663 131.5 1.674 130.6
Acetotoluidides + 1.623 134.8 1.602 136.6 1.614 135.5
1.610 135.9 1.587 137.9 1.599 136.8
Phthalic acid methyl esters + 1.540 138.5 1.450 148.0 1.466 146.5
1. 527 140.6 1.481 145.0 1.497 143.5

* 1, The original samples were analysed by gas chromatographic injection of the


peaks into the plasma chromatograph; 2, the same samples injected in methanol
solvent directly into the plasma chromatograph inlet three days later.

pair. Both compounds ionize at 266 nm and the two peaks differing in their
mobilities partially overlap. The N-methylaniline peak only appears at 310 nm.
The combination of gas chromatography with an IMD allows us to distinguish
25
ortho-, meta- and para-isomers and also ais- and trans-isomers ,32. Isomers
give different drift times. Meta-substituted isomers are larger than para- and
ortho-isomers (compare their reduced ion mobilities and average collision cross-
sections in Table 12.2). Therefore, plasma chromatography is isomer selec-
tive 25 ,32,33. The sensitivity of detection is also structure dependent 25 ; for
instance, the detection limit for p-chlorodiphenyl oxide is about 0.01 ng and
for o-chlorodiphenyl oxide it is about 0.4 ng.

12.4. EFFECT OF BACKGROUND

Column bleed from different liquid stationary phases exhibits different


degrees of reactivity towards the reactant ions, positive or negative 25 ,34.
This contributes to the baselines and results in discrete peaks of ions that
can show different reactivity toward the sample molecules compared with the
original reactant ions.
287

Dexsil410 a
tO yDexsil300
c It. OY-210
0.9 OY-101 ......; UCW-98
08 -30 -410 SP-525
---- ... ...-
0.7
06 QY-17 .
05 HI-EFF 2GP(EGP)
04
0.3
0.2 OY-25
Carbowax E-20M
0.1

40 50 60
Time(min)

b
Carbowax E-20M
Carbowax E-4ooo
....-___~~,..a:::=--.-6- OY-7

RL--:,=--~~~~~~------;.,-
20 30 40 50 60
-- _J.
70
~-;!;;-'------;!',
80 90
Time(min)

Time!mini

Fig. 12.10. Relative conditioning rates for GC stationary phase in plasma chro-
matograph. c, Normalized reactant ion concentration. (a) Positive mode in air;
(b) positive mode in nitrogen; (c) negative mode in air. Temperature, 200 0 C.
DEGS = Diethylene glycol succinate. EGP = Ethylene glycol phthalate. (Reprinted
from ref. 34, with permission.)
288

Fig. 12.10 shows plots of the fractional return of the reactant ion signal
to its initial intensity as a function of time following the introduction of
0.1 mg of stationary phase into the sampling port of the plasma chromatograph
for positive and negative modes 34 • For the positive mode with air as carrier
and drift gas, the plasma chromatograph gives little response to Dexsil 410,
OV-210, Dexsil 300 and OV-l0l. These stationary phases, together with DC 410.
UCW-98, SE-30 and OV-7, if thermally conditioned to remove volatiles, should
be acceptable for plasma chromatography (the faster the return of the reactant
ion concentration, the less is the interaction of the plasma chromatograph with
the bleed from the stationary phase). With nitrogen as carrier and drift gas,
Carbowax E-20M is also acceptable. A comparison of the relationships in the
positive and negative modes for OV-210 and SP-525 clearly indicates differences
in reactivity for volatiles towards the corresponding reactant ions. The highly
electronegative trifluoropropyl group in OV-210 shifts this phase to the right
in the negative mode with air. However, the curve for SP-525 shifts to the left,
which demonstrates its applicability to plasma chromatography in the negative
mode.
The introduction of a dopant compound (hexane or tetrachloromethane) into
the detector usually causes a decrease in the detection sensitivity in both the
selective positive and selective negative mode 35 . At a hexane mass rate of
3.10- 7 g/sec this decrease amounts to ca. 50% in the positive mode with
naphthalene. In the negative mode the decrease is ca. 66% at a mass rate of
8.10- 10 g/sec of carbon tetrachloride with hexachloroethane.

REFERENCES

1 F.W. Karasek, Res./DeveZop., 21, March (1970) 34.


2 M.J. Cohen and F.W. Karasek, J. Chromatogr. Sei., 8 (1970) 330.
3 F.W. Karasek, Res./DeveZop., 21, (1970) 25.
4 F.W. Karasek, AnaZ. Chem., 46 (1974) 710A.
5 F.W. Karasek and D.W. Denney, AnaZ. Chem., 46 (1974) 633.
6 0.1. Carroll, R.N. Dzidic, R.N. Stilwell and E.C. Horning, AnaZ. Chem., 47
(1975) 1956. .
7 M.A. Baim and H.H. Hill, Jr., AnaL. Chern., 54 (1982) 38.
8 LA. Mason and H.W. Schamp, Jr., Ann. Phys. (N.Y.), 4 (1958) 233.
9 F.W. Karasek, 0.5. Tatone and D.W. Denney, J. Chromatogr., 87 (1973) 137.
10 F.W. Karasek, D.M. Kane and 0.5. Tatone, AnaZ. Chem., 45 (1973) 1210.
11 F.W. Karasek and D.M. Kane, AnaZ. Chem., 46 (1974) 780.
12 F.W. Karasek and D.W. Denney, J. Chromatog~., 93 ,tl9]4) 141.
13 F.W. Karasek, D.W. Denney and LH. DeDecker, Anaz'\ Ch~m., 46 (1974) 970.
14 F.W. Karasek and S.H. Kim, J. Chromatogr., 99 (1974) ~:
15 F.W. Karasek and S.H. Kim, AnaZ. Chem., 47 (1975) 1166.
16 F.W. Karasek, D.E. Karasek and S.H. Kim, J. Chromatogr., 105 (1975) 345.
17 F.W. Karasek, A. Maican and 0.5. Tatone, J. Chromatogr., 110 (1975) 295.
18 F.W. Karasek, H.H. Hill, Jr. and S.H. Kim, J. Chromatogr., 117 (1976) 327.
19 F.W. Karasek and D.M. Kane, J. Chromatogr. Sei., 10 (1972) 673.
20 F.W. Karasek and D.W. Denney, AnaZ. Chem., 46 (1974) 1312.
289

21 D.S. Ithakissios, J. Chromatogr. Sci., 18 (1980) 88.


22 F.W. Karasek, Anal. Chem., 43 (1971) 1982.
23 F.W. Karasek and R.A. Keller, J. Chromatogr. Sci., 10 (1972) 626.
24 S.P. Cram and S.N. Chesler, J. Chromatogr. Sci., 11 (1973) 391.
25 J.C. Tou and G.U. Boggs, Anal. Chem., 48 (1976) 1351.
26 F.W. Karasek, H.H. Hill, Jr., S.H. Kim and S. Rokushika, J. Chromatogr., 135
(1977) 329.
27 M.A. Bairn and H.H. Hill, Jr., J. High Resolut. Chromatogr. Chromatogr.
Commun., 6 (1983) 4.
28 M.A. Bairn and H.H. Hill, Jr., J. Chromatogr., 279 (1983) 631.
29 M.A. Bairn, R.L. Eatherton and H.H. Hill, Jr., Anal. Chern., 55 (1983) 1761.
30 D.M. Lubman and M.N. Kronick, Anal. Chern., 54 (1982) 1546.
31 D.M. Lubman and M.N. Kronick, Anal. Chern., 55 (1983) 867.
32 D.F. Hagen, Anal. Chem., 51 (1979) 870.
33 T.W. Carr, J. Chromatogr. Sci., 15 (1977) 85.
34 T. Ramstad, T.J. Mestrick and J.C. Tou, J. Chromatogr. Sci., 16 (1978) 240.
35 M.A. Bairn and H.H. Hill, Jr., J. Chromatogr., 299 (1984) 309.
This Page Intentionally Left Blank
291

Chapter 13

t4ISCELLANEOUS DETECTORS

CONTENTS

13.1. Introduction . . • . • • . • . 291


13.2. Plasma-emission spectrometry. 291
13.3. Atomic-absorption spectrometry 294
13.4. Ion-selective electrodes ••• 294
13.5. Piezoelectric sorption detector 295
13.6. Mass and infrared spectrometry . . • • . . . . . • • • • • . 296
13.6.1. Interfacing gas chromatography and mass spectrometry 297
13.6.2. Mass spectrometer as a GC detector 297
13.6.3. Methods of ion production 298
13.6.4. Infrared spectrometer 300
References . . . . . . . . . • . . . . . 305

13.1. INTRODUCTION

The previous chapters described selective detectors that are commonly used
in gas chromatography (GC) and manufactured commercially. In addition to these
detectors, many other physico-chemical principles can be applied for selective
detection, e.g., polarographyl, nuclear magnetic resonance 2 ,3, atomic-fluo-
rescence spectrometry4, fluorescence spectrometry5-15, ultraviolet spectro-
PhotometryI6-21, elemental analysis 22 ,23, and laser absorption 24 . However,
these detectors have so far been used for research purposes. Detectors that
have found wider use, e.g., atomic-absorption spectrometers, ion selective
electrodes, piezoelectric detector and, above all, plasma emission spectrometers,
are briefly described in this chapter, which also contains concise data on
infrared and mass spectrometry.

13.2. PLASMA-EMISSION SPECTROMETRY

Detection by means of emission spectrometry is based on the same principle


as in flame photometric detection, but it is a plasma that serves as the
emission source in this instance.
The plasma consists of a mass of predominantly ionized gas at a temperature
of 4000-10 000 K. This state can be maintained by an electrical discharge
through the gas [d.c. plasma (OCP)] or indirectly via inductive heating of the
292

Argon-Hydrogen

Gas Supply

Microwove Stobilized

IOscilloscope Generator
Power
Supply

Strip Chart
Recorder

Fig. 13.1. Schematic diagram of the GC-MIP system. 1 = Microwave cavity and
plasma; 2 = quartz lens; 3 = sample injection port; 4 = reflected power meter;
5 = diffraction grating; 6 = photomultiplier power supply; 7 = vibrating plate.
(Reprinted with permission from ref. 52.)

gas by means of an electromagnetic field established using, power generated at


radiofrequencies [inductively coupled plasma (rCP)] or microwave frequencies
[microwave-induced plasma (MIP)J. The analyte,excitation results from electron
impact and from collision with metastable atoms of the plasma support gas
(usually argon or helium). Emissions from these species, selected by means of
a monochromator, are used to detect the presence of the compound in the effluent
rare gas stream25 - 79 •
In order to achieve selective detection, it is necessary to form predominantly
atomic rather than molecular species, as the latter often overlap the major
atomic spectral lines. Most of the literature is concerned with MIP emission
detectors. A schematic diagram of a GC-MIP interface is given in Fig. 13.1.
The GC-MIP combination has been used for the determination of phosphorus 26 ,28,
31,38,54,68, halogens25,27-31 ,33,37,38,54,68,71 ,76,77, nitrogen 35 ,39, chromi-
um41 ,47,63,68, 1ead 44 ,52 ,54 ,67 ,68 ,72,74, mercury36 ,43 ,46,54,61,68,75, s il icon 48 ,
54,68, selenium42 ,68, arsenic44 ,45,68 , manganese54 ,68 , boron 68 ,73, sulphur 25 ,29,
32,35,38,39,54,62,68,70, iron 63 ,68, tin68,72, molYbdenum68 , vanadium68 , ger-
manium68 ,72, niobium68 , tungsten 68 , ruthenium68 , osmium68 , cobalt68 , nicke1 68 ,
aluminium68 , gallium40 compounds, including metal chelates, and hydrogen iso-
topes39,51 ,68. Examples of the sensitivity, linear dynamic range and selectivity
for individual elements are given in Table 13.1. Of course, the detector can be
made non-selective by observing C lines25,35,38,39,61 ,63,67,68,76 . If a poly-
chromator/microcomputer system is used, the simultaneous monitoring of multiple
atomic-emission wavelengths can be carried out through an entire chromatographic
run 75 •
293

TABLE 13.1

DETECTION LIMITS, SELECTIVITIES AND LINEAR DYNAMIC RANGES

Reprinted with permission from ref. 68.

Element Emission Detection Minimum Selectivity Linear


wavelength 1imi t detect- dynamic
(nm) (pg) abil ity range
(pg/sec)

H 486.1 45 16 74 5-10 2
H 656.3 22 7.5 160 5-10 2
2H( I) 656.1 20 7.4 194 5-10 2
V(I I) 268.8 26 10 5.69 -1 04 10 2
Nb(I1) 288.3 335 69 3.21-10 4 10 2
Cr(I1) 267.7 19 7.5 1.08-10 5 10 3
Mo(I1) 281.6 25 5.5 2.42-10 4 5-~02
W(II) 255.5 646 51 5.45 -10 3 10
Mn(I I) 257.6 7.7 1.6 1.11-10 5 10 3
Fe(I!) 259.9 0.89 0.28 2.80-10 5 10 3
Ru (I I) 240.3 35 7.8 1.34-10 5 10 3
Os (I I) 225.6 34 6.3 5.00-10 4 10 3
Coli) 240.7 18 6.2 1.82'10 5 5-10 2
Ni (I 1) 231.6 5.9 2.6 6.47-10 3 10 3
Hg(I) 253.7 60 0.60 7.69-10 4 10 3
B(I) 249.8 27 3.6 9.25-10 3 5-10 2
Al (I) 396.2 19 5.0 3.90-10 3 5 -10 2
C(I) 247.9 12 2.7 1.00 10 3
Si (I) 251.6 18 9.3 1.58-10 3 5-10 2
Ge(I) 265.1 3.9 1.3 7.57 -10 4 10 3
Sn( I) 284.0 6.1 1.6 3.58-10 5 10 3
Pb(I) 283.3 0.71 0.17 2.46-10 5 10 3
Pb(I) 405.8 7.2 2.3 2.00-10 5 10 3
P(I) 253.6 56 3.3 1.06 -10 4 5-10 2
As(I) 228.8 155 6.5 4.70-10 4 5-10 2
S(I I) 545.4 76
Se(I ) 204.0 62 5.3 1.09-10 4 10 3
F( I) 685.6 64 20 573 5-10 2
C1{II) 479.5 155 43 610 5-10 2
Br (I 1) 470.5 106 33 274 5-10 2
Br(I I) 478.6 106 34 599 5-10 2
I{I) 206.2 56 21 5.01-10 4 5-10 2

. 56 57 .. 55-57 55-57
The GC-ICP system has been used for t1n ' , slllcon , lead ,
.1ron 56,57. G6 d 65 compounds. The GC-DCP system has been used
, n1trogen an oxygen
f or c hrom1um
. 50 , copper 50. d· 53 , 1ea d60 , t1n
, me ke153 , pa 11 alum . 60 , Sl·1 lcon
. 60 ,
60 62
germanium , sulphur and manganese compounds • 49
294

13.3. ATOMIC-ABSORPTION SPECTROMETRY

The interfacing of an atomic-absorption spectrometer (AAS) to a gas chro-


matograph is accomplished 58 ,80 through the atomizer (direct connection of the
column to the burner gas flow, connection of a heated transfer line from the
column to the injection port of the electrothermal device, or cold vapour tube
atomizer interface for mercury compounds). Thus, both flame 81 - 84 and flame-
less 85 - 91 atomization are used. However, the minimum detectability with the
flameless mode is two to three orders of magnitude 10wer80
The AAS is particularly attractive for. detection of organometallic compounds.
This detector has been used for lead81 ,83-85,87 (217 nm), tin 88 ,90,91 (224.6 nm),
mercury92-94 (253.7 nm), chromium95 ,96 , arsenic 88 (193.7), silicon 82 (251.6 nm)
and selenium86 ,88 (196 nm), the detection limits being 0.1,0.1,0.02, 1,5,
100, and 1 ng for Pb 87 , Sn 90 , Hg 93 , Cr 95 , As 88 , Si 82 and Se 86 compounds,
respectively. The linear dynamic range covers about four orders of magnitude 91

13.4. ION-SELECTIVE ELECTRODES

The gas chromatographic effluent is drawn through a reaction tube at 800-


1000 0C. The separated components undergo hydrogenolysis in the presence of a
catalyst, forming hydrogen sulphide, hydrogen chloride, hydrogen fluoride and
ammonia. The gases are dissolved in a suitable absorption solution and the
concentrations of the ions produced are monitored continuously in a flow-through
a cell with an appropriate ion-selective electrode (Fig. 13.2). Non-absorbed

:-----l
I
I
I I
L ____ --.J

Fig. 13.2. Schematic diagram of the combination of gas chromatograph, reaction


tube and ion-selective electrode. 1 = Gas chromatograph; 2 = reaction tube;
3 = tube consisting of a gas-liquid contact area and a gas-liquid separation
area; 4 = absorption solution; 5 = micropump; 6 = stream buffer; 7 = ion-
selective electrode; 8 = reference electrode. (Reprinted with permission from
ref. 99.)
295

gases, such as methane, are separated together with the carrier gas from the
absorption solution in a gas-liquid separator and vented from the system.
Selective responses are obtained for sulphur 97 ,98, chlorine 97 , fluorine 99 ,
101
bromine 100 and nitrogen compounds • A detector equipped with two ion-selective
electrodes allows the simultaneous determination of two types of compound
through dual-channel operation (chlorine/fluorine-containing compounds 102 or
bromine/fluorine-containing compounds 103 ). The products of hydrogenolysis can
also be split into two parts. One part passes through a granular silver absorber
into a flame-ionization detector and the other passes through a dual ion-
selective electrode cell (chloride and bromine). Thus, the atomic ratios of
chlorine, bromine and carbon in halogen-containing organic compounds can be
determined 104 in this way. Mercaptans can be detected selectively without
hydrogenolysis by using a silver/sulphide ion-selective electrode 105 • The de-
tection limits are 10- 12 mole of sulphur compounds 103 , 10- 11 mole of fluorine
compounds, 10- 10 mole of chlorine compounds and 5.10- 11 mole of bromine com-
pounds 100 • The linear dynamic range is about 10 4 and the selectivity for sulphur
compounds 97 and the N:Cl response ratio 101 are 2.10 3• The response time of the
detector is high 99 •

13.5. PIEZOELECTRIC SORPTION DETECTOR

A quartz crystal vibrating at a constant frequency in the megacycle range


exhibits a decrease in frequency when substances are adsorbed directly on the
surface of the crystal coated with a thin film of a liquid or solid sorbent.
When used as a detector for gas chromatography 106-116 , the response 1S
. 0 bta1ne
. d
by coating the crystal with a film of the same liquid phase as that commonly
used as the stationary phase for the column. The eluted compounds passing over
the quartz crystal surface dissolve in the coating, thus changing the resonant
frequency of the oscillating piezoelectric crystal. The detector response is
given by the equation

( 13.1)

where w is the total weight of the eluent, y is the activity coefficient of


the eluent in the crystal coating, po is the vapour pressure of the eluent at
the operating temperature, F is the carrier gas flow-rate and c is a constant
that is characteristic of the detector temperature, the crystal and the liquid
phase used to coat the crystal. Thus, the detector response increases with
increasing molecular weight and boiling point of the eluting compounds, making
the wide, low peaks at the end of the chromatogram more detectable.
296

THERMAL DETECTOR

Fig. 13.3. Chromatograms from differential piezoelectric sorption detector and


thermal conductivity detector. (From ref. 107.)

The detector can be made selective by the choice of the crystal coating,
e.g., Sl"1 lca
" ge 1 , mo 1ecu 1ar Sleve
" 106 or a hygroscoplC
"1po ymer 107 f or th e de t er-
mination of water, lead acetate for the determination of hydrogen sulphide 106
and inorganic salts 111 for organophosphates. It is obvious that the detector
selectivity is limited by being given by the relative sorption properties of
the system only. The chromatogram in Fig. 13.3 shows the results obtained with
a differential piezoelectric sorption detector. The response of the polar
detector (dinonyl phthalate coating) is made equal to that of the non-polar
detector (DC 200 silicone oil) for alkanes. The response is obtained for polar
compounds only. The detection limit 106 ,114 is about 10- 9 g and the linear
dynamic range 107 is 10 4 .

13.6. MASS AND INFRARED SPECTROMETRY

As already stated in Chapter 1, a great advantage of GC is its high separa-


tion efficiency. This means that GC is able to separate mUlti-component mixtures
into single components. However, the identification ability of GC does not
match the level of its separation ability. On the other hand, mass spectrometry
(MS) is an analytical technique with a high identification ability. By combining
these two techniques, we obtain an almost ideal analytical unit where MS
operates as a chromatographic detector. Much time has passed since combined
GC-MS was used for the first time 117 , and nowadays it is a highly developed
independent technique. There is a large literature on this subject and to deal
with it in detail would require a separate book. Therefore, only the principles
are briefly described here and readers are referred to the numerous books
available, e.g., refs. 118-128.
297

13.6.1. Interfacing gas chromatography and mass spectrometry

It has been stated that "The development of a gas chromatographic-mass


spectrometric analysis system is not a mere combining of the two techniques.
There are inherent incompatibilities of operational procedures when the
techniques are performed separately that must be resolved when they are com-
bined"121,129. The GC-MS interface has frequently been a point of difficulty
in combining these techniques. The difficulty arises from the fact that a mass
spectrometer must operate with sufficient vacuum for the ions to traverse the
analyser without collisions (10- 4-10- 5 Torr), whereas a typical gas chroma-
graph operates with its exit at atmospheric pressure.
With packed columns it is mostly necessary to use so-called molecular
separators as interfaces between the GC and MS instruments 125 ,127,128,130-132.
Thelr task is to remove as many molecules of the carrier gas as possible from
the column effluent and to transport the maximum amount of the organic solute
into the mass spectrometer ion source. These two functions are measured by
an enrichment factor N, indicating the amount of the carrier gas removed from
the GC peak, and by the yield Y, indicating the percentage of the sample that
actually reaches the MS ion source. The jet separator133-136 is a based on the
fractionation of gases in an expanding jet stream. Effusion-type separa-
tors137-142 are based on selective effusion through fine pores or through a
narrow slit. In membrane separators, the preferential diffusion of the carrier
gas or of the sample takes place through a semi-permeable membrane (PTFE
separator 137 ,143, palladium-silver separator144-146 or silicone separator147-150).
GC-MS interfacing without a molecular separator can be accomplished either
as an open split coupling or as a direct coupling. The former seems to be the
preferred interfacing technique at present151-153. Direct coupling can be used
for a well pumped chemical ionization system and also for packed columns,
whereas capillary columns can be connected directly to the most modern GC-MS
systems.

13.6.2. Mass spectrometer as a GC detector

There are two ways in which chromatograms can be obtained from the output
of a mass spectrometer: by recording either the total ion current or the ion
current of a selected ion mass.
A general chromatographic record, similar to those provided by other non-
selective GC detectors, can be obtained by measuring the total ion current (TIC).
The TIC is formed from the solute molecules eluting from the gas chromatograph
and is recorded as a function of time. In sel'ected ion monitoring (SIM), the
298

intensities of pre-selected ions, characteristic of a class or of a particular


compound, are recorded as a function of time, SIM is made possible by rapid
switching from one mass to another in a very short time. Chromatographic peaks
appear only if the compound producing the chosen ion is present. In this mode
the mass spectrometer operates as a classical selective detector responding
only to a certain type of compound. The sensitivity that can be obtained in
the single ion mode is up to two orders of magnitude higher than that with
the TIC (the ion current is integrated for a longer time), e.g., the detection
limit for steroids 154 (mass 436.319) is 2.10- 14 g, for benzophenone 155 with
chemical ionization (mass 183) 1.10- 14 g and for dopamine derivatives 156 with
negative ion chemical ionization 1.10- 14 g.
It is possible to obtain a mass spectrum for every chromatographic peak in
real time. Hence, in this mode, a mass spectrometer is a highly specific GC
detector. The spectra produced are compared with those of known compounds by
a computer system. The information obtained indicates what compound has eluted
in an observed peak.
The chromatograms produced by a GC-MS system may either be generated in
real time or reconstructed-by a computer. With computer control, the spectra
can be scanned and acquired repetitively every few seconds during the chroma-
tographic run. In this way, three-dimensional ion signals are generated, each
characterized by mass, intensity and time. All the spectra are stored in the
computer. The total ion chromatogram can be reconstructed by summing the ion
intensities obtained in each repetitive scan. By retrieVing the intensities
of the chosen ion from each scan and by plotting them as a function of time,
mass fragmentograms (selected ion chromatograms) can be reconstructed. Fourier
transform Ms 157 ,158 and MS_MS 161 can also be used in connection with gas chro-
matography150,160.

13.6.3. Methods of ion produation

The following methods of ion production are mostly used in GC-MS: electron-
impact ionization (EI), chemical ionization (CI), field ionization (FI), and
atmospheric pressure ionization (API)123,128,162.
EI is the commonest method. Electrons of 70-eV energy are produced in a
collimated stream by thermionic emission from a filament. This electron beam
falls on neutral molecules entering the ion source, and the energy transmitted
brings about ionization and fragmentation of the sample molecules. The molec-
ular ionization is initiated when the electron energy is greater than the
ionization potential of the compound. A disadvantage of this mode of ionization
299

is that in most instances very complicated spectra arise in which the molecular
ion is mostly missing. For structural analyses this is, however, an advantage.
Methane and isobutane are the most often used reactant gases in CI_MS163-167
The CI mass spectra often show abundant characteristic protonated molecular
ions, even when the corresponding EI mass spectra show no detectable molecular
ions. The primary ions produced in CI by the EI of the reactant gas are, in
comparison with the analyte, in a large excess at pressures of about 1 Torr.
The ionized reactant gas undergoes ion-molecule self-reactions to form a steady-
state plasma. The ions produced react by proton and hydride transfer to give
M+1 and M-1 ions that may further dissociate. The CI mass spectrum of the
sample is dependent on the ions of the reactant gas. The ion molecular reaction
is a much milder process than EI. CI mass spectra are less complex and, there-
fore, easier to interpret than El mass spectra. Negative CI gives a lower
detection limit for certain types of compounds, because the negative quasi-
molecular ion spectra of some compounds are one to three orders of magnitude
more intense than the positive ion spectra1~2. It is possible to obtain both
the positive and negative spectra at the same time by pulsing the polarity of
the ion source potential and focusing the lens potentia1 168 .
In the FI source 169 ,170, soft ionization is promoted by an extremely high
potential gradient (about 10 7-10 8 V cm- 1). The molecules passing through this
gradient will modify the field and allow an electron to escape from the molecule,
thus producing an ion with a small excess of energy available for fragmentation.
i~ass spectra are generally characterized by the presence of prominent ion or

"parent ion" peaks with only a few minor fragment ion peaks. FI is five to ten
times less sensitive than EI {ca. 50 nm)171.
In the API source ion-molecule reactions occur in an inert gas weak plasma
at a pressure of 1 atm '56 ,172,173. The source can work in both positive and
negative ion modes.
Two mass detectors, small compact units, are currently commercially available.
The ions of a certain mass can be stored in stable paths for many seconds in an
ion trap174,175. A GC detector 176 , the so-called ion trap detector 21 ,177 (ITO)
(Finnigan-MAT), has been constructed on this principle. The ion trap is scanned
over a mass range whereby the ions are ejected from the ion storage region
sequentially from low to high mass. The ejected ions are detected by a conven-
tional electron multiplier. The ITO also gives a universal gas chromatogram
(the instrument is scanned repetitively over a user-selected mass range).
complete mass spectra of all the eluting compounds and selected ion detection
(up to sixteen pre-selected ions during each scan). The Hewlett-Packard mass-
selective detector 178 ,179 with a hyperbolic quadrupole analyser can also operate
in the TIC mode so that a spectrum of any peak can be obtained, or in the SIM
mode with twenty channels.
300

13.6.4. Infrared spectrometer

In the last few years interest has focused on the combination of GC with
infrared (IR) spectrometry180,222, particularly with the use of capillary
columns. The wider application of modern GC-IR for analyses of multi-component
mixtures has resulted especially from: (1) the development of Fourier transform
(FT) IR spectrometry, (2) the introduction of a narrow-range mercury-cadmium
telluride photodetector and (3) the construction of small, gold-coated glass
"light pipes,,193. The path length is very long, which results in a relatively
low detection limit. At present, the detection limit 182 ,184,191 ,202,215 ranges
from 10 to 100 ng. A schematic diagram of the GC-FTIR system is shown in Fig.
13.4.
There are two ways to obtain chromatograms from interferometric data. (1) In
the Gram-Schmidt reconstruction each interferogram is treated as a vector and
the orthogonal components of the chromatogram vectors are computed with respect
to a set of background basis vectors representative of the GC baseline. A plot
of the length of the orthogonal component against the interferogram number
will form a reconstructed chromatogram for a particular set of interferograms 181.
Specific functional groups cannot be monitored. (2) 512 (or 1024) point sec-
tions of each interferogram are transformed to obtain a low-resolution IR
absorbance spectrum, which can then be integrated to determine a chromatogram

IR Source

Interferometer

/ MGT Detector
/
/
/
Paraboloid

Inlet

Gas Chromatograph

Fig. 13.4. Schematic diagram of GC-FTIR instrumentation. MCT = mercury-cadmium


telluride photodetector. (Reprinted with permission from ref. 212.)
301

detector response. The total absorbance is calculated for various bands or


windows in the spectrum and plotted atainst the interferogram number. The so-
called chemigram l85 ,21o is the plot of the integrated absorbances within five
user-selected spectral windows as a function of elution time. Therefore,
selective detection (analogous to SIM in GC-MS) is possible 21o (with the earlier
IR spectrometers a desired wavelength could also be selected by a filter and
thereby a selective chromatogram obtained I8o ,195,211). The reconstructed
chromatogra~ can be compared to total ion monitoring in GC-MS. The peak heights

in TIC chromatograms are often similar to those of the peaks obtained from
an FLO, but this is not the case when comparing·the IR reconstructed chromato-
gram and the FLO chromatogram, as the IR molar absorptivities of bands of
polar molecules tend to be much greater than those of non-polar molecules (see
Fig. 13.5).

Scan
Sets 91 182 279 362 453 543 634 724 815 905 996 1086
i I I I I I I I 1 I 1 I 1 -,
ClOO t23 2.47 170 4.94 6.17 7.41 8.64 9.88 11.11 1235 1358 14.81 11i05
Time (min)

Fig. 13.5. Comparison of (a) the chromatogram measured o~ an FLO located ~fter
the light pipe with (b) IR reconstructed chromatogram uSlng the Gram-Schmldt
algorithm. (Reprinted with permission from ref. 212.)
302

A very successful search algorithm for GC-FTIR spectra involves first


normalizing the absorbance of each measured spectrum, so that the most intense
band has an absorbance of unity. Each spectrum in the reference library is
normalized in the same fashion. The sample spectrum is then subtracted from
each reference spectrum, and the sum of the squares of each point in the
differential spectrum is computed. This sum is sometimes known as the "hit
index"; the reference spectrum giving the smallest hit index has the greatest
212
probability of yielding the identity of the component of interest •
A subnanogram detection limit, i.e •• a limit 100 times lower than that
obtained with a commercial light pipe GC-IR, is reported for the use of
matrix isolation 223 . The combined use of gas chromatography and matrix iso-
lation IR spectrometry was first demonstrated in 1979 224 , 225. A commercial
unit, Cryolect. has been developed 226 • Each molecule of the GC effluent is
captured into an argon cage at 12 ~ (up to 5 h of chromatography can be
frozen indefinitely). Thus each molecule is isolated from the others and
intermolecular interactions that give rise to band broadening are eliminated.
The Cryolect unit is schematically presented in Fig. 13.6. The GC effluent
is directed onto a gold-plated copper disk. After spraying helium. argon and
the separated GC components onto the disk surface, helium is removed by the
vacuum system. Argon and the GC components condense in a solid spiral band.

VACUUM CHAMBER
r="";;::::::::::;;"-=
IR DETECTOR

FOCUSING MIRRORS

OPTIC_'_": yftiii==fI!LA GAS CHROMATOGRAPH


GC OVEN

COllECTION DISK _-..+-;

-
~ ......
A~ •

A
''''
"'ao:
z~
... ~
1'rtTr======:::;rI~
~,
FLAME IONIZATION
0
~
...
...

o DETECTOR

Fig. 13.6. Schematic diagram of Cryolect GC and collection chambers. (From


ref. 226.)
303

The optical beam from the spectrometer passes through the sample matrix, is
reflected by the gold disk, passes back through the sample to the second mir-
ror and passes into an IR detector. All the IR measurements are performed af-
ter the chromatograms have been deposited 223 ,226
The complementary nature of the GC-MS and GC-FTIR techniques is well recog-
nized 189 ,197,200,203,219. Both mass and IR spectrometry suffer from certain
weaknesses in identifying compounds: the former in distinguishing chemical
isomers and the latter in distinguishing long-chain homologues, as illustrated
in Figs. 13.7 and 13.8, demonstrating the mass and IR spectra of tetrachloro-
benzene isomers and n-alkanes, respectively. The mass spectra of the tetra-
chlorobenzene isomers are identical whereas the IR spectra yield data that
can he uspd for identification. In contrast, for n-alkanes the IR spectra are

B
1.2000
a

~ -0.1199 J<.. lit L\


I- 1.2000
c{
:J
...
Z
l- b
I-
<l
...u

~
A z
c{

,~
a III
1049] IX
0
~ -0.1t..00
...u
z j" i •
• i i - c{
tOOO

f49] ... ~
b c

~ i
M
i
L
>

r"] i
. .•
c

, i ~ - 0.1399
3200 2t..00 1600 800
M/E 50 100 150 200 250 300
WAVENUMBERS

Fig. 13.7. Electron-impact mass spectra (A) and vapour-phase IR spectra (B) of
the isomeric tetrachlorobenzenes: a = 1,2,3,5-tetrachlorobenzene; b = 1,2,3,4-
tetrachlorobenzene; c = 1,2,4,5-tetrachlorobenzene. (Reprinted with permission
from ref. 204.)
304

43 n-PENTANE
100 "-PENTANE

80

57

60 80 100 3200 2400 1600 800


m/z WAVENUMBERS
57
100 "-HEXANE "-HEXANE

80
>
I-
~ 60 43
UJ
UCD
zeJ!
w
I-
«0
aI"",
! ~r-:
~ 40 VIC>
i= aiel)
~ «~
~

~ 20 86
('I

~
~~~~~~~~~==r=~~~~~
'4000 3200 2400 1600 800
60 80 100 WAV£NUMBERS
m/z
43
100 "-HEPTANE "-HEPTANE

80
UJ
u~

z51
71 «
57 ~~
Od
VI
aI(II
«~
d

100

85 3200 2400
WAVE NUMBERS
O~~. .~~~~-+~
40 60 80 100
m/z

Fig. 13.8. Low-resolution GC-mass spectra and gas-phase FTIR spectra (at 4 cm- 1
resolution) of pentane, hexane and heptane. (Reprinted with permission from
ref. 196.)
305

Needl e Digilab
...----1 FTS-20C .....- . - o j Nova 3
FTIR Computer

HP
5985 B
GC/MS

GC oven

HP21MX
Computer

Fig. 13.9. Schematic diagram of GC-FTIR-MS system. HP = Hewlett-Packard; SCOT =


support-coated open tubular. (Reprinted with permission from ref. 203. Credit
is given to the University of California, Lawrence Livermore National Laboratory,
and to the Department of Energy under whose auspices the work was performed.)

identical whereas the MS spectra differ. Hence it is obvious that the use of
the two spectral methods combined with gas chromatography offers the optimum
solution to the analysis of complex mixtures and the identification of indi-
vidual compounds. There are two approaches to this problem: either sample
analysis separately with both instrument combinations using the same chromato-
graphic column, or tandem connection of the two spectral devices at the column
outlet, i.e., GC-FTIR-MS. The latter solution is more advantageous, of course.
The latest publications 196 ,203,207,220 dealing with this problem are based on
the tandem arrangement (see Fig. 13.9).

REFERENCES

1 B. Fleer and T.H. Risby, Talanta, 16 (1969) 839.


2 J. Buddrus and H. Herzog, Org. Magn. Reson., 15 (1981) 211.
3 H. Herzog and J. Buddrus, Chromatographia, 18 (1984) 31.
4 J.C. Van Loon, J. Lichwa and B. Radziuk, J. Chromatogl'., 136 (1977) 301.
5 M.C. Bowman and M. Beroza, Anal. Chern., 40 (1968) 535.
6 H.P. Burchfield, R.J. Wheeler and J.B. Bernos, Anal. Chern., 43 (1971) 1976.
7 D.J. Freed and L.R. Faulkner, Anal. Chern., 44 (1972) 1194.
8 J.W. Robinson and J.P. Goodbread, Anal. Chim. Acta, 66 (1973) 239.
9 H.P. Burchfield, E.E. Green, R.J. Wheeler and S.M. Billedeau, J. Chl'omatogl'.,
99 (1974) 697.
10 J. Mulik, M. Cooke, M.F. Guyer, G.M. Semeniuk and E. Sawicki, Anal. Lett.,
8 (1975) 511.
11 R.P. Cooney and J.D. Winefordner, Anal. Chern., 49 (1977) 1057.
12 F. Chow and A. Karmen, Clin. Chern., 26 (1980) 1480.
306

13 L.C. Thomas and A.K. Adams, Ana~. Chem., 54 (1982) 2597.


14 J.M. Hayes and G.J. Small, Ana~. Chem., 54 (1982) 1204.
15 W.B. Conrad, W.J. Carter, E.L. Wehry and G. Mamantov, Ana~. Chem., 55 (1983)
1340.
16 W. Kaye, Ana~. Chem., 34 (1962) 287.
17 J. Merritt, F. Comendant, S.T. Abrams and V.N. Smith, Ana~. Chem., 35 (1963)
1461.
18 W. Kaye, Ana~. Chem., 36 (1964) 2380.
19 Spectroscopic Detectors for Gas Chromatography, L-513, Perkin-Elmer, Norwalk,
CT, 1977.
20 M. Novotny, F.J. Schwende, M.J. Hartigan and J.E. Purcell, Anal. Chem., 52
(1980) 736.
21 J.A. Borman, Ana~. Chem., 55 (1973) 726A.
22 S.A. Liebman, D.H. Ahlstrom, C.D. Nauman, R. Averitt, J.L. Walker and
E.J. Levy, in A. Zlatkis (Editor), Advances in Chromatography 1973, Chromato-
graphy Symposium, Houston, TX, 1973, p. 202.
23 V. Rezl and J. Bursa, J. Chromatogr., 126 (1976) 723.
24 J.D. Parli, D.W. Paul and R.B. Green, Ana~. Chem., 54 (1982) 1969.
25 A.J. McCormack, S.C. Tong and W.O. Cooke, Anal. Chem., 37 (1965) 1470.
26 C.A. Bache and D.J. Lisk, Ana~. Chern., 37 (1965) 1477.
27 C.A. Bache and D.J. Lisk, Anal. Chem., 38 (1966) 783.
28 C.A. Bache and D.J. Lisk, Ana~. Chem., 38 (1966) 1757.
29 C.A. Bache and D.J. Lisk, Anal. Chern., 39 (1967) 786.
30 H.A. Moye, Anal. Chem., 39 (1967) 1441.
31 ,C.A. Bache and D.J. Lisk, J. Ass. Offic. Ana~. Chern., 50 (1967) 1246.
32 C.A. Bache and D.J. Lisk, J. Gas Chromatogr., 6 (1968) 301.
33 C.A. Bache, L.E. St. John and D.J.,Lisk, Ana~. Chern., 40 (1968) 1241.
34 C.A. Bache and D.J. Lisk, Ana~. Chern., 40 (1968) 2224.
35 W. Braun, N.C. Peterson, A.M. Bass and M.J. Kurylo, J. Chromatogr., 55 (1971)
237.
36 C.A. Bache and D.J. Lisk, Ana~. Chern., 43 (1971) 950.
37 R.M. Dagnall, T.S. West and P. Whitehead, Ana~. Chim. Acta, 60 (1972) 25.
38 R.M. Dagnall, T.S. West and P. Whitehead, Ana~. Chem., 44 (1972) 2074.
39 W.R. McLean, D.L. Stanton and G.E. Panketh, Ana~yst (London), 98 (1973) 432.
40 R.M. Dagnall, T.S. West and P. Whitehead, Ana~yst (London), 98 (1973) 647:
41 F.A. Serravallo and T.H. Risby, J. Chromatogr. Sci., 12 (1974) 585.
42 Y. Talmi and A.W. Andren, Ana~. Chern., 46 (1974) 2122.
43 Y. Talmi, Ana~. Chirn. Acta, 74 (1975) 107.
44 Y. Talmi and V.E. Norvell, Ana~. Chem., 47 (1975) 1510.
45 Y. Talmi and D.T. Bostick, Ana~. Chern., 47 (1975) 2145.
46 Y. Talmi and V.E. Norvell, Ana~. Chirn. Acta, 85 (1976) 203.
47 M.S. Black and R.E. Sievers, Ana~. Chern., 48 (1976) 1872.
48 D.T. Bostick and Y. Talmi, J. Chromatogr. Sci., 15 (1977) 164.
49 P.C. Uden, R.M. Barnes and F.P. DiSanzo, Ana~. Chern., 50 (1978) 852.
50 K.S. Brenner, J. Chrornatogr., 167 (1978) 365.
51 F.P. Schwarz, Anal. Chern., 50 (1978) 1006.
52 D.C. Reamer, W.H. Zoller and T.C. O'Haver, Anal. Chem., 50 (1978) 1449.
53 R.J. Lloyd, R.M. Barnes, P.C. Uden and W.G. Elliot, Ana~. Chern., 50 (1978)
2025.
54 B.D. Quimby, P.C. Uden and R.M. Barnes, Ana~. Chern., 50 (1978) 2112.
55 D. Sommer and K. Ohls, Fresenius Z. Ana~. Chern., 295 (1979) 337.
56 D.L. Windsor and M.B. Denton, Anal. Chern., 51 (1979) 1116.
57 D.L. Windsor and M.B. Denton, J. Chrornatogr. Sci., 17 (1979) 492.
58 J.C. Van Loon, Ana~. Chem., 51 (1979) 1139A.
59 B.D. Quimby, M.F. Delaney, P.C. Uden and R.M. Barnes, Ana~. Chem., 52 (1980)
259.
60 S.A. Estes, C.A. Poirier, P.C. Uden and R.M. Barnes, J. Chromatogr., 196
(1980) 265.
61 S.P. Wasik and F.P. Schwarz, J. Chrornatogr. Sci., 18 (1980) 660.
62 D.S. Treybig and S.R. Ellebracht, Anal. Chem., 52 (1980) 1633.
307

63 S.A. Estes, P.C. Uden, M.D. Rausch and R.t1. Barnes, J. High Resolut.
Chrornatogp. Chpornatogp. Commun., 3 (1980) 471.
64 I.S. Krull and S. Jordan, Int. Lab., Nov./Dec. (1980) 13.
65 R.M. Brown, Jr., and R.C. Fry, Anal. Chern., 53 (1981) 532.
66 R.M. Brown, Jr., S.J. Northway and R.C. Fry, Anal. Chern., 53 (1981) 934.
67 S.A. Estes, P.C. Uden and R.M. Barnes, Anal. Chern., 53 (1981) 1336.
68 S.A. Estes, P.C. Uden and R.M. Barnes, Anal. Chern., 53 (1981) 1829.
69 M.A. Eckhoff, J.P. McCarthy and J.A. Carus, Anal. Chern., 54 (1982) 165.
70 J.L. Genna, W.O. McAninck and R.A. Reich, J. Chpornatogp., 238 (1982) 103.
71 K. Chiba, K. Yoshida, K. Tanabe, M. Ozaki, H. Haraguchi, J.D. Winefordner
and K. Fuwa, Anal. Chern., 54 (1982) 761.
72 S.A. Estes, P.C. Uden and R.M. Barnes, J. Chpornatogp., 239 (1982) 181.
73 I.S. Krull, S.W. Jordan, S. Kahland and S.B. Smith, Jr., J. Chpornatogp. Sci.,
20 (1982) 489.
74 S.A. Estes, P.C. Uden and R.M. Barnes, Anal. Chern., 54 (1982) 2402.
75 K. Chiba, K. Yoshida, K. Tanabe, H. Haraguchi and K. Fuwa, Anal. Chern., 55
(1983) 450.
76 M.A. Eckhoff, T.H. Ridgway and J.A. Caruso, Anal. Chern., 55 (1983) 1004.
77 K. Chiba and H. Haraguchi, Anal. Chern., 55 (1983) 1504.
78 L. Lendero, K. Cammann and K. Ballschmiter, Micpochirn. Acta, I (1984) 107.
79 D.S. Ballantine, Jr. and W.H. Zoller, Anal. Chern., 56 (1984) 1288.
80 M.J. Bychovskij and A.J. Braude, Zh. Anal. Chirn., 38 (1983) 2236.
81 B. Kolb, G. Kemmner, F.H. Schlesser and E. Wiedeking, Fpesenius Z. Anal.
Chern., 221 (1966) 166.
82 R.W. Morrow, J.A. Dean, W.O. Shults and M.R. Guerin, J. Chrornatogp. Sci.,
7 (1969) 572.
83 D.T. Coker, Anal. Chern., 47 (1975) 386.
84 Y.K. Chau, P.T.S. Wong and H. Saitoh, J. Chpornatogp. Sci., 14 (1976) 162.
85 D.A. Segar, Anal. Lett., 7 (1974) 89.
86 Y.K. Chau, P.T.S. Wong and P.O. Goulden, Anal. Chern., 47 (1975) 2279.
87 Y.K. Chau, P.T.S. Wong and P.O. Goulden, Anal. Chirn. Acta, 85 (1976) 421.
88 G.E. Parris, W.R. Blair and F.E. Brinckman, Anal. Chern., 49 (1977) 378.
89 J.W. Robinson, E.L. Kiesel, J.P. Goodbread, R. Bliss and R. Marshall, Anal.
Chirn. Acta, 92 (1977) 321.
90 Y.K. Chau, P.T.S. Wong and G.A. Bengert, Anal. Chern., 54 (1982) 246.
91 R.J. MaguireandR.,). Tkacz, J. ChPOrnatogp., 268 (1983) 99.
92 J.C. Gonzales and R.T. Ross, Anal. Lett., 5 (1972) 683.
93 J.E. Longbottom, Anal. Chern., 44 (1972) 1111.
94 R.C. Dressman, J. Chpornatogp. Sci., 10 (1972) 472.
95 W.R. Wolf, Anal. Chern., 48 (1976) 1717.
96 W.R. Wolf, J. Chpornatogp., 134 (1977) 159.
97 T. KOjima, M. Ichise and Y. Seo, Bunseki Kagaku (Jap. Anal.), 20 (1971) 20.
98 T. KOjima, T. Seo and J. Sato, Bunseki Kagaku (Jap. Anal.), 23 (1974) 1389.
99 T. Kojima, M. Ichise and Y. Sea, Talanta, 19 (1972) 539.
100 Y. Seo, Bunseki iagaku (Jap. Anal.), 28 (1979) 334.
101 T. Kojima, M. Ichise and Y. Sea, Bunseki Kagaku (Jap. Anal.), 22 (1973) 208.
102 T. Kojima, M. Ichise and Y. Seo, Bunseki Kagaku (Jap. Anal.), 24 (1975) 7.
103 T. KOjima, M. Ichise and Y. Sea, Anal. Chim. Acta, 101 (1978) 273.
104 Y. Seo, Bunseki iagaku (Jap. Anal.), 33 (1984) 252.
105 T. Kojima, Y. Seo and J. Sato, Bunseki Kagaku (Jap. Anal.), 24 (1975) 772.
106 W.A. King, Jr., Anal. Chern., 36 (1964) 1735.
107 W.H. King, Jr., Res. /Develop., 20, No.4 (1969) 28.
108 W.A. Ki"g, Jr., Envipon. Sci. Technol., 4 (1970) 1136.
109 G.G. Guilbault and A. Lopez-Roman, Envipon. Lett., 2 (1971) 35.
110 F.W. Karasek and K.R. Gibbins, J. Chpomatogp. Sci., 9 (1971) 535.
111 A. Lopez-Roman and G.G. Guilbault, Anal. Lett., 5 (1972) 225.
112 E.P. Scheide and G.G. Guilbault, Anal. Chern., 44 (1972) 1765.
113 M. Janghorbani and H. Freud, Anal. Chern., 45 (1973) 325.
114 F.W. Karasek and J.M. Tiernay, J. Chpornatogp., 89 (1974) 31.
115 F.W. Karasek, P. Guy, H.H. Hill, Jr. and J.M. Tiernay, J. Chpornatogp., 124
(1976) 179.
308

116 J. Hlavay and G.G. Guilbault, Anal. Chem., 49 (1977) 1890.


117 J.C. Holmes and F.A. Morrell, Appl. SpeatY'Osa .• 11 (1957) 86.
118 W.H. McFadden, Advan. Chromatogr., 4 (1967) 265.
119 J.T. Watson, in L.S. Ettre and W.H. McFadden (Editors), Anaillary Teahniques
of Gas Chromatography, Wiley, New York, 1969, p. 145.
120 S. Stallberg-Stenhagen and E. Stenhagen, in A.L. Burlingame (Editor),
Topias in Organia Mass Speatrometry, Wiley, New York, 1970, p. 167.
121 C. Merrit, Jr., in E.G. Brame, Jr. (Editor), Applied Speatrosaopy Reviews,
Vol. 3, Marcel Dekker, New York, 1970, p. 261.
122 G.A. Junk, Int. J. Mass Speatrom. Ion Phy"s., 8 (1972) 1.
123 W. McFadden, Teahniques of Combined Gas Chromatography/Mass Speatrometry:
Appliaations in Organia Analysis, Wiley, New York, 1973.
124 E.O. Oswald, P.W. Albro and J.D. McKinney, J. Chromatogr., 98 (1974) 363.
125 B.J. Gudzinowicz, r~.J. Gudzinowicz and H.F. Martin, Fundamentals of
Integrated GC-MS, Part III: The Integrated GC-MS Analytiaal System, Marcel
Dekker, New York, 1977.
126 C. Fenselau, Anal. Chem., 49 (1977) 563A.
127 M.C. ten Noever de Brauw, J. Chromatogr., 165 (1979) 207.
128 A.M. Greenway and C.F. Simpson, J. Ph~s. E, 13 (1980) 1131.
129 C. Merritt, Jr., in R.I. Reed (Editor), Recent Topias in Mass Spectrometry,
Gordon and Breach, New York, 1971.
130 D. Rees, Talanta, 16 (1969) 903.
131 A.N. Freedman, Anal. Chim. Aata, 59 (1972) 19.
132 W.H. McFadden, J. Chromatogr. Sai., 17 (1979) 2.
133 R. Ryhage, Anal. Chem., 36 (1964) 759.
134 E. Stenhagen, Fresenius Z. Anal. Chem., 205 (1964) 109.
135 R. Ryhage, S. Wikstrom and G.R. Waller, Anal. Chern., 47 (1965) 435.
136 R. Ryhage, Ark. Kem., 26 (1967) 305.
137 M.A. Grayson and C.J. Wolf, Anal. Chem., 39 (1967) 1438.
138 J.T. Watson and K. Biemann, Anal. Chem., 36 (1964) 1135.
139 J.T. Watson and K. Biemann, Anal. Chem., 37 (1965) 844.
140 M.C. ten Noever de Brauw and C. Brunnee, Fresenius Z. Anal. Chem., 229
(1967) 321.
141 A. Copet and J. Evans, Org. Mass Speatrom., 3 (1970) 1457.
142 M.A. Grayson and R.L. Levy, J. Chromatogr. Sci., 9 (1971) 687.
143 S.R. Lipsky, C.G. Horvath and W.J. McMurray, Anal. Chem., 38 (1966) 1585.
144 D.P. Lucero and F.C. Haley, J. Gas Chromatogr., 6 (1968) 477.
145 J.E. Lovelock, K.W. Charlton and P.G. Simmonds, Anal. Chern., 41 (1969) 1048.
146 P.G. Simmonds, G.R. Shoemake and J.E. Lovelock, Anal. Chem., 42 (1970) 881.
147 P.M. Llewellyn and D.P. Littlejohn, presented at Pittsburgh Conference on
Analytical Chemistry and Applied Speatrosaopy, Febryary 1966.
148 D.R. Black, R.A. Flath and R. Teranishi, J. Chrornatogr. Sci., 7 (1969) 284.
149 J.E. Hawes, R. Mallaby and W.P. Williams, J. Chrornato~r. Sai., 7 (1969) 690.
150 T.A. Gou9h and C.F. Simpson, J. Chrornatogr., 68 (1972) 31.
151 K. Grob and A. Jaeggi, Anal. Chern., 45 (1973) 1788.
152 D. Henneberg, U. Henrichs and G. Schomburg, Chrornatographia, 8 (1975) 449.
153 D. Henneberg, U. Henrichs, H. Husmann and G. Schomburg, J. Chrornatogr.,
167 (1978) 139.
154 D.S. Millington, M.E. Buoy, G. Brooks, M.E. Harper and K. Griffiths,
Biomed. Mass Speatrorn., 2 (1975) 219.
155 E.M. Chait and P.A. Strauss, Int. Lab., July/August (1979) 71.
156 D.F. Hunt and F.W. Crow, Anal. Chern., 50 (1978) 1781.
157 A.G. Marshall and M.B. Comisarow, Anal. Chern., 47 (1975) 491A.
158 R.L. White, E.B. Ledford, Jr., S. Ghaderi, C.L. Wilikins and M.L. Gross,
Anal. Chern., 52 (1980) 1525.
159 E.B. Ledford, Jr., S. Ghaderi, R.L. White, R.B. Spencer, P.S. Kulkarni,
C.L. Wilkins and t4.L. Gross, AnaL Chern., 52 (1980) 463.
160 E.B. Ledford, Jr., R.L. White, S. Ghaderi, C.L. Wilkins and M.L. Gross,
Anal. Chern., 52 (1980) 2450.
161 Model TSQ-46 GC/MS/MS/DS, Finnigan-MAT, San Jose, CA, 1982.
309

162 R.M. Milberg and J.C. Cook, Jr., J. Chromatogr. Sci., 17 (1979) 17.
163 B. Munson and F.H. Field, J. Amer. Chern. Soc., 88 (1966) 2621.
164 F.H. Field, Ace. Chern. Res., 1 (1968) 42.
165 D.M. Shoengold and B. Munson, Anal. Chern., 42 (1970) 1811.
166 G.P. Arsenault, J.J. Dolhun and K. Biemann, Chern. Commun., (1970) 1512.
167 G.P. Arsenault, J.J. Dolhun and K. Biemann, Anal. Chern., 43 (1971) 1720.
168 F.W. Crow and J.W. Russell, Anal. Chern., 48 (1976) 2098.
169 J.N. Damico and R.P. Barron, Anal. Chern., 43 (1971) 17.
170 H.D. Beckey, in R.I. Reed (Editor), Mass Spectrometry, Academic Press,
London, 1965, p. 93.
171 R.M. Milberg and J.C. Cook, Jr., J. Chromatogr. Sci., 17 (1979) 43.
172 E.C. Horning, M.G. Horning, 0.1. Carroll, 1.0. Dzidic and R.N. Stillwell,
Anal. Chern., 45 (1973) 936.
173 M.W. Siegel and M.C. McKeown, J. Chromatogr., 122 (1976) 397.
174 P.H. Dawson (Editor), Quadrupole Mass Spectrometry and Its Applications,
Elsevier, Amsterdam, 1976.
175 D. Price and J.F.J. Todd (Editors), Dynamic Mass Spectrometry, Heyden &Son,
London, Vol. 5, 1980 and Vol. 6, 1981.
176 G.C. Statford, Jr., P.E. Kelley, E.P.J. Syka, W.E. Reynolds and J.F.J.
Todd, Int. J. Mass Spectrom. Ions ~ocesses, 60 (1984) 85.
177 G.C. Stafford, Jr., P.E. Kelley and D.C. Bradford, Intern. Lab., September
(1983) 84.
178 5970A Mass SeZective Detector, PUbZication No. 25-5955-8028, Hewlett-
Packard, Avondale, PA, 1983.
179 HP 5970B Mass Selective Detector and HP 5995C Benchtop GC/MS System,
Publication No. ~5-5953-8091, Hewlett-Packard, Avondale, PA, 1984.
180 H.H. Hausdorff, J. Chromatogr., 134 (1977) 13.
181 J.A. de Haseth and T.L. Isenhour, Anal. Chern., 49 (1977) 1977.
182 D.L. Wall and A.W. Mantz, Appl. Spectrosc., 31 (1977) 525.
183 P.J. Coffey, D.R. Mattson and J.C. Wright, ArneI'. Lab., 10 (1978) 126.
184 K. Krishnan, R. Curbelo, P. Chiha and R.C. Noonan, J. Chromatogr. Sci., 17
(1979) 413.
185 D.R. Mattson and R.L. Julian, J. Ch~omatogr. Sci., 17 (1979) 416.
186 D.A. Hanna, G. Hangae, B.A. Hohne, G.W. Small, R.C. Wieboldt and T.L.
Isenhour, J. Chromatogr. Sci., 17 (1979) 423.
187 D.A. Hanna, J.C. Marshall and T.L. Isenhour, J. Chromatogr. Sci., 17 (1979)
434.
188 S. Bourne, G.T. Reddy and P.T. Cunningham, J. Chromatogr. Sci., 17 (1979)
460.
189 K.H. Shafer, S.V. Lucas and R.J. Jakobsen, J. Chrornatogr. Sci., 17 (1979)
464.
190 R.C. Wieboldt, B.A. Hohne and T.L. Isenhour, Appl. Spectrosc., 34 (1980) 7.
191 D. Kuehl, G.J. Kemeny and P.R. Griffiths, Appl. Spectrosc., 34 (1980) 222.
192 W. Herres, Bruker Rep., 2 (1980) 4.
193 L.V. Azarraga, AppZ. Spectrosc., 34 (1980) 224.
194 K.H. Shafer, A. Bj~rseth, J. Tabor and R.J. Jakobsen, J. High Resolut.
Chrornatogr. Chrornatogr. Commun., 3 (1980) 87.
195 Miran Infrared GC Detector, PSS 6-12Z1C GC Detector, Foxboro, North Haven,
CT, 1981.
196 C.L. Wilkins, G.N. Giss, G.M. Brissey and S. Steiner, Anal. Chern., 53 (1981)
113.
197 S.R. Lowry and D.A. Huplel", Anal. Chern., 53 (1981) 889.
198 D.M. Hembree, A.A. Garrison, R.A. Crocombe, R.A. Yokley, E.L. Wehry and
G. Mamantov, Anal. Chern., 53 (1981) 1783.
199 R.L. White, G.N. Giss, G.M. Brissey and C.L. Wilkins, Anal. Chern., 53 (1981)
1778.
200 K.H. Shafer, M. Cooke, F. deRoos, R.J. Jakobsen, O. Rossarion and J.D.
Mulik, Appl. Spect~osc., 35 (1981) 469. '
201 V. Rossiter, Int. Lab., March (1982) 42.
202 V. Rossiter, Int. Lab., June (1982) 40.
310

203 R.W. Crawford, T. Hirschfeld, R.H. Sanborn and C.M. Wong, AnaZ. Chem., 54
(1982) 817.
204 D.F. Gurka and L.D. Betowski, AnaZ. Chem., 54 (1982) 1819.
205 S.T. Sparks, R.B. Lam and T.L. Isenhour, AnaZ. Chem., 54 (1982) 1922.
206 R.B. Lam, D.T. Sparks and T.L. Isenhour, AnaZ. Chem., 54 (1982) 1927.
207 C.L. Wilkins, G.N. Giss, R.L. White, G.M. Vrissey and E.C. Onyiriuka, AnaZ.
Chem., 54 (1982) 2260.
208 P.M. Owens, R.B. Lam and T.L. Isenhour, AnaZ. Chem., 54 (1982) 2344.
209 D.F. Gurka, P.R. Laska and R. Titus, J. Chromatogr. Sci., 20 (1982) 145.
210 G. Hangac, B.A. Hohne and T.L. Isenhour, J. Chromatogr. Sci., 21 (1983) 241.
211 J.A. Yancey, ISA Trans., 22 (1983) 33.
212 P.R. Griffiths, J.A. deHaseth and L.V. Azzarraga, AnaZ. Chem., 55 (1983)
1361A.
213 R.L. White, G.N. Giss, G.M. Brissey and C.L. Wilkins, AnaZ. Chem., 55 (1983)
998.
214 S.S. Williams, R.B. Lam and T.L. Isenhour, AnaZ. Chem., 55 (1983) 1117.
215 S.L. Smith, S.E. Garlock and G.E. Adams, AppZ. Spectrosc., 37 (1983) 192.
216 W. Herres, H. Idstein and P. Schreir, J. High ResoZut. Chromatogr.
Chrornatogr. Commun., 3 (1983) 590.
217 V. Rositer, Int. Lab., May (1984) 70.
218 K.H. Shafer, T.L. Hayes, J.W. Brasch and R.J. Jakobsen, AnaZ. Chem., 56
(1984) 237.
219 D.F. Gurka, M. Hiatt and R. Titus, AnaZ. Chern., 56 (1984) 1102.
220 D.A. Laude, Jr., G.M. Brissey, C.F. Ijames, R.S. Brown and C.L. Wilkins,
AnaZ. Chern., 56 (1984) 1163.
221 J.R. Cooper and L.T. Taylor, AppZ. Spectrosc., 38 (1984) 366.
222 B.A. Hohne, G. Hangac, G.W. Small and T.L. Isenhour, J. Chrornatogr. Sci.,
19 (1981) 283.
223 G.T. Reedy, D.G. Ettinger, J.F. Schneider and S. Bourne, Anal. Chern.,
57 (1985) 1602
224 G.T. Reedy, S. Bourne and P .T. Cunningham, Anal. Chern., 51 (1979) 1535.
225 S. Bourne, G.T. Reedy and P .T. Cunningham, J. Chrornatogr. Sci., 17
(1979) 460.
226 S. Bourne, G. Reedy, P. Coffey and D. Mattson, Arn. Lab., 16 (June)
(1984) 90.
311

Chapter 14

CONCLUSION

Chapter 1 discussed the advantages of using selective detectors. Another


important fact follows from the chapters dealing with the individual selective
detectors. With most of these detectors the minimum detectability is either
approximately equal to or, in many instances, lower than that obtainable with
the commonly used non-selective flame-ionization detector (FlD). Thus, in
addition to a selective response, selective detectors usually also yield a
lower detection limit.
The application of selective detectors is most advantageous if they are used
in an on-line combination with a non-selective detector. Two chromatograms can
thus be obtained from a single chromatographic analysis; a non-selective record
containing all peaks, and a selective record. This combination of two detectors
is commonly realized with parallel connection, where the effluent from the
chromatographic column is split and supplied to two detectors. With series
connection of the detectors the effluent passes first through a non-destructive
detector ~lectron-capture detector 1 (ECD)].
A direct dual non-selective-selective record is possible with detectors that
simultaneously combine an FlD and a selective detector in a single detection
system: an alkali flame-ionization detector (AFlD) or flame-photometric detector
(FPD) (see the respective chapters).
The second way to obtain simultaneously selective and non-selective records
is to use a detection system that provides both records. Plasma emission
spectroscopy can give carbon- and element-selective chromatograms simulta-
neously2.

REFERENCES

1 P. Gagliardi and G.R. Verga, J. Chromatogr., 279 (1983) 323.


2 K.J. Slatkavitz, P.C. Uden, L.D. Hoey and R.M. Barnes, J. Chromatogr., 302
(1984) 227.
312

The following section was added in proof (see p. 174)

8.5A. REDOX CHH1ILUt1INESCENCE DETECTOR

The redox chemiluminescence detector 1 (RCD) is based on catalyzed redox reac-


tions of nitrogen dioxide with reducing agents to form nitroge~ oxide, which is
detected by an ozone chemiluminescence technique. Nitrogen dioxide in helium
(ca. 100 ppm) is introduced post-column just prior to a heated reaction zone,
where a catalyst (gold) is present. In the presence of any solute that reduces
nitrogen dioxide, nitrogen oxide is produced 2 :

reducing solute + N0 2 Au 0" oxidized species + NO


150-400 C

Nitrogen oxide produced is re-oxidized with ozone and the emission arlslng from
the decay of excited N0 2 back to its ground state is detected. The RCD thus
provides response to the compounds that serve as reducing agents.
Compound that can be detected are 1 alcohols, aldehydes, ketones, acids,
phenols, olefins, aromatic hydrocarbons, amines, thiols, sulphides, hydrogen,
ammonia, hydrogen peroxide and sulphur dioxide. No response is obtained with
low-molecular weight paraffins, water, carbon dioxide and argon. Common com-
pounds such as methylene chloride, l,2-dichloroethane, tetrachloroethylene and
tetrahydrofuran do not generate any appreciable response.
Not only the type of compound, but also the reaction temperature 3 affects
the selectivity of response. For instance, in the analysis of a gasohol only
methanol is detected at 360 0 C. At 390°C other components of gasoline, such as
the aromatic fractions, are detected, but not the saturated alkanes. At 420°C
the chromatogram becomes very complex and resembles that obtained with an FID.
When palladium used as a catalyst instead of gold, alkanes produce a response
at a temperature as low as 250 0 C. The response to unsaturated hydrocarbons is
much greater than that to saturated ones.
The detection limit 1 is 200 pg and the linear dynamic range is three orders
of magn itude.

REFERENCES

1 S.A. Nyarady, R.M. Barkley and R.E. Sievers, Anal. Chem., 57 (1985) 2074.
2 S.A. Nyarady and R.E. Sievers, J. Am. Chem. Soc., 107 (1985) 3726
3 R.E. Sievers, S.A. Nyarady, R.L. Shearer, J.J. DeAngelis, R.11. Barkley and
R.S. Hutte, J. Chromatogr., 349 (1985) 395.
313

LIST OF ABBREVIATIONS

AAS atomic absorption spectrometer


AFIO alkali flame-ionization detector
API atmospheric pressure ionization
CO coulometric detector
CFID catalytic flame-ionization detector
CI chemical ionization
CLO chemiluminescence detector
d.c. direct current
OCP d.c. plasma
ECO electron-capture detector
EI electron impact ionization
ELCD electrolytic conductivity detector
FASO flameless alkali sensitized detector
FI field ionization
FlO flame-ionization detector
FPO flame ohotometric detector
FTIR Fourier transform infrared spectroscopy
Fn1S Fourier transform mass spectrometry
GC gas chromatoqraohy
ICP inductively coupled plasma
IMO ion mobility detector
IP ionization potential
IR infrared
ITO ion trap detector
HAFIO hydrogen atmosphere flame-ionization detector
HAFID-Si hydrogen atmosphere flame-ionization detector for silicon compounds
M general flame gas molecule
MIP microwave-induced plasma
MPI multi-photon ionization
MS mass spectrometry
NPO nitrogen-phosphorus detector
N mode nitrogen mode
NP mode nitrogen-ohosphorus mode
PIO photoionization detector
P mode phosphorus mode
314

R response
RCD redox chemiluminescence detector
RMR relative molar response
RWR relative weight response
S sensitivity
SECS selective electron-capture sensitization
SIM selected ion monitoring
TEA thermal energy analyser
TIC total ion current
TID thermionic detector
TZ Trennzahl (separation number)
UV ultraviolet
315

SUBJECT INDEX

A - , FPD 149
- , HAFID 93-95, 100
Alkali flame-ionization detector - . PID 116
15-59, 63 Barbi turates
- , air in 18-20, 34, 53, 54 ,ELCP 188
- , carrier gas 18-20, 28, 29, 31, - , FASD 66
37, 38, 53, 54 - , IMD 278
- , collector electrode 25, 26, 28, -,PID117,125
34-36, 45-48 Boron compounds
- , design 16-23 AFID 39
- , double flame 23, 32, 37, 38, ,ECD 245-247
40, 44, 54 --, FPD 149
- , effect of anions 49-52
- , effect of cations 49-52
- , effect of compound structure C
41-43
, effect of temperature 52 Chemigram 301
- , flame 15, 16, 21 Chemi-ionization detector 83, 85
- , gate electrode 22, 32, 45 Chemiluminescence detector 161-179,
- , hydrogen in 16, 18-20, 24-26, 312
28-31, 33-35, 38, 39 - , design 162, 173, 176
- , jet tip bore 31, 36,48,49 - , effect of temperature 163, 168,
- , life 16, 23, 24 169,312
- , negative response 29-33, 34, - , fluorine-induced detector 177-179
36-38, 40, 45, 52-54, 59 - for nitroaromatic compounds
, peak tailing 40 169, 170
- , principle parameters 18-21 -- for nitrogen-containing compounds
- , single flame 23 170-174,312
- , three-electrode construction - for N-nitroso compounds 161-168
21, 22, 25, 32, 34 - with sodium metal 174-177
- , temperature of the flame 24, 25 Concentration-sensitive detector 6
- , voltage 21, 22, 32, 43-45 Coulometric detector 209-215
Amines - , design 209
,CLD 169, 171, 172, 312 , effect of temperature 212-214
- , ECD 223, 255, 257 - , nitrogen mode 211
- , IMD 284 - , oxidative mode 210
- , RCD 312 --, reductive mode 210
Arsenic compounds Coulometric response in ECD 263-265
, AFID 19, 38, 50 Coulson detector 182, 183, 196, 201,
- , FPD 149 204
- , HAFID 94
- , PID 117
Atomic-absorption spectrometer 294 D

Detection limit 7,8


B ,AAS 294
,CD 210
Background current CLD 165, 177, 178,312
- , AFID 18-21, 24, 29, 30, 32, ECD 222, 224, 239, 241, 246, 252
36-40,44,47,48,51-54 ElCD 181, 194, 199, 205
- , FASD 69, 72-74, 76, 77, 79, 85 FPD 145, 149
316

HAFID 94-96, 98, 99, 101, 103 --, direct current mode with constant
,IMD 286 current 229
IR 300 --, effect of compound structure
ion-selective electrode 295 235-239, 243-248
,MS 298 --, effect of detector volume 266-269
--, PID 116, 117 --, effect of temperature 239-241,
--, piezoelectric sorrtion detector 244, 247
296 --, effect of voltage 224, 226
--, plasma-emission spectrometer --, gases in 221, 250, 267-269
293 --, impurities in 248
--, RCD 312 -- pulse mode with constant current
Detector characterization parameters 227-229, 249, 262
5-13 --, pulse mode with constant frequency
Detector 225-227, 249, 262
--, halogen compounds 87-89 --, pulse period 226
--, nitroaromatic compounds 170, --, source of electrons 219-223
171 --, space charge effect 225
--, nitrogen-containing compounds
171-174
--, N-nitroso compounds 161-168 F
--, reducing compounds 312
Detector linearity 10 Flame-ionization detection 91-106
Detector noise 6, 8-10 Flame-ionization detector 15, 20,
AFID 20-22 27 37,64,91-106
J

,ECD 220 --, flow-rate of hydrocarbon in


,FASD 70,71,73, 74 105, 106
HAFID 93 Flameless alkali sensitized detector
PID 116 63-89
, temperature dependence 10 --, bead current 86
Detector response 5, 6 --, design 64, 72, 74, 79, 85, 88
Detector sensitivity 5, 6 --, effect of compound structure
Detector, specific 13 66, 68, 81-83
Detector, substance-selective 13 --, effect of voltage 64, 73, 75, 76,
Drift speed 276 88
Drift time 276 --, gases in 65,69,72,74,76,77,
Dynamic detector range 11, 12 79, 80, 88
- , heatin9 current 69, 86
-,life 84-87
E - , principal parameters 70-74,80,81
- , temperature of alkali source 75,
Electrolytic conductivity detector 79
181-206 Flame photometric detector 133-158
--, design 181-185 --, design 134, 135
--, effect of compound structure --, dual channel 135, 144
109-192 - , dual flame 147, 148, 154-156
--, effect of tempera ture 187, 189, - , effect of compound concen tra t ion
191-193, 195,201,203-206 139-141
--, electrode 195, 196 - , effect of compound structure
, gases in 186-188, 201-203 138, 139
--,life 193, 205, 206 - , effect of temperature 142
--, principal parameters of 194 - , flame stability 157
--, solvent in 196-201 --, gases in 135, 137, 138
Electron capture coefficient 235, - , interference fil ter 142, 143
237, 249 -,photomultiplier 143
Electron capture detector 217-269 --, sul phur background 151, 152
--, bipolar pulsed mode 251 Fluorine-induced detector 177-179
--, contact potential 225
--, design 218, 219
--, direct-current mode 224, 225,
249
317

G
Germanium compounds Infrared spectroscopy 300-305
- , FPD 146, 147 - , matrix isolation 302,303
- , HAFID 94 Ion mobility detector 275-288
Gram-Schmidt reconstruction 300 - , different modes of operation
280-282
- , source of ionization 275, 282,
H 283, 285
Ion-selective electrodes 294, 295
Hall detector 182-184, 196, 200, Ion trap detector 299
201, 204 Iron compounds
Halogen compounds ,FPD 149
- , AFID 18,19,26-29,31,33,36, - , AFID 94,95, 103
37,41-43,49,50,52,53,55,57,
59
- , CD 210, 213, 214 L
--, CLD 176, 177, 312
- , ECD 238, 239, 242-244, 246, Lead compounds
247,253,257,259-262 ,AFID 18, 29, 39, 40, 55
ELCD 186, 188, 199, 204 - , FPD 149
FASD 71, 87-89 - , HAFID 94, 97, 100
FID 106 - , PID 117
FPD 147-149 Linear dynamic range 11
HAFID 94,95, 103 AAS 294
IMD 278, 280, 281, 286 CD 212
PID 114, 123, 124 CLD 165, 169, 171, 177, 312
,RCD 312 ECD 249-251
Heteroatom 7,9, 13, 15,20,21, ELCD 193
30 FASD 70, 72-74, 81
Hydrocarbons FPD 150, 151
AFrD 19, 22, 29, 37, 53 HAFID 97, 103
CLD 174, 175, 312 ion-selective electrode 295
ECD 253, 254, 257 PID 115-117
FASD 71,88 , piezoelectric sorption detector
FPD 144 296
HAFID 93, 95, 96, 98-100, 102 - , plasma-emission spectrometer 293
IMD 278, 284 - , RCD 312
PID 112-114, 117, 119, 120 Long-term noise 9
,RCD 312
Hydrogen atmosphere flame-ionization
detector 92-104 M
air in 102
carrier gas 102 Mass detector 299
collector electrode 100, 101 Mass spectrometry 296-299
design 93, 102 - , atmospheric pressure ionization
for silicon compounds 102-104 299
hydrogen in 102 - , chemical ionization 299
iron compounds 102, 103 - , electron impact ionization 298
, negative response 100, 101, , field ionization 299 .
103, 104 - , mass fragmentography 298
oxygen in 102 - , selected ion monitoring 297
, peak tailing 102 - , total ion current 297
- , potential 100 Mass-rate sensitive detector 6
- , principal parameters 98 Matrix isolation, IR spectroscopy
- , silicon compounds 93, 101, 102 302, 303
Hypercoulometric response in ECD Minimum detectability 6,8, 9
265 - , AFID 18-20, 33, 35-39
318

CLD 175 Photoionization detector 109-131


ECD 239 carrier gas 126-131
ELCD 194 design 109, 110
FASD 70-74, 81 principal parameters 115
-, FlO 20 radiation source 109, 110
-, FPD 137-143 , solvents with no response 123
-, HAFID 93-95 Photoionization cross-section 111,
-, PID 117 112, 114
- , plasma-emission spectrometer 293 Photoionization efficiency 11, 112,
~1inimum detectable solute concentra- 113
tion 6-8, 11 Piezoelectric sorption detector
Minimum detectable solute mass-rate 295, 296
6-8, 10 .. 11 Plasma chromatography 275
Molecular separators 297 Plasma-emission spectrometry 291,
293
N
Q
Nitro compounds
AFID 23, 35 Qualitative analysis 1-3
CLD 167-170
ECD 237
ELCD 199 R
FASD 66
IMD 278, 284 Reaction gas chromatography 2
,PlO 125 Redox chemiluminescence detector 312
Nitr0gen compounds --, effect of catalyst 312
- , AFID 18-22, 29, 34-36, 49, 50, --, effect of reaction temperature
53, 57 312
,CD 211,212 Response
- , CLD 170-174 --, molar 6
,ELCD 186-188, 199 --, negative 13,29-34,36-38,40,
- , FASD 66,68,69, 71,73 45, 52-54, 59, 100, 101, 103, 104
- , HAFID 95 of detector 5, 6
- , PID 117, 128 ,polarity 13
Nitrogen-phosphorus detector 16, 63 --, relative molar 12
Nitroso compounds - , relative weight 12
,CLD 163-169 --, specific 13
- , ELCD 187-190 Response mechanism
--, AFID 54-59
--, CD 210-212
o --, CLD 161-164, 170, 174, 176,
177, 312
Organometallic compounds ECD 230-235, 252, 258
- , ECD 237, 239 --, ELCD 186, 187
- , FPD 149 --, FASD 64-67,72,77-79,84,85,88
- , HAFID 93-96, 102 FlD 91, 92
FPD 136, 137
HAFID 92, 102
P lMD 275-278
ion selective electrode 294
Phosphorus compounds --, lR 300, 301
- , AFID 18-22, 26, 33, 34, 41, 43, MS 297, 298
49, 50, 53, 55, 57, 58 , piezoelectric sorption detector
CD 211 295
ELCD 186 --, plasma-emission spectrometer
FASD 64,65,68,71,73 291, 292
FPD 140, 144, 157 --, RCD 312
,HAFID 94, 103
- , PID 117
319

Response quenching T
ECO 248, 249
,ELCO 190, 192 Thermal energy analyzer 163, 169,
--, FPO 152-156 170
--, IMO 287, 288 Thermionic detector 16, 163
--, PID 122 Tin compounds
Response reproducibility --, AFID 18, 19, 39, 40, 55
--, AFIO 23,24 --, FPD 147-147
--, ELCO 200 --, HAFIO 94, 98-100
--, FASO 84-87 Trennzahl (see Separation number)
Response selctivity 12, 13
--, AFIO 13, 33, 36, 37
--, CLO 167-169, 312 W
--, ECO 237-239, 242, 244, 253,
254, 260, 261 Weight response 6
ELCD 186-189
FASO 70-74, 76, 81
FPO 144, 145
HAFID 94, 94, 98, 99, 102-105
Hm 281, 282
,IR 301
--, MS 298
--, PIO 118-126
--, piezoelectric sorption detector
296
--, plasma-emission spectrometer
293
--, RCD 312
Response sensitization in ECO
251-263
aromatic hydrocarbons 263
, effect of nitrous oxide 252-256
--, effect of oxygen 257-263

Selenium compounds in FPO 149, 155


Separation number 1
Short-term noise 8,9
Silicon compounds
--, AFID 39, 40
--, FlO 105, 106
--, HAFIO 102-105
Solute switching 265, 266
Substance selective response 7
Sulphur compounds
AFIO 18, 20, 29-32, 38, 41, 53
CD 210,211
CLO 174, 178, 179, 312
ELCD 188, 200
FPO 141, 144, 157
HAFID 95
PID 117
,RCD 312
Surface ionization detector 87
This Page Intentionally Left Blank

You might also like