You are on page 1of 135

Diabetes

Introduction to Diabetes

1. Introduction to Diabetes
Created: July 7, 2004

Diabetes mellitus is characterized by abnormally high levels of sugar (glucose) in the blood.
When the amount of glucose in the blood increases, e.g., after a meal, it triggers the release of
the hormone insulin from the pancreas. Insulin stimulates muscle and fat cells to remove glucose
from the blood and stimulates the liver to metabolize glucose, causing the blood sugar level to
decrease to normal levels [http://www.ncbi.nlm.nih.gov:80/books/bv.fcgi?call=bv.View..
ShowSection&rid=mcb.figgrp.5903].
In people with diabetes, blood sugar levels remain high. This may be because insulin is not
being produced at all, is not made at sufficient levels, or is not as effective as it should be. The most
common forms of diabetes are type 1 diabetes (5%), which is an autoimmune disorder, and type 2
diabetes (95%), which is associated with obesity. Gestational diabetes is a form of diabetes that
occurs in pregnancy, and other forms of diabetes are very rare and are caused by a single gene
mutation.
For many years, scientists have been searching for clues in our genetic makeup that may
explain why some people are more likely to get diabetes than others are. "The Genetic Landscape
of Diabetes" introduces some of the genes that have been suggested to play a role in the
development of diabetes.

1-1

Diabetes

Introduction to Diabetes

Classification
Diabetes is classified by underlying cause. The categories are: type 1 diabetesan autoimmune
disease in which the body's own immune system attacks the pancreas [http://www.ncbi.nlm.nih.
gov:80/books/bv.fcgi?tool=bookshelf&call=bv.View..ShowSection&rid=imm.figgrp.1942], rendering it unable to produce insulin; type 2 diabetesin which a resistance to the effects of insulin or
a defect in insulin secretion may be seen; gestational diabetes; and other types. Table 1 compares the presentation (phenotype) of type 1 and type 2 diabetes.
Table 1. Comparison of Type 1 and Type 2 Diabetes

Phenotype

Type 1 diabetes

Type 2 diabetes

Onset primarily in childhood and adolescence


Often thin or normal weight
Prone to ketoacidosis
Insulin administration required for survival
Pancreas is damaged by an autoimmune
attack
Absolute insulin deficiency

Onset predominantly after 40 years of age*


Often obese
No ketoacidosis
Insulin administration not required for survival
Pancreas is not damaged by an autoimmune
attack
Relative insulin deficiency and/or insulin
resistance
Treatment: (1) healthy diet and increased
exercise; (2) hypoglycemic tablets; (3)
insulin injections
Increased prevalence in relatives
Identical twin studies: usually above 70%
concordance
HLA association: No

Treatment: insulin injections

Genotype

Increased prevalence in relatives


Identical twin studies: <50% concordance
HLA association: Yes

* Type 2 diabetes is increasingly diagnosed in younger patients.

Type 2 diabetes commonly occurs in adults who are obese. There are many underlying factors that contribute to the high blood glucose levels in these individuals. An important factor is the
body's resistance to insulin in the body, essentially ignoring its insulin secretions. A second factor
is the falling production of insulin by the beta cells of the pancreas. Therefore, an individual with
type 2 diabetes may have a combination of deficient secretion and deficient action of insulin.
In contrast to type 2, type 1 diabetes most commonly occurs in children and is a result of the
body's immune system attacking and destroying the beta cells. The trigger for this autoimmune
attack is not clear, but the result is the end of insulin production.

References
1. Expert Committee on the Diagnosis and Classification of Diabetes Mellitus. Report of the expert
committee on the diagnosis and classification of diabetes mellitus. Diabetes Care 26:S5-S20; 2003.
(PubMed)

1-2

Diabetes

Introduction to Diabetes

History of Diabetes
Physicians have observed the effects of diabetes for thousands of years. For much of this time,
little was known about this fatal disease that caused wasting away of the body, extreme thirst,
and frequent urination. It wasn't until 1922 that the first patient was successfully treated with
insulin.
One of the effects of diabetes is the presence of glucose in the urine (glucosuria). Ancient
Hindu writings, many thousands of years old, document how black ants and flies were attracted to
the urine of diabetics. The Indian physician Sushruta in 400 B.C. described the sweet taste of
urine from affected individuals, and for many centuries to come, the sweet taste of urine was key
to diagnosis.
Around 250 B.C., the name diabetes was first used. It is a Greek word that means to
syphon, reflecting how diabetes seemed to rapidly drain fluid from the affected individual. The
Greek physician Aretaeus noted that as affected individuals wasted away, they passed increasing
amounts of urine as if there was liquefaction of flesh and bones into urine. The complete term
diabetes mellitus was coined in 1674 by Thomas Willis, personal physician to King Charles II.
Mellitus is Latin for honey, which is how Willis described the urine of diabetics (as if imbued with
honey and sugar).
Up until the mid-1800s, the treatments offered for diabetes varied tremendously. Various fad
diets were prescribed, and the use of opium was suggested, as were bleeding and other therapies. The most successful treatments were starvation diets in which calorie intake was severely
restricted. Naturally, this was intolerable for the patient and at best extended life expectancy for a
few years.
A breakthrough in the puzzle of diabetes came in 1889. German physicians Joseph von Mering and Oskar Minkowski surgically removed the pancreas from dogs. The dogs immediately
developed diabetes. Now that a link was established between the pancreas gland and diabetes,
research focused on isolating the pancreatic extract that could treat diabetes.
When Dr. Frederick Banting took up the challenge of isolating a pancreatic extract, he was
met with much skepticism. Many great physiologists had tried and failed to isolate an internal
secretion from the pancreas. But Banting, a surgeon, persisted and in May 1921, he began work
in the laboratory of Professor John Macloed in Toronto, Canada. Charles Best, a medical student
at the time, worked as his assistant.
To concentrate what we now know as insulin, Banting tied the pancreatic ducts of dogs. The
pancreatic cells that released digestive enzymes (and could also destroy insulin) degenerated,
but the cells that secreted insulin were spared. Over several weeks the pancreas degenerated
into a residue from which insulin could be extracted. In July 1921, a dog that had had its pancreas
surgically removed was injected with an extract collected from a duct-tied dog. In the two hours
that followed the injection, the blood sugar level of the dog fell, and its condition improved.
Another de-pancreatized (diabetic-like) dog was kept alive for eight days by regular injections
until supplies of the extract, at that time called "isletin", were exhausted.
Further experiments on dogs showed that extracts from the pancreas caused a drop in blood
sugar, caused glucose in the urine to disappear, and produced a marked improvement in clinical
condition. So long as the extract was being given, the dogs were kept alive. The supply of the
1-3

Diabetes

Introduction to Diabetes
extract was improved: the pancreas of different animals were used until that of the cow was settled upon. This extract kept a de-pancreatized dog alive for 70 days. Dr. J. Collip, a biochemist,
was drafted to continue improving the purity of the pancreas extract, and later, Best carried on
this work.
A young boy, Leonard Thompson, was the first patient to receive insulin treatment. On January 11, 1922, aged 14 and weighing only 64 pounds, he was extremely ill. The first injections of
insulin only produced a slight lowering of blood sugar level. The extract still was not pure enough,
and abscesses developed at the injection site. Collip continued to refine the extract. Several
weeks later, Leonard was treated again and showed a remarkable recovery. His blood sugar levels fell, he gained weight and lived for another 13 years. He died from pneumonia at the age of
27.
During the spring of 1922, Best increased the production of insulin to enable the treatment of
diabetic patients coming to the Toronto clinic. Over the next 60 years, insulin was further refined
and purified, and long-acting and intermediate types were developed to provide more flexibility. A
revolution came with the production of recombinant human DNA insulin in 1978. Instead of collecting insulin from animals, new human insulin could be synthesized.
In 1923, Banting and Macloed were awarded the Nobel Prize for the discovery of insulin.
Banting split his prize with Best, and Macloed split his prize with Collip. In his Nobel Lecture,
Banting concluded the following about their discovery:
Insulin is not a cure for diabetes; it is a treatment. It enables the diabetic to burn sufficient
carbohydrates, so that proteins and fats may be added to the diet in sufficient quantities to provide energy for the economic burdens of life.

Link Roundup
Discovery of Insulin
Nobel Prize [www.nobel.se/medicine/laureates/1923/index.html] for the discovery of insulin
Development of Insulin [http://digital.library.utoronto.ca/insulin/], University of Toronto
Banting Digital library [http://www.newtecumseth.library.on.ca/banting/], New Tecumseth Public Library, Canada
Discovery of insulin [www.discoveryofinsulin.com/Home.htm]

1-4

Diabetes

Introduction to Diabetes

Epidemiology
Over 18 million Americans have diabetes; of these, about 5 million do not know they have the
disease (1).
Type 1 diabetes accounts for 5-10% of cases, affecting 1 of 400 children and adolescents.
Type 2 diabetes is extremely common, accounting for 90-95% of all cases of diabetes. This
form of diabetes can go undiagnosed for many years, but the number of cases that are being
diagnosed is rising rapidly, leading to reports of a diabetes epidemic.

The Type 2 Diabetes Epidemic


When people think of epidemics, they often think of infectious diseases such as SARS, HIV, or
the flu. However, the prevalence of type 2 diabetes is now at epidemic proportions. In the United
States, diabetes accounts for over 130 billion dollars of health care costs and is the fifth leading
cause of death (2). The number of new cases being diagnosed continues to rise. It has been
estimated that of the children born in the year 2000, 1 of 3 will suffer from diabetes at some point
in their lifetime (3). Diabetes is predicted to become one of the most common diseases in the
world within a couple of decades, affecting at least half a billion people (4).
Estimate your risk of developing Type 2 Diabetes [www.healthcalculators.org/calculators/diabetes.asp]

In the past, type 2 was rarely seen in the young, hence its original name of adult-onset diabetes. But now type 2 diabetes is increasingly being diagnosed in young adults and even in
children. In Japan, more children suffer from type 2 than type 1 (juvenile onset) diabetes. This
young generation of diabetics will have many decades in which to develop the complications of
diabetes.
In 1990, 4.9% of the American population were diagnosed with diabetes (see Flash Animation 1). This increased to 7.9% by the year 2001 (5).

Obesity
The driving force behind the high prevalence of diabetes is the rise of obesity [http://www.ncbi.
nlm.nih.gov:80/books/bv.fcgi?call=bv.View..ShowTOC&rid=obesity.TOC] in the population. In
today's society, it can be difficult to maintain a healthy weight. We have the combination of ample
food and a sedentary lifestyle. This is in stark contrast to only a couple of hundred years ago,
when people were more active and food supplies were not as abundant. As a result, many of us
are heavier than we should be.
Calculate your ideal weight [www.drkoop.com/template.asp?page=ibw&ap=93]

Being overweight or obese is defined by a calculation called the Body Mass Index (BMI). It is
a calculation that takes your height and weight into consideration and gives you a score. A score
of 1824.9 is a healthy weight. If you are overweight, your score lies within the range to 2529.9;
a score of 30 and above indicates obesity.
1-5

Diabetes

Introduction to Diabetes
Calculate your BMI [http://nhlbisupport.com/bmi/bmicalc.htm]

In 1991, it was estimated that 12% of the population were obese (5). By the year 2001, this
had increased to an estimated 20.9% of the population; this represents over 44 million obese
adult Americans. A more recent study estimated that a record 30% of the American population
are now obese (6) (see Flash Animation 2).
Obesity is a major problem for the United states. Every year, an estimated 300,000 US adults
die of causes related to obesity (7). Obesity is also a huge economic burden, accounting for up to
4% of healthcare costs in the United States (8).

Thrifty Genes
Epidemics of infectious diseases increase when there is increased spread of the infectious agent
and decrease when the number of victims who are susceptible falls (they either become immune
or they die). An epidemic of a genetic disease such as type 2 diabetes is similar. The number of
cases rises when there is a rise in environmental risk (abundant food supplies, lack of activity)
and decreases when the number of susceptible individuals falls (by deaths from the complications
of diabetes).
The classic example of an epidemic of diabetes is found on an remote island in the Pacific
Ocean, the island of Nauru. Before the turn of the 20th century, the lifestyle of Nauruans was
harsh. The soil was poor, agriculture was difficult, and frequent episodes of starvation were common. Despite these adverse conditions, the islanders were noted to be heavy. In 1922, it was
discovered that Nauru contained phosphate rock, which was then mined for use in fertilizer, and
for which the islanders received royalties. Over several decades, the Nauruans became
extremely wealthy, and with their new-found riches came major lifestyle changes. Food was now
abundant and could be bought from stores. Instead of fishing and farming, Nauruans now led
sedentary lives. By the 1950s, type 2 diabetes exploded from being non-existent in this population to affecting 2 of 3 adults over the age of 55 and becoming a common cause of death.
The case of the Nauruans is an extreme case of how type 2 diabetes can rapidly reach epidemic proportions, and thrifty genes may be involved. It has been postulated by Neel (9) that
genes that are metabolically thrifty give a survival advantage in times when there is a constant
threat of famine and starvation. When food is abundant, these genes aid the efficient metabolism
of the food, enabling rapid build up of fat stores. This enabled people like the Nauruans to survive
food shortages later on. But when food is always abundant, a thrifty genetic makeup turns into a
survival disadvantage. Thrifty genes cause obesity, which in turn predisposes to diabetes. The
epidemic that took hold of the island of Nauru is now emerging in developing countries and
already has a firm hold on the developed world.

References
1. National Diabetes Statistics fact sheet: general information and national estimates on diabetes in the
United States. National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of
Health. U.S. Department of Health and Human Services.

1-6

Diabetes

Introduction to Diabetes

2. Hogan P, Dall T, Nikolov P. Economic costs of diabetes in the US in 2002. Diabetes Care 26(3):917932;
2003. (PubMed)
3. Narayan KM, Boyle JP, Thompson TJ, Sorensen SW, Williamson DF. Lifetime risk for diabetes mellitus in
the United States. JAMA 290(14):18841890; 2003. (PubMed)
4. King H, Aubert RE, Herman WH. Global burden of diabetes, 1995-2025: prevalence, numerical
estimates, and projections. Diabetes Care 21(9):14141431; 1998. (PubMed)
5. Mokdad AH, Ford ES, Bowman BA, Dietz WH, Vinicor F, Bales VS, Marks JS. Prevalence of obesity,
diabetes, and obesity-related health risk factors, 2001. JAMA 289(1):7679; 2001. (PubMed)
6. Flegal KM, Carroll MD, Ogden CL, Johnson CL. Prevalence and trends in obesity among US adults,
1999-2000. JAMA 288(14):17231727; 2002. (PubMed)
7. Allison DB, Fontaine KR, Manson JE, Stevens J, VanItallie TB. Annual deaths attributable to obesity in
the United States. JAMA 282(16):15301538; 1999. (PubMed)
8. Allison DB, Zannolli R, Narayan KM. The direct health care costs of obesity in the United States. Am J
Public Health 89(8):11941199; 1999. (PubMed)
9. Neel JV. Diabetes mellitus: a "thrifty" genotype rendered detrimental by "progress"? JAMA 14:353362;
1962. (PubMed)

Box 1: Increase of diabetes in adults in the United States.


References
Mokdad AH, Ford ES, Bowman BA, Dietz WH, Vinicor F, Bales VS, Marks JS. Prevalence of obesity, diabetes, and obesity-related
health risk factors, 2001. JAMA 289(1):7679; 2001. (PubMed)
To view this you will need to have Flash [http://www.macromedia.com/go/getflashplayer] installed on your computer.

Box 2: Rise of obesity in the United States.


References
Mokdad AH, Ford ES, Bowman BA, Dietz WH, Vinicor F, Bales VS, Marks JS. Prevalence of obesity, diabetes, and obesity-related
health risk factors, 2001. JAMA 289(1):7679; 2001. (PubMed)
To view this you will need to have Flash [http://www.macromedia.com/go/getflashplayer] installed on your computer.

Link Roundup
Calculators
Diabetes Type 2 Risk Calculator [www.healthcalculators.org/calculators/diabetes.asp], University of Maryland Medical Center.
Calculate your ideal weight [www.drkoop.com/template.asp?page=ibw&ap=93], according to Dr. Koop.
Calculate your BMI [nhlbisupport.com/bmi/bmicalc.htm] at the National Heart, Lung, and Blood Institute.

Healthy Living
The DASH diet [www.nhlbi.nih.gov/health/public/heart/hbp/dash/], National Heart, Lung, and Blood Institute.
Advice from the Obesity Education Initiative [www.nhlbi.nih.gov/about/oei/index.htm], National Heart, Lung, and Blood Institute.
The President's challenge [www.presidentschallenge.org/].
Advice from the Surgeon General [www.surgeongeneral.gov/topics/obesity/calltoaction/fact_whatcanyoudo.htm].

1-7

Diabetes

Introduction to Diabetes

Physiology and Biochemistry of Sugar Regulation


Overview of Glucose Metabolism
Glucose is an essential fuel for the body (Figure 1).The amount of glucose in the bloodstream is
regulated by many hormones, the most important being insulin.
Insulin is the hormone of plentyit is released when glucose is abundant and stimulates the
following:

muscle and fat cells to remove glucose from the blood


cells to breakdown glucose, releasing its energy in the form of ATP (via glycolysis and the
citric acid cycle)

the liver and muscle to store glucose as glycogen (short-term energy reserve)
adipose tissue to store glucose as fat (long-term energy reserve)
cells to use glucose in protein synthesis

1-8

Diabetes

Introduction to Diabetes

Figure 1: Overview of glucose metabolism.


Glucose is used for many purposes in the body. It can be converted into energy via pyruvate and the tricarboxylic acid
(TCA) cycle, as well as being converted to fat (long-term storage) and glycogen (short-term storage). Some amino acids
may also be synthesized directly from pyruvate; thus, glucose may also indirectly contribute to protein synthesis.

Glucagon is the main hormone opposing the action [http://www.ncbi.nlm.nih.gov:80/books/bv.


fcgi?tool=bookshelf&call=bv.View..ShowSection&rid=mcb.figgrp.5903] of insulin and is released
when food is scarce. Whereas insulin triggers the formation of glycogen (an energy-requiring process, or an anabolic effect), glucagon triggers glycogen breakdown, which releases energy (a
catabolic effect). Glucagon also helps the body to switch to using resources other than glucose,
such as fat and protein (Figure 2).

1-9

Diabetes

Introduction to Diabetes

Figure 2: Anabolism and catabolism of glucose.


Glucose metabolism involves both energy-producing (catabolic, shown in orange) and energy-consuming (anabolic,
shown in green) processes.

Regulation of Blood Glucose


Blood glucose levels are not constantthey rise and fall depending on the body's needs, regulated by hormones. This results in glucose levels normally ranging from 70 to 110 mg/dl.
The blood glucose level can rise for three reasons: diet, breakdown of glycogen, or through
hepatic synthesis of glucose.
Eating produces a rise in blood glucose, the extent of which depends on a number of factors
such as the amount and the type of carbohydrate eaten (i.e., the glycemic index), the rate of
digestion, and the rate of absorption. Because glucose is a polar molecule, its absorption across
the hydrophobic gut wall requires specialized glucose transporters (GLUTS) [http://www.ncbi.nlm.
nih.gov:80/books/bv.fcgi?call=bv.View..ShowSection&rid=endocrin.box.54] of which there are five
types. In the gut, GLUT2 and GLUT5 [http://www.ncbi.nlm.nih.gov:80/books/bv.fcgi?call=bv.
View..ShowSection&rid=endocrin.box.53] are the most common.
The liver is a major producer of glucoseit releases glucose from the breakdown of glycogen
and also makes glucose from intermediates of carbohydrate, protein, and fat metabolism. The
liver is also a major consumer of glucose and can buffer glucose levels (see Box 1). It receives
glucose-rich blood directly from the digestive tract via the portal vein (Figure 3). The liver quickly
removes large amounts of glucose from the circulation so that even after a meal, the blood glucose levels rarely rise above 110 mg/dl in a non-diabetic.

1-10

Diabetes

Introduction to Diabetes

Figure 3: The portal circulation.


The portal vein drains almost all of the blood from the digestive tract and empties directly into the liver. This circulation of
nutrient-rich blood between the gut and liver is called the portal circulation. It enables the liver to remove any harmful
substances that may have been digested before the blood enters the main blood circulation around the bodythe
systemic circulation.

After a Mealthe Role of Insulin


The rise in blood glucose following a meal is detected by the pancreatic beta cells, which respond
by releasing insulin. Insulin increases the uptake and use of glucose by tissues such as skeletal
muscle and fat cells. This rise in glucose also inhibits the release of glucagon, inhibiting the production of glucose from other sources, e.g., glycogen break down (Figure 4).

1-11

Diabetes

Introduction to Diabetes

Figure 4: Changes in key hormones after a meal.


Changes in blood levels of glucose, insulin, and glucagon after a carbohyrate-rich meal (ingested at time 0 minutes).

1. Use Glucose
Once inside the cell, some of the glucose is used immediately via glycolysis. This is a central
pathway of carbohydrate metabolism because it occurs in all cells in the body, and because all
sugars can be converted into glucose and enter this pathway. During the well-fed state, the high
levels of insulin and low levels of glucagon stimulate glycolysis, which releases energy and produces carbohydrate intermediates that can be used in other metabolic pathways.
Glycolysis [http://www.ncbi.nlm.nih.gov:80/books/bv.fcgi?call=bv.View..ShowSection&rid=stryer.section.2206] in
Stryer's Biochemistry

2. Make Glycogen
Any glucose that is not used immediately is taken up by the liver and muscle where it can be converted into glycogen [http://www.ncbi.nlm.nih.gov:80/books/bv.fcgi?call=bv.View..
ShowSection&rid=stryer.chapter.2911] (glycogenesis). Insulin stimulates glycogenesis in the liver
by:

1-12

Diabetes

Introduction to Diabetes

stimulating hepatic glycogen synthetase (the enzyme that catalyzes glycogen synthesis in
the liver)

inhibiting hepatic glycogen phosphorylase (the enzyme that catalyzes glycogen breakdown
in the liver)

inhibiting glucose synthesis from other sources (inhibits gluconeogenesis)


Insulin also encourages glycogen formation in muscle, but by a different method. Here it
increases the number of glucose transporters (GLUT4) on the cell surface. This leads to a rapid
uptake of glucose that is converted into muscle glycogen.
Glycogen metabolism [http://www.ncbi.nlm.nih.gov:80/books/bv.fcgi?call=bv.View..ShowSection&rid=stryer.
chapter.2911] in Stryer's Biochemistry

3. Make Fat
When glycogen stores are fully replenished, excess glucose is converted into fat in a process
called lipogenesis. Glucose is converted into fatty acids that are stored as triglycerides (three fatty
acid molecules attached to one glycerol molecule) for storage. Insulin promotes lipogenesis by:

increasing the number of glucose transporters (GLUT4) expressed on the surface of the fat
cell, causing a rapid uptake of glucose

increasing lipoprotein lipase activity, which frees up more fatty acids for triglyceride synthesis
In addition to promoting fat synthesis, insulin also inhibits fat breakdown [http://www.ncbi.nlm.nih.
gov:80/books/bv.fcgi?call=bv.View..ShowSection&rid=stryer.chapter.3038] by inhibiting hormonesensitive lipase (an enzyme that breaks down fat stores). As a result, there are lower levels of
fatty acids in the blood stream.
Fatty acid metabolism [http://www.ncbi.nlm.nih.gov:80/books/bv.fcgi?call=bv.View..ShowSection&rid=stryer.
chapter.3038] in Stryer's Biochemistry

Insulin also has an anabolic effect on protein metabolism. It stimulates the entry of amino
acids into cells and stimulates protein production from amino acids.

Fastingthe Role of Glucagon


Fasting is defined as more than eight hours without food. The resulting fall in blood sugar levels
inhibits insulin secretion and stimulates glucagon release. Glucagon opposes many actions of
insulin. Most importantly, glucagon raises blood sugar levels by stimulating the mobilization of
glycogen stores in the liver, providing a rapid burst of glucose. In 1018 hours, the glycogen

1-13

Diabetes

Introduction to Diabetes
stores are depleted, and if fasting continues, glucagon continues to stimulate glucose production
by favoring the hepatic uptake of amino acids, the carbon skeletons of which are used to make
glucose.
In addition to low blood glucose levels, many other stimuli stimulate glucagon release including eating a protein-rich meal (the presence of amino acids in the stomach stimulates the release
of both insulin and glucagon, glucagon prevents hypoglycemia that could result from unopposed
insulin) and stress (the body anticipates an increased glucose demand in times of stress).

Starvation
The metabolic state of starvation in the USA is more commonly found in people trying to lose
weight rapidly or in those who are too unwell to eat. After a couple of days without food, the liver
will have exhausted its stores of glycogen but continues to make glucose from protein (amino
acids) and fat (glycerol).
The metabolism of fatty acids (from adipose tissue) is a major source of energy for organs
such as the liver. Fatty acids are broken down to acetyl-CoA, which is channeled into the citric
acid cycle and generates ATP. As starvation continues, the levels of acetyl-CoA increase until the
oxidative capacity of the citric acid cycle is exceeded. The liver processes these excess fatty
acids into ketone bodies (3-hydroxybutyrate) to be used by many tissues as an energy source.
The most important organ that relies on ketone production is the brain because it is unable to
metabolize fatty acids. During the first few days of starvation, the brain uses glucose as a fuel. If
starvation continues for more than two weeks, the level of circulating ketone bodies is high
enough to be used by the brain. This slows down the need for glucose production from amino
acid skeletons, thus slowing down the loss of essential proteins.

Starvation in the Midst of Plenty


Diabetes is often referred to as starvation in the midst of plenty because the intracellular levels
of glucose are low, although the extracellular levels may be extremely high.
As in starvation, type 1 diabetics use non-glucose sources of energy, such as fatty acids and
ketone bodies, in their peripheral tissues. But in contrast to the starvation state, the production of
ketone bodies can spiral out of control. Because the ketones are weak acids, they acidify the
blood. The result is the metabolic state of diabetic ketoacidosis (DKA). Hyperglycemia and
ketoacidosis are the hallmark of type 1 diabetes (Figure 5).
Hypertriglyceridemia is also seen in DKA. The liver combines triglycerol with protein to form
very low density lipoprotein (VLDL). It then releases VLDL into the blood. In diabetics, the
enzyme that normally degrades lipoproteins (lipoprotein lipase) is inhibited by the low level of
insulin and the high level of glucagon. As a result, the levels of VLDL and chylomicrons (made
from lipid from the diet) are high in DKA.

1-14

Diabetes

Introduction to Diabetes

Figure 5: Metabolic changes in diabetic ketoacidosis.


Hyperglycemia is caused by the increased production of glucose by the liver (driven by glucagon) and the decreased use
of glucose of insulin by peripheral tissues (because of the lack of insulin).

Low-Carbohydrate Diet
Low-carbohydrate diets, such as the Atkins and South Beach diets, are currently popular ways
to lose weight. Such diet plans involve restricting the type and amount of carbohydrate eaten.
One of the earliest descriptions of a low-carbohydrate diet was by William Banting in the
1860s in England. At the age of 66, Banting found success in following a carbohydrate-restricted
diet: in the course of one year, he lost 46 pounds of his initial weight of 202 pounds. His impression was that any starchy or saccharine matter tends to the disease of corpulence in advanced
life. He claimed he was never hungry and that the great charms and comfort of the system are
that its effects are palpable within a week of trial and creates a natural stimulus to persevere for a
few weeks more.
In a recent small trial, 63 obese men and women were assigned to either a low-carbohydrate
diet or a low-fat diet (1). People on both diets lost weight. The carbohydrate-restricted group initially lost weight at a faster rate, but when reviewed at the end of the year there was no significant
difference in weight loss between the two groups (1). It was found that low-carbohydrate dieters
(who were allowed unrestricted amounts of protein and fat) actually had a lower energy intake
than the low-fat diets (who were limited in their calorie intake). It may be that when carbohydrates
are restricted, weight loss is due to a lower calorie intake due to the monotony of the diet. It is
also possible that the lower calorie intake may be because of a change in peripheral or central
saiety signals, leaving people feeling more full after a meal.
A second study compared the effects of a carbohydrate-restricted diet on the risk of developing atherosclerosis (2). 132 severely obese men and women were assigned to either a lowcarbohydrate or low-fat diet. Again, after a 6-month period both groups lost weight. They became
1-15

Diabetes

Introduction to Diabetes
more sensitive to insulin, and their triglyceride (TG) levels, a type of fat that is a risk factor for
atherosclerosis, improved. However, the carbohydrate-restricted group lost more weight and
showed a greater improvement in insulin sensitivity and TG levels. After one year, the weight loss
between the two groups was similar, but the cardiogenic risk factors were improved in the lowcarbohydrate dieters, TG levels were lower, and levels of HDL cholesterol, a type of fat that
protects against atherosclerosis, were higher (3). Also, long-term sugar control, which can be
measured by checking for the amount of glycosylated hemoglobin (HbA1c), was better in people
on the low-carbohydrate diet. However, it remains unclear whether these beneficial effects would
continue after 1 year.
At present, the risks of obesity are well known, and the benefits of weight loss by traditional
low-calorie, low-fat, and high-complex carbohydrate diets are also well documented. Future
research will clarify the long-term outcomes of a low-carbohydrate diet for achieving and maintaing a healthy weight together with the effects on the heart and other systems of the body.

References
1. Foster GD, Wyatt HR, Hill JO, McGuckin BG, Brill C, Mohammed BS, Szapary PO, Rader DJ, Edman JS,
Klein S. A randomized trial of a low-carbohydrate diet for obesity. N Engl J Med 348:20822090; 2003.
(PubMed)
2. Samaha FF, Iqbal N, Seshadri P, Chicano KL, Daily DA, McGrory J, Williams T, Williams M, Gracely EJ,
Stern L. A low-carbohydrate as compared with a low-fat diet in severe obesity. N Engl J Med 348:2074
2081; 2003. (PubMed)
3. Stern L, Iqbal N, Prakash Seshadri, Chicano KL, Daily DA, McGrory J, Williams T, Williams M, Gracely
EJ, Samaha FF. The effects of low-carbohydrate versus conventional weight loss diets in severely obese
adults: one-year follow-up of a randomized trial. Ann Intern Med 140:778785; 2004. (PubMed)

Box 1: The liver buffers glucose levels.

The liver receives glucose-rich blood from the digestive tract via the portal vein (Figure 3).

The liver has a high amount of GLUT2 transporters that do not need the presence of insulin to transport glucose into the liver
cells. GLUT2 has a low affinity for glucose which enables a rapid influx of glucose when sugar levels are high. Therefore, in
the liver, levels of glucose inside and outside the cell can become equal (an equilibrium is reached).

The first step for trapping glucose inside the cell involves phosphorylation to produce glucose-6-phosphate (G6P). The liver
differs from the rest of the body in that it uses the enzyme glucokinase, rather than hexokinase. Glucokinase can produce G6P
at a faster rate and also is not inhibited by its product (this is because in the liver, G6P can be channeled into making
glycogen).

Glucose and insulin both modulate metabolic enzymes in such a way that glycogen formation is promoted. This drives forward
the process of bringing more glucose into the liver. Insulin promotes glycogen synthesis by stimulating glycogen synthetase
and inhibiting glycogen phosphorylase.

Link Roundup
The glycemic index [www.joslin.harvard.edu/education/library/glycemic_index.shtml], Joslin Diabetes Center, Boston.

1-16

Diabetes

Introduction to Diabetes

The Story of Insulin


Insulin Synthesis
The insulin-making cells of the body are called beta cells, and they are found in the pancreas
gland. These cells clump together to form the "islets of Langerhans", named for the German medical student who described them.
The synthesis of insulin begins at the translation of the insulin gene, which resides on chromosome 11. During translation, two introns are spliced out of the mRNA product, which encodes
a protein of 110 amino acids in length. This primary translation product is called preproinsulin and
is inactive. It contains a signal peptide of 24 amino acids in length, which is required for the protein to cross the cell membrane.
Once the preproinsulin reaches the endoplasmic reticulum, a protease cleaves off the signal
peptide to create proinsulin. Proinsulin consists of three domains: an amino-terminal B chain, a
carboxyl-terminal A chain, and a connecting peptide in the middle known as the C-peptide.
Within the endoplasmic reticulum, proinsulin is exposed to several specific peptidases that
remove the C-peptide and generate the mature and active form of insulin. In the Golgi apparatus,
insulin and free C-peptide are packaged into secretory granules, which accumulate in the cytoplasm of the beta cells. Exocytosis of the granules is triggered by the entry of glucose into the
beta cells. The secretion of insulin has a broad impact on metabolism.

Insulin Structure
In 1958, Frederick Sanger was awarded his first Nobel Prize [http://www.nobel.se/chemistry/
laureates/1958/] for determining the sequence of the amino acids that make up insulin. This
marked the first time that a protein had had the order of its amino acids (the primary sequence)
determined.
Insulin is composed of two chains of amino acids named chain A (21 amino acids) and chain
B (30 amino acids) that are linked together by two disulfide bridges. There is a 3rd disulfide
bridge within the A chain that links the 6th and 11th residues of the A chain together.
In most species, the length and amino acid compositions of chains A and B are similar, and
the positions of the three disulfide bonds are highly conserved. For this reason, pig insulin can be
used to replace deficient human insulin levels in diabetes patients. Today, porcine insulin has
largely been replaced by the mass production of human proinsulin by bacteria (recombinant
insulin).
Insulin molecules have a tendency to form dimers in solution, and in the presence of zinc
ions, insulin dimers associate into hexamers. Whereas monomers of insulin readily diffuse
through the blood and have a rapid effect, hexamers diffuse slowly and have a delayed onset of
action. In the design of recombinant insulin, the structure of insulin can be modified in a way that
reduces the tendency of the insulin molecule to form dimers and hexamers but that does not
interrupt binding to the insulin receptor. In this way, a range of preparations of insulin is made,
varying from short acting to long acting.

1-17

Diabetes

Introduction to Diabetes

Insulin secretion
Rising levels of glucose inside the pancreatic beta cells trigger the release of insulin:

1.

Glucose is transported into the beta cell by type 2 glucose transporters (GLUT2). Once inside, the
first step in glucose metabolism is the phosphorylation of glucose to produce glucose-6-phosphate.
This step is catalyzed by glucokinaseit is the rate-limiting step in glycolysis, and it effectively traps
glucose inside the cell.

2.

As glucose metabolism proceeds, ATP is produced in the mitochondria.

3.

The increase in the ATP:ADP ratio closes ATP-gated potassium channels in the beta cell membrane.
Positively charged potassium ions (K+) are now prevented from leaving the beta cell.

4.

The rise in positive charge inside the beta cell causes depolarization.

5.

Voltage-gated calcium channels open, allowing calcium ions (Ca2+) to flood into the cell.

6.

The increase in intracellular calcium concentration triggers the secretion of insulin via exocytosis.
There are two phases of insulin release in response to a rise in glucose. The first is an immediate release of insulin. This is attributable to the release of preformed insulin, which is stored in
secretory granules. After a short delay, there is a second, more prolonged release of newly synthesized insulin.
Once released, insulin is active for a only a brief time before it is degraded by enzymes.
Insulinase found in the liver and kidneys breaks down insulin circulating in the plasma, and as a
result, insulin has a half-life of only about 6 minutes. This short duration of action allows rapid
changes in the circulating levels of insulin.

1-18

Diabetes

Introduction to Diabetes

Insulin Receptor
The net effect of insulin binding is to trigger a cascade of phosphorylation and dephosphorylation
reactions. These actions are terminated by dephosphorylation of the insulin receptor.
Similar to the receptors for other polypeptide hormones, the receptor for insulin is embedded
in the plasma membrane and is composed of a pair of alpha subunits and a pair of beta subunits
(Figure 1). The alpha subunits are extracellular and contain the insulin-binding site. The beta
subunits span the membrane and contain the enzyme tyrosine kinase. Kinases are a group of
enzymes that phosphorylate proteins (the reverse reaction is catalyzed by a group of enzymes
called phosphatases).

Figure 1: The insulin receptor.


The insulin receptor is a tyrosine kinase receptor and is composed of a pair of alpha subunits and a pair of beta subunits.
Insulin binds to the alpha subunits and induces a conformational change that is transmitted to the beta subunits that
autophosphorylate and initiate a cascade of phosphorylation and dephosphorylation reactions.

Insulin binding to the alpha subunits induces a conformational change that is transmitted to
the beta subunits and causes them to phosphorylate themselves (autophosphorylation). A specific tyrosine of each beta subunit is phosphorylated along with other target proteins, such as
insulin receptor substrate (IRS). As these and other proteins inside the cell are phosphorylated,
this in turn alters their activity, bringing about the wide biological effects of insulin.

Insulin Action
The binding of insulin results in a wide range of actions that take place over different periods of
time.
Almost immediately, insulin promotes the uptake of glucose into many tissues that express
GLUT4 glucose transporters, such as skeletal muscle and fat. Insulin increases the the activity of
these transporters and increases their numbers by stimulating their recruitment from an intracellu1-19

Diabetes

Introduction to Diabetes
lar pool to the cell surface. Not all tissues require insulin for glucose uptake. Tissues such as liver
cells, red blood cells, the gut mucosa, the kidneys, and cells of the nervous system use a glucose
transporter that is not insulin dependent.
Over minutes to hours, insulin alters the activity of various enzymes as a result of changes in
their phosphorylation status.
Over a period of days, insulin increases the amounts of many metabolic enzymes. These
reflect an increase in gene transcription, mRNA, and enzyme synthesis.

Link Roundup
Read more about Insulin on the Bookshelf
In Human Molecular Genetics 2, see the post-translation processing [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..
ShowSection&rid=hmg.figgrp.50] and the primary structure [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..
ShowSection&rid=hmg.figgrp.51] of insulin.
In Stryer's Biochemistry, read about the receptors that contain tyrosine kinase domains [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?
call=bv.View..ShowSection&rid=stryer.section.2093#2101] | See the release of insulin [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?
call=bv.View..ShowSection&rid=stryer.figgrp.4354], insulin crystals [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..
ShowSection&rid=stryer.figgrp.288], and the synthesis of proinsulin by bacteria [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.
View..ShowSection&rid=stryer.figgrp.828].

1-20

Diabetes

Genetic Factors in Type 1 Diabetes

2. Genetic Factors in Type 1 Diabetes


Created: July 7, 2004

Type 1 diabetes is an autoimmune disorder in which the body attacks its pancreatic beta cells. The
onset of type 1 diabetes is attributed to both an inherited risk and external triggers, such as diet or
an infection. The hunt for these genetic and enviromental risk factors is on-going.
About 18 regions of the genome have been linked with influencing type 1 diabetes risk. These
regions, each of which may contain several genes, have been labeled IDDM1 to IDDM18.
The most well studied is IDDM1, which contains the HLA genes that encode immune response
proteins. Variations in HLA genes are an important genetic risk factor, but they alone do not account
for the disease and other genes are involved.
There are two other non-HLA genes which have been identified thus far. One of these non-HLA
genes, IDDM2, is the insulin gene, and the other non-HLA gene maps close to CTLA4, which has a
regulatory role in the immune response.

2-1

Diabetes

Genetic Factors in Type 1 Diabetes

IDDM1 Contains the HLA Genes


Summary
HLA genes encode molecules that are crucial to the immune system. These molecules hold small
chains of amino acids on the cell surface so that immune cells can analyze these chains. When
the immune cells find an inappropriate chain, they begin attacking. Without HLA genes, immune
cells would not find the chains of viruses, bacteria, or tumor cells. On the other hand, inheriting
certain versions (alleles) of the HLA genes increases the probability that immune cells will attack
the body's healthy cells. This is how IDDM1 contributes to the immune attack of the beta cells and
thus type 1 diabetes.

Background
The HLA region is a cluster of genes on chromosome 6. The genes encode glycoproteins that are
found on the surfaces of most cells and help the immune system to distinguish between self (its
own cells, e.g., beta cells of the pancreas) and non-self (e.g., bacteria, viruses).
Autoimmune disease results when the immune system launches an attack against the body's
tissues. The risk of developing autoimmune disease is sometimes related to the alleles of HLA
genes in the body. Type 1 diabetes is unique among these diseases in that HLA alleles may
increase the risk of diabetes, have no effect, or even be protective.
The HLA genes encode proteins called major histocompatibility complex (MHC), and there
are two main classes of MHC proteins, both of which display chains of amino acids. The chains
are called antigens, and immune cells (called T cells) analyze them. Class I MHC present chains
from inside cells, whereas MHC class 2 present chains from outside the cells. If T cells bind to the
chain presented on an MHC, the T cell immediately orchestrates powerful attacks by the body's
other immune cells. Ideally, the body only contains T cells that bind to chains from infectious
organisms (viruses, bacteria, etc.) and tumor cells. Healthy development largely achieves this
ideal. The alternative is found in autoimmune diseases such as diabetes: T cells bind to chains
from the body's healthy cells.
There are many different alleles of the HLA genes, leading to many different variants of MHC
proteins and allowing a variety of chains to be presented to cells. The inheritance of particular
HLA alleles can account for over half of the genetic risk of developing type 1 diabetes (1). The
genes encoding class II MHC proteins are most strongly linked with diabetes, and these genes
are called HLA-DR, HLA-DQ, and HLA-DP.
In the general population, only half of the people inherit a copy (allele) of DR gene called DR3
and DR4, and less than 3% of the people have two alleles. However, in type 1 diabetes at least
one allele of DR3 or DR4 is found in 95% of Caucasians, and individuals with both DR3 and DR4
are particularly susceptible to type 1 diabetes (2). Conversely, the DR2 allele is protective (3).
Similar to the DR gene, certain alleles of the DQ gene are risk factors for developing the disease, whereas other alleles of DQ are protective. There is also a tendency for people who inherit
DR3 or DR4 to inherit DQ, which adds to their genetic risk of developing diabetes. Conversely,
the protective alleles of DR and DQ tend to be inherited together. These tendencies have complicated the study of the effects of individual HLA-DR or HLA-DQ genes.
2-2

Diabetes

Genetic Factors in Type 1 Diabetes

IDDM1 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on IDDM1 and diabetes, search PubMed.
Associations between HLA alleles and diabetes began to be documented in the 1970s when
serological markers were used. This association was later confirmed with genome-wide scans.
The IDDM1 locus contains many diabetes susceptibility genes, and it remains difficult to identify the specific risk alleles because of linkage disequilibrium; certain alleles tend to be coinherited with other alleles, making it difficult to distinguish between the effects of either on
diabetes susceptibility.
Fine mapping of these regions suggests that the two alleles DQB1 and DRB1 are the most
important (4). Alleles in the DQB1 gene are often tightly associated with alleles in the DRB1 gene,
and variants of both or either allele may confer an increased risk of diabetes.
Sequences in the DQB1 gene that code for an amino acid other than aspartic acid at position
57 (non-ASP57) are highly associated with type 1 diabetes (5). Crystal structures suggest that
loss of aspartic acid at this position creates an "oxyanion hole". This may be occupied by the T
cell during the interaction between HLA and the T-cell receptor (6). The diabetes risk of nonASP57 is further increased when the haplotype also contains the DRB1*0401 allele, suggesting
the possible existence of at least two separate loci of susceptibility (7).
One of the protective HLA haplotypes is DQA1*0102, DQB1*0602. Aproximately 20% of
Americans and Europeans have this haplotype, whereas less than 1% of children with type 1 diabetes do (8).
A well-known marker for type 1 diabetes is the presence of islet cell autoantibodies. However,
even in the presence of islet cell autoantibodies, the haplotype DQA1*0102, DQB1*0602 has a
protective effect. But once the diabetes disease process begins, the mechanism that protected
these individuals from diabetes is lost, suggesting that inheriting these alleles does not prevent
diabetes but may somehow delay or arrest the progression of the disease (9).

References
1. Todd JA, Bell JI, McDevitt HO et al. HLA-DQ beta gene contributes to susceptibility and resistance to
insulin-dependent diabetes mellitus. Nature 329:599604; 1987. (PubMed)
2. Wolf E, Spencer KM, Cudworth AG et al. The genetic susceptibility to type 1 (insulin-dependent)
diabetes: analysis of the HLA-DR association. Diabetologia 24:224230; 1983. (PubMed)
3. Platz P, Jakobsen BK, Morling N et al. HLA-D and -DR antigens in genetic analysis of insulin dependent
diabetes mellitus. Diabetologia 21:108115; 1981. (PubMed)
4. Herr M, Dudbridge F, Zavattari P et al. Evaluation of fine mapping strategies for a multifactorial disease
locus: systematic linkage and association analysis of IDDM1 in the HLA region on chromosome 6p21.
Hum Mol Genet 9:12911301; 2000. (PubMed)
5. Todd JA, Bell JI, McDevitt HO. HLA-DQ beta gene contributes to susceptibility and resistance to insulindependent diabetes mellitus. Nature 329:559604; 1987. (PubMed)
6. Corper AL, Stratmann T, Apostolopoulos V et al. A structural framework for deciphering the link between
I-Ag7 and autoimmune diabetes. Science 288:505511; 2000. (PubMed)
7. Florez JC, Hirschhorn J, Altshuler D et al. The inherited basis of diabetes mellitus: implications for the
genetic analysis of complex traits. Annu Rev Genomics Hum Genet 4:257291; 2003. (PubMed)

2-3

Diabetes

Genetic Factors in Type 1 Diabetes

8. Redondo MJ, Kawasaki E, Mulgrew CL. DR- and DQ-associated protection from type 1A diabetes:
comparison of DRB1*1401 and DQA1*0102-DQB1*0602*. J Clin Endocrinol Metab 85:37933797; 2000.
(PubMed)
9. Greenbaum CJ, Schatz DA, Cuthbertson D et al. Islet cell antibody-positive relatives with human
leukocyte antigen DQA1*0102, DQB1*0602: identification by the Diabetes Prevention Trial-type 1. J Clin
Endocrinol Metab 85:12551260; 2000. (PubMed)

Link Roundup
Live Searches
Diabetes and IDDM1 in PubMed | PubMed Central | Books

Background Information
IDDM1 in OMIM
Type 1 diabetes in Genes and Disease [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..ShowSection&rid=gnd.section.137]

Molecular Biology
IDDM1 in Entrez Gene | Map Viewer

2-4

Diabetes

Genetic Factors in Type 1 Diabetes

IDDM2 Contains the Insulin Gene (INS)


Summary
The IDDM2 locus contains the insulin gene (INS). Mutations of INS cause a rare form of diabetes
that is similar to MODY (Maturity Onset Diabetes in the Young). Other variations of the insulin
gene (variable number tandem repeats and SNPs) may play a role in susceptibility to type 1 and
type 2 diabetes.

Background
Insulin is a hormone that has a wide range of effects on metabolism. Its overall action is to
encourage the body to store energy rather than use it, e.g., insulin favors the storage of glucose
as glycogen or fat as opposed to breaking down glucose to release ATP. For a summary of the
actions of insulin, see the Physiology and Biochemistry of Sugar Regulation.
Insulin is composed of two distinct polypeptide chains, chain A and chain B, which are linked
by disulfide bonds. Many proteins that contain subunits, such as hemoglobin, are the products of
several genes. However, insulin is the product of one gene, INS.
INS actually encodes an inactive precursor called preproinsulin. Preproinsulin is processed
into proinsulin by removal of a signaling peptide; however, proinsulin is also inactive. The final
processing step involves removal of a C-peptide (a connecting peptide that links chain A to chain
B), and this process produces the mature and active form of insulin. For further information, see
The Story of Insulin.

Molecular Information
Several species, including the rat, mouse, and some species of fish, have two insulin genes. In
contrast, in humans there is a single insulin gene that is located on chromosome 11 (Figure 1). It
has three exons (coding regions) that span about 2,200 bases (see evidence [http://www.ncbi.
nlm.nih.gov/sutils/evv.cgi?taxid=9606&contig=NT_009237.16&gene=INS&graphiconly=TRUE]).
Exon 2 encodes the B chain, along with the signal peptide and part of the C-peptide found in the
precursors of insulin. Exon 3 encodes the A chain and the remainder of the C-peptide.
C-peptide is secreted in equal amounts to insulin, but it has long been thought that it has no
biological role. However, in diabetic rats C-peptide has been shown to reduce the dysfunction of
blood vessels and the nervous system that is common in diabetes (1). C-peptide contains the
greatest variation among species, whereas regions of insulin that bind to the insulin receptor are
highly conserved.
Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=3630&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the INS gene, none (at the time of writing) of which cause non-synonymous amino
acid changes in the mature protein (see the allelic variants that are known to be associated with
disease).

2-5

Diabetes

Genetic Factors in Type 1 Diabetes


A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4557671&cut=100&org=1]
using human proinsulin precursor as a query finds proteins in 107 different species, which are all
metazoans apart from three plants and one bacterium. However, potential true homologous
genes have thus far been identified only in the mouse and rat.

Figure 1: Location of INS on the human genome.


INS maps to chromosome 11, approximately between 21442148 kilobases (kb). Click

on the figure or here for a

current and interactive view of the location of INS in the human genome.
Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of INS may fluctuate; therefore, the live Web site may not appear exactly as in this
figure.

IDDM2 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on INS and diabetes, search PubMed.
The IDDM2 locus contributes about 10% toward type 1 diabetes susceptibility (2). The "risk
area" of this locus is localized to a region flanking the insulin gene that contains a short sequence
of DNA that is repeated many times (3, 4). The repeats are found 0.5 kb upstream from the site
where transcription of INS begins. Because the repeated sequences follow one behind the other
(in tandem) and because the number of repeats varies between individuals, this phenomenon is
called variable number tandem repeats (VNTRs).
There are three classes of VNTR in the insulin gene (5):

Class I has alleles that range from 26 to 63 repeat units.


Class II has alleles that average around 80 repeat units.
Class III has alleles ranging from 141 to 209 repeat units.

2-6

Diabetes

Genetic Factors in Type 1 Diabetes


The class I VNTRs are most common in Caucasians, with around 70% of alleles being in the
range of 30-44 repeats, and nearly all other alleles are longer than 110 repeats (class III). The
intermediate lengths (class II) are rare.
The class of VNTR is associated with susceptibility to type 1 diabetes. Short class I alleles
are associated with a higher risk of developing type 1 diabetes, whereas the longer class III alleles are protective. The presence of at least one class III allele is associated with a 3-fold reduction in the risk of type 1 diabetes, compared with common I/I homozygote genotype (6).
Because the VNTR occurs in a non-coding region, its influence on diabetes risk cannot be
attributed to an alteration of the protein sequence. Instead, the VNTR probably affects the transcription of the insulin gene in some way. Indeed in the pancreas, the class III alleles are associated with 15-30% lower INS mRNA.
In contrast, class III alleles are associated with higher levels of INS mRNA in the thymus. This
gland has an important role in training the immune system in the developing embryo. Immature T
cells are presented with chains of amino acids, such as insulin, and T cells that form a response
to them (and thus are autoreactive) are deleted. Because the longer VNTRs cause more insulin
to be produced in the thymus, the detection and deletion of autoreactive T cells may be more efficient. This improved immune tolerance to insulin would lessen the risk of a future onset of type 1
diabetes caused by anti-insulin antibodies.

References
1. Ido Y, Vindigni A, Chang K et al. Prevention of vascular and neural dysfunction in diabetic rats by Cpeptide. Science 277:563566; 1997. (PubMed)
2. Bennett ST, Lucassen AM, Gough SC et al. Susceptibility to human type 1 diabetes at IDDM2 is
determined by tandem repeat variation at the insulin gene minisatellite locus. Nat Genet 9:284292;
1995. (PubMed)
3. Owerbach D, Gabbay KH. Localization of a type I diabetes susceptibility locus to the variable tandem
repeat region flanking the insulin gene. Diabetes 42:17081714; 1993. (PubMed)
4. Cox NJ, Wapelhorst B, Morrison VA et al. Seven regions of the genome show evidence of linkage to type
1 diabetes in a consensus analysis of 767 multiplex families. Am J Hum Genet 69:820830; 2001.
(PubMed)
5. Bennett ST, Todd JA. Human type 1 diabetes and the insulin gene: principles of mapping polygenes.
Annu Rev Genet 30:343370; 1996. (PubMed)
6. Vafiadis P, Ounissi-Benkalha H, Palumbo M et al. Class III alleles of the variable number of tandem
repeat insulin polymorphism associated with silencing of thymic insulin predispose to type 1 diabetes. J
Clin Endocrinol Metab 86:37053710; 2001. (PubMed)

Link Roundup
Live Searches
Diabetes and insulin in PubMed | PubMed Central | Books

Background Information
Insulin in OMIM

Molecular Biology

2-7

Diabetes

Genetic Factors in Type 1 Diabetes

IDDM2 in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?


taxid=9606&contig=NT_009237.16&gene=INS&graphiconly=TRUE] | Map Viewer | Insulin Domain [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4557671] | SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=3630&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseR] | Allelic Variants | BLink [http://www.ncbi.nlm.nih.gov/
sutils/blink.cgi?pid=4557671&cut=100&org=1] | HomoloGene

2-8

Diabetes

Genetic Factors in Type 1 Diabetes

Other Type 1 Diabetes Susceptibility Loci: IDDM3IDDM18


IDDM1 (containing the HLA system) and IDDM2 (containing the insulin gene) were both originally
identified by investigating the suspected genes, HLA genes and INS, respectively, using casecontrol studies. The remaining type 1 diabetes susceptibility loci, IDDM3IDDM18, were mainly
discovered by genome scan linkage studies, e.g., looking for linkage between regions of the
genome and disease in affected sib-pairs.
The IDDM loci are found on several different chromosomes and contain many genes, many
of which have now been identified. Some of these genes are suspected to play a role in susceptibility to type 1 diabetes, and they are discussed below.

IDDM3
No diabetes susceptibility genes have been identified in the IDDM3 locus, which is found on
chromosome 15.

IDDM4
Several potential candidate genes lie near the IDDM4 locus on chromosome 11. These include
ZFM1 (zinc finger protein 162), which encodes a transcription factor found in the pancreas, and
FADD (Fas-associated death protein). The transmission of the "cell death" signal involves the
interaction between FAS and FADD, and in type 1 diabetes, the apoptosis of pancreatic beta cells
may involve the FADD. Apoptosis of the beta cell may be triggered by the binding of Fas
(expressed on the beta cell) with Fas ligand (expressed on the cytotoxic T cell) (1).
2-9

Diabetes

Genetic Factors in Type 1 Diabetes


Other candidate genes in this region include LRP5, which encodes a novel transmembrane
protein that is similar to receptors belonging to the low-density lipoprotein family (2).

IDDM5
The region of chromosome 6 that contains the IDDM5 locus includes the SOD2 gene, which
encodes mitochondrial superoxide dismutase. SOD2 metabolizes harmful oxygen free radicals,
which are intermediates in many biological reactions, and converts them into less reactive and
less harmful molecules. There is some evidence that free oxygen radicals may play a role in the
destruction of beta cells. Enzymes such as SOD2 may therefore offer protection against type 1
diabetes, and genetic variants of SOD2 may increase susceptibility to disease (1).

IDDM6
Several candidate diabetes susceptibility genes have been identified in the IDDM6 locus. They
include a gene associated with colorectal cancer (DCC) that may be linked with autoimmune disease, a gene that encodes a zinc finger DNA binding domain (ZNF236) that may be linked with
diabetic kidney disease, and a molecule that opposes apoptosis (bcl-2) (1).

IDDM7
Within the IDDM7 locus on chromosome 2 are several candidate diabetes risk genes. One is
NEUROD1 (3), a transcription factor that is expressed widely in the developing brain and pancreas. NEUROD1 regulates the transcription of the insulin gene, and in addition to its association
with type 1 diabetes, variants of this gene have been implicated in susceptibility to type 2 diabetes; a mutation of this gene causes MODY6.
Other genes located within the IDDM7 locus include IGRP (islet-specific glucose-6phosphatase catalytic subunit-related protein), which encodes the beta cell-specific version of the
enzyme glucose-6-phosphatase. IGRP has emerged as a major target of cell-mediated autoimmunity in type 1 diabetes (4).
Many other candidate genes (interleukin-1 gene cluster, HOXD8, GAD1, GALNT3) in this
region have been investigated but none of these genes have been shown to be associated with
type 1 diabetes (1).

IDDM8
No diabetes susceptibility genes have been identified in the IDDM8 locus, which is found on
chromosome 6.

IDDM9
The symbol IDDM9 has not yet been approved.

IDDM10
The gene GAD2 is closely linked to the region of chromosome 10 designated as IDDM10. Glutamic acid decarboxylase (GAD) catalyzes formation of the neurotransmitter GABA. Targeting of
this enzyme by autoantibodies has been implicated in the pathogenesis of stiff-man syndrome
2-10

Diabetes

Genetic Factors in Type 1 Diabetes


and type 1 diabetes (5, 6). Both diseases feature insulin deficiency, but stiff-man syndrome also
bears the features of an autoimmune attack against the central nervous system, characterized by
painful muscle spasms and increasing stiffness of axial muscles. The difference between stiffman syndrome and type 1 diabetes may be because GAD is expressed in two different isoforms:
one is expressed in the central nervous system, and the other is in the beta cells (7). The nature
of the immune attack against these two isoforms also appears to be different (8).
The GAD2 gene encodes GAD65, and this protein contains a 24-amino acid segment that is
similar to an amino acid sequence found in the Coxsackie virus, a suspected environmental trigger for the onset of type 1 diabetes. Autoimmunity in IDDM may thus arise by "molecular mimicry"
between GAD and a viral polypeptide (9). However, evidence of cross-reactivity has not been
demonstrated in immune cells from patients with diabetes.
Autoantibodies against GAD have been found in patients who have had preclinical type 1
diabetes (10). In the type 1 diabetes mouse, the expression of GAD by beta cells is required for
the development of autoimmune diabetes. Complete suppression of beta-cell GAD expression
blocks the generation of diabetogenic T cells, leading to the theory that modulation of GAD might
have therapeutic value in type I diabetes (11).

IDDM11
One candidate gene in the IDDM11 locus is the ENSA gene, which encodes alpha-endosulphine.
This protein is thought to be an endogenous regulator of the beta cell potassium channel (KATP
channel).
The KATP channels co-ordinate a rise in blood glucose with insulin secretion. As glucose
levels rise, the corresponding rise in ATP shuts the channel, leading to a change in membrane
polarity. Voltage-sensitive calcium channels flip open, allowing Ca2+ ions to enter into the beta
cells, triggering exocytosis of insulin. The KATP channel pore is encoded by the KIR gene, and
the channel is regulated by the sulfonylurea receptor encoded by the ABCC8 gene.
Recombinant alpha-endosulphine has been shown to inhibit the binding of the diabetes drug
sulfonylurea to its receptor, to reduce the flow of K+ through the KATP channel, and to stimulate
insulin secretion (12).
Another candidate gene in this region is the SEL1L gene. It is a negative regulator of the
Notch signalling pathway which controls the differentiation of pancreatic endocrine cells (13).

IDDM12
Several candidate genes have been located in the IDDM12 locus, and the strongest candidates
encode co-stimulatory receptors on the T cell. When the T cell is presented with a chain of amino
acids, its T-cell receptor binds to the HLA molecules that are presenting the chains. For the T cell
to become fully activated, there is additional signaling between co-stimulatory receptors and corresponding ligands on the antigen-presenting cell. These co-stimulatory receptors are encoded
by the candidate genes for type 1 diabetes susceptibility CTLA4, CD28, and ICOS.
Read more about the role of CTLA4 in type 1 diabetes.

2-11

Diabetes

Genetic Factors in Type 1 Diabetes

IDDM13
Several IDDM13 candidate genes have been investigated, but variants of these genes have yet
to be associated with type 1 diabetes.

IDDM14
The symbol IDDM14 has not yet been approved.

IDDM15
The IDDM15 locus has been linked with type 1 diabetes, and mutations near this region are
associated with a rare form of diabetes called transient neonatal diabetes (14).

IDDM16
One of the candidate genes in the IDDM16 locus is the immunoglobulin heavy chain.
Immunoglobulins (antibodies) have a central role in the immune response against foreign antigens and in error can also attack self antigens, resulting in autoimmune disease. Immunoglobulins are known to interact with HLA molecules, variants of which are associated with diabetes
protection or susceptibility (IDDM1 contains the HLA genes). Immunoglobulins are composed of
two heavy chains and two light chains, and the IDDM16 locus contains the gene that encodes the
heavy chain. Genetically controlled differences in the immunoglobulin heavy chain may affect an
individual's immune response to self antigens and thus alter the risk of developing autoimmune
diseases such as type 1 diabetes (1).

IDDM17
The IDDM17 locus was discovered to be linked to type 1 diabetes, but the candidate gene(s) is
not yet known. The FAS gene maps to this genomic area, but it has been excluded as a possible
diabetes susceptibility gene.

IDDM18
A candidate diabetes susceptibility gene in the IDDM18 locus is ILB12. This gene encodes a
subunit of IL-12p40, a signaling molecule secreted by white blood cells. In animal models, IL-12
plays an important role in the induction of diabetes. In humans, variation in IL-12p40 production
may influence the reactivity of T cells and initiate or protect against autoimmune diseases such as
type 1 diabetes (15, 6).

References
1. Pociot F, McDermott MF et al. Genetics of type 1 diabetes mellitus. Genes Immun 3:235249; 2002.
(PubMed)
2. Hey PJ, Twells RC, Phillips MS et al. Cloning of a novel member of the low-density lipoprotein receptor
family. Gene 216:103111; 1998. (PubMed)
3. Tamimi R, Steingrimsson E, Copeland NG. The NEUROD gene maps to human chromosome 2q32 and
mouse chromosome 2. Genomics 34:418421; 1996. (PubMed)
2-12

Diabetes

Genetic Factors in Type 1 Diabetes

4. Hutton JC, Eisenbarth GS et al. A pancreatic beta-cell-specific homolog of glucose-6-phosphatase


emerges as a major target of cell-mediated autoimmunity in diabetes. Proc Natl Acad Sci U S A
100:86268628; 2003. (PubMed)
5. Baekkeskov S, Aanstoot HJ, Christgau S et al. Identification of the 64K autoantigen in insulin-dependent
diabetes as the GABA-synthesizing enzyme glutamic acid decarboxylase. Nature 347:151156; 1990.
(PubMed)
6. von Boehmer H, Sarukhan A et al. GAD, a single autoantigen for diabetes. Science 284:11351137;
1999. (PubMed)
7. Kaufman DL, Erlander MG, Clare-Salzler M et al. Autoimmunity to two forms of glutamate decarboxylase
in insulin-dependent diabetes mellitus. J Clin Invest 89:283292; 1992. (PubMed)
8. Lohmann T, Hawa M, Leslie RD et al. Immune reactivity to glutamic acid decarboxylase 65 in stiffman
syndrome and type 1 diabetes mellitus. Lancet 356:3135; 2000. (PubMed)
9. Albert LJ, Inman RD et al. Molecular mimicry and autoimmunity. N Engl J Med 341:20682074; 1999.
(PubMed)
10. De Aizpurua HJ, Wilson YM, Harrison LC. Glutamic acid decarboxylase autoantibodies in preclinical
insulin-dependent diabetes. Proc Natl Acad Sci U S A 89:98419845; 1992. (PubMed)
11. Yoon JW, Yoon CS, Lim HW et al. Control of autoimmune diabetes in NOD mice by GAD expression or
suppression in beta cells. Science 284:11831187; 1999. (PubMed)
12. Heron L, Virsolvy A, Apiou F et al. Isolation, characterization, and chromosomal localization of the human
ENSA gene that encodes alpha-endosulfine, a regulator of beta-cell K(ATP) channels. Diabetes
48:18731876; 1999. (PubMed)
13. Apelqvist A, Li H, Sommer L et al. Notch signalling controls pancreatic cell differentiation. Nature
400:877881; 1999. (PubMed)
14. Arima T, Drewell RA, Arney KL et al. A conserved imprinting control region at the HYMAI/ZAC domain is
implicated in transient neonatal diabetes mellitus. Hum Mol Genet 10:14751483; 2001. (PubMed)
15. Adorini L. Interleukin 12 and autoimmune diabetes. Nat Genet 27:131132; 2001. (PubMed)
16. Morahan G, Ymer SI, Cancilla MR et al. Linkage disequilibrium of a type 1 diabetes susceptibility locus
with a regulatory IL12B allele. Nat Genet 27:218221; 2001. (PubMed)

Link Roundup
Background information in Entrez Gene
IDDM3

IDDM4

IDDM5

IDDM6

IDDM7

IDDM8

IDDM9

IDDM10

IDDM11

IDDM12

IDDM13

IDDM14

IDDM15

IDDM16

IDDM17

IDDM18

Background information in OMIM


IDDM3

IDDM4

IDDM5

IDDM6

IDDM7

IDDM8

[IDDM9]

[IDDM10]

IDDM11

IDDM12

IDDM15 [IDDM16]

IDDM13 [IDDM14]
IDDM17

IDDM18

2-13

Diabetes

Genetic Factors in Type 1 Diabetes

An Inhibitor of the Immune Response (CTLA4)


Summary
Immune cells are continually analyzing small chains of amino acids to detect infectious agents or
tumor cells. When a foreign chain is found, the immune cells become activated and begin to
attack. The CTLA4 gene encodes a molecule that hinders the activation of immune cells. The
region of the chromosome that contains CTLA4 has been linked with susceptibility to many
autoimmune diseases including type 1 diabetes.

Nomenclature
Official gene name: cytotoxic T-lymphocyte-associated protein 4
Official gene symbol: CTLA4
Alias: CD152

Background
One of the steps in mounting an immune response involves an interaction between two cells. The
first cell, called the antigen-presenting cell (APC), displays small chains of amino acids (antigens)
on its surface, and they present these antigens to the second type of cell, T cells. Once the T cell
has analyzed the antigen, it can either become activated and launch an immune attack or be
deactivated. In a healthy immune system, T cells become activated only to foreign antigens, such
as fragments from bacteria or viruses. If the T cells become activated in response to self antigens, autoimmune diseases such as diabetes results.
Optimal activation of the T cell requires a two-way interaction between the T-cell receptor and
the antigen (the first signal) and between co-stimulatory receptors on the surface of the T cell with
co-stimulatory ligands expressed by APCs (the second signal). Failure of the T cell to receive a
second signal can lead to its deactivation.
One of the co-stimulatory molecules on the T cell is called cytotoxic T lymphocyte-associated
antigen 4 (CTLA4). CTLA4 has a negative regulatory effect on the immune system because it
down-regulates T-cell activation by interfering with the second signal. Mice with a targeted disruption of the CTLA4 gene develop a fatal disorder characterized by massive lymphocyte proliferation (1).
Unlike other co-stimulator receptors on the T cell, CTLA4 is only expressed when the T cell
has been activated after antigen presentation. Because it is only expressed in activated T cells,
and because it down regulates the function of T cells, it is likely that CTLA4 has a role in guarding
against autoimmunity (2). Loss of this gene may result in activated T cells attacking self antigens.
Indeed, genetic variants of CTLA4 have been linked with autoimmune disorders such as autoimmune hypothyroid disease, Graves' disease (3), systemic lupus erythematosus (SLE) (4), celiac
disease (5), and type 1 diabetes (6-8).

2-14

Diabetes

Genetic Factors in Type 1 Diabetes

Molecular Information
CTL4A maps to IDDM12 on chromosome 2 (Figure 1), and the equivalent area of the mouse
genome has been linked to type 1 diabetes in the non-obese diabetic (NOD) mouse (9). It has
four exons (coding regions) that span around 7,000 bases (see evidence [http://www.ncbi.nlm.
nih.gov/sutils/evv.cgi?taxid=9606&contig=NT_005403.14&gene=CTLA4&graphiconly=TRUE]).
The gene encodes a protein of 223 amino acids.
The CTLA4 protein contains an immunoglobulin V-like domain (view domain [http://www.ncbi.
nlm.nih.gov:80/Structure/cdd/cddsrv.cgi?uid=3897]), a transmembrane region, and a putative
cytoplasmic region that is identical to the mouse CTLA4 protein. This conservation of the cytoplasmic region between species suggests that it has an important role in the functioning of CTLA4
(10).
Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=1493&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the CTLA4 gene, two (at the time of writing) of which cause non-synonymous amino
acid changes in the mature protein (Figure 2). None of these variants have yet to be associated
with disease (see known allelic variants).
A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=21361212&cut=100&org=1]
using human CTLA4 as a query finds proteins in 19 different species, which are all metazoans
(multicellular). However, potential true homologous genes have thus far been identified only in the
mouse and rat.

Figure 1: Location of CTLA4 on the human genome.


CTLA4 maps to chromosome 2, between approximately 204,930204,945 kilobases (kb). Click

on the image or here

for a current and interactive view of the location of CTLA4 in the human genome.

2-15

Diabetes

Genetic Factors in Type 1 Diabetes


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of CTLA4 may fluctuate. Therefore, the live Web site may not appear exactly as in this
figure.

Figure 2: SNP positions of CTLA4 mapped to the 3D structure of human Ctla-4B7-2 complex.
The figure shows the positions of non-synonymous amino acid changes (green residues) caused by SNPs in the coding
sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

CTLA4 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on CTLA4 and diabetes, search PubMed.
There are several SNPs in the 3' untranslated region of the CTLA4 sequence that have been
implicated in determining the risk for several common autoimmune disorders, including type 1
diabetes.
One of the SNPs, termed CT60, encodes a genotype that is either protective (A/A) or predisposes (G/G) toward autoimmune disease. The disease susceptibility G allele is common, being
found in 50% of individuals without autoimmune disorders, and is more common in individuals
with Graves' disease (63%). The G/G haplotype correlated with lower production of the soluble
alternative splice from of CTLA4 (sCTLA4) compared with the protective A/A haplotype. This
reduction of sCTLA4 levels could lead to reduced blocking of signals between T cells and APCs,
leading to increased activation of T cells. This allele was also associated with type 1 diabetes but
the effect was small (11).

References
2-16

Diabetes

Genetic Factors in Type 1 Diabetes

1. Waterhouse P, Penninger JM, Timms E et al. Lymphoproliferative disorders with early lethality in mice
deficient in Ctla-4. Science 270:985988; 1995. (PubMed)
2. Kristiansen OP, Larsen ZM, Pociot F. CTLA-4 in autoimmune diseases--a general susceptibility gene to
autoimmunity? Genes Immun 1:170184; 2000. (PubMed)
3. Vaidya B, Imrie H, Perros P et al. The cytotoxic T lymphocyte antigen-4 is a major Graves' disease locus.
Hum Mol Genet 8:11951199; 1999. (PubMed)
4. Hudson LL, Rocca K, Song YW et al. CTLA-4 gene polymorphisms in systemic lupus erythematosus: a
highly significant association with a determinant in the promoter region. Hum Genet 111:452455; 2002.
(PubMed)
5. Djilali-Saiah I, Schmitz J, Harfouch-Hammoud E et al. CTLA-4 gene polymorphism is associated with
predisposition to coeliac disease. Gut 43:187189; 1998. (PubMed)
6. Nistico L, Buzzetti R, Pritchard LE et al. The CTLA-4 gene region of chromosome 2q33 is linked to, and
associated with, type 1 diabetes. Belgian Diabetes Registry. Hum Mol Genet 5:10751080; 1996.
(PubMed)
7. Marron MP, Raffel LJ, Garchon HJ et al. Insulin-dependent diabetes mellitus (IDDM) is associated with
CTLA4 polymorphisms in multiple ethnic groups. Hum Mol Genet 6:12751282; 1997. (PubMed)
8. Lohmueller KE, Pearce CL, Pike M et al. Meta-analysis of genetic association studies supports a
contribution of common variants to susceptibility to common disease. Nat Genet 33:177182; 2003.
(PubMed)
9. Lamhamedi-Cherradi SE, Boulard O, Gonzalez C et al. Further mapping of the Idd5.1 locus for
autoimmune diabetes in NOD mice. Diabetes 50:28742878; 2001. (PubMed)
10. Dariavach P, Mattei MG, Golstein P et al. Human Ig superfamily CTLA-4 gene: chromosomal localization
and identity of protein sequence between murine and human CTLA-4 cytoplasmic domains. Eur J
Immunol 18:19011915; 1988. (PubMed)
11. Ueda H, Howson JM, Esposito L et al. Association of the T-cell regulatory gene CTLA4 with susceptibility
to autoimmune disease. Nature 423:506511; 2003. (PubMed)

Link Roundup
Live Searches
Diabetes and CTLA4 in PubMed | PubMed Central | Books

Background Information
CTLA4 in OMIM

Molecular Biology
CTLA4 in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_005403.14&gene=CTLA4&graphiconly=TRUE] | Map Viewer | Immunoglobulin V-type Domain [http://www.
ncbi.nlm.nih.gov:80/Structure/cdd/cddsrv.cgi?uid=3897] | SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?locusId=1493&view
+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?
pid=21361212&cut=100&org=1] | HomoloGene

2-17

Diabetes

Genetic Factors in Type 2 Diabetes

3. Genetic Factors in Type 2 Diabetes


Created: July 7, 2004

Type 2 diabetes has been loosely defined as "adult onset" diabetes, although as diabetes becomes
more common throughout the world, cases of type 2 diabetes are being observed in younger
people. It is increasingly common in children.
In determining the risk of developing diabetes, environmental factors such as food intake and
exercise play an important role. The majority of individuals with type 2 diabetes are either
overweight or obese. Inherited factors are also important, but the genes involved remain poorly
defined.
In rare forms of diabetes, mutations of one gene can result in disease. However, in type 2
diabetes, many genes are thought to be involved. "Diabetes genes" may show only a subtle
variation in the gene sequence, and these variations may be extremely common. The difficulty lies
in linking such common gene variations, known as single nucleotide polymorphisms (SNPs), with an
increased risk of developing diabetes.
One method of finding the diabetes susceptibility genes is by whole-genome linkage studies.
The entire genome of affected family members is scanned, and the families are followed over
several generations and/or large numbers of affected sibling-pairs are studied. Associations
between parts of the genome and the risk of developing diabetes are looked for. To date only two
genes, calpain 10 (CAPN10) and hepatocyte nuclear factor 4 alpha (HNF4A), have been identified
by this method.

3-1

Diabetes

Genetic Factors in Type 2 Diabetes

The Sulfonylurea Receptor (ABCC8)


Summary
Sulfonylureas are a class of drugs used to lower blood glucose in the treatment of type 2 diabetes. These drugs interact with the sulfonylurea receptor of pancreatic beta cells and stimulate
insulin release. The sulfonylurea receptor is encoded by the ABCC8 gene, and genetic variation
of ABCC8 may impair the release of insulin.

Nomenclature
Official name: ATP-binding cassette, sub-family C, member 8
Official gene symbol: ABCC8
Alias: sulfonylurea receptor, SUR, SUR1

Background
The protein encoded by the ABCC8 gene is a member of the ATP-binding cassette transporters.
These proteins use energy in the form of ATP to drive the transport of various molecules across
cell membranes. ABCC8 belongs to a subfamily of transporters that contains the chloride channel
that is mutated in cystic fibrosis (CFTR) and also the proteins that are involved in multi-drug resistance.
Read more: The Human ATP-Binding Cassette (ABC) Transporter Superfamily [www.ncbi.nlm.nih.gov:80/books/bv.
fcgi?call=bv.View..ShowTOC&rid=mono_001.TOC&depth=2]

The ABCC8 protein is also known as the sulfonylurea urea receptor (SUR). SUR is one of the
proteins that composes the ATP-sensitive potassium channel (KATP channel) found in the pancreas (1). The other protein, called Kir6.2, forms the core of the channel and is encoded by the
KCNJ11 gene. KATP channels play a central role in glucose-induced insulin secretion by linking
signals derived from glucose metabolism (a rise in ATP) to membrane depolarization (due to
KATP channels closing) and the secretion of insulin.
The activity of the KATP channel regulates the release of insulin. The sulfonylureas are drugs
that can modulate KATP channel activity and are used in the treatment of type 2 diabetes. By
binding to SUR, they inhibit the channel and stimulate the release of insulin. This leads to a lowering of blood glucose levels.
Further information on sulfonylureas from MEDLINEplus [www.nlm.nih.gov/medlineplus/druginfo/uspdi/202742.html]

The activity of the KATP channel is also modulated by the subtype of SUR (SUR, also known
as SUR1, is encoded by ABCC8; or SUR2A and SUR2B, which are encoded by ABCC9). In the
pancreas, most KATP channels are thought to be a complex of four SUR1 proteins and four
Kir6.2 proteins.

3-2

Diabetes

Genetic Factors in Type 2 Diabetes


Mutations in either ABCC8 or KCNJ11 can result in up-regulated insulin secretion, a condition
termed familial persistent hyperinsulinemic hypoglycemia of infancy (PHHI) (2-4). Genetic variation in ABCC8 has also been implicated in the impaired release of insulin that is seen in type 2
diabetes.

Molecular Information
ABC genes are found in many different eukaryotic species and are highly conserved between
species, indicating that many of these genes existed early in eukaryotic evolution. A BLAST
search [www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4507317&cut=100&org=1] using human
ABCC8 as a query finds proteins in 30 different species, which include multicellular organisms
(metazoans), fungi, and plants. Potential true homologous genes have been identified in the
mouse and rat.
By fluorescence in situ hybridization, it was found that the ABBC8 gene maps to the short
arm of chromosome 11 (Figure 1) (5). It has 41 exons (coding regions) that span over 84,000
bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_009237.16&gene=ABCC8&graphiconly=TRUE]).

3-3

Diabetes

Genetic Factors in Type 2 Diabetes

Figure 1: Location of ABCC8 on the human genome.


ABCC8 maps to chromosome 11, between approximately 17,37017,470 kilobases (kb). Click

or here for a current and

interactive view of the location of ABCC8 in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of ABCC8 may fluctuate. The live Web site may, therefore, not appear exactly as in this
figure.

The ABC transporter proteins, such as ABCC8, typically contain two ATP-binding domains
and two transmembrane domains (view domains [www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi?
INPUT_TYPE=precalc&SEQUENCE=4507317]).
The ATP-binding domains are also known as nucleotide binding folds (NBFs), and mutations
in either NBF1 or NBF2 can lead to PHHI (6). This suggests that both NBF regions of the SUR
are needed for the normal regulation of KATP channel activity.
As found in all proteins that bind ATP, the nucleotide binding domains of the ABC family of
proteins contain characteristic motifs called Walker A and B. The Walker A motif contains a lysine
residue that is critical for activating the KATP channel. When this lysine residue is mutated in
NBF1, but not NBF2, the KATP channel can no longer be activated (7). In addition, ABC genes
also contain a signature C motif.

3-4

Diabetes

Genetic Factors in Type 2 Diabetes


The transmembrane domains contain 611 membrane spanning helices, and the ABCC protein contains 6. These helices provide the protein with specificity for the molecule they transport
across the membrane.
Several single nucleotide polymorphisms (SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=6833&view=view&chooseRs=coding&.cgifields=chooseRs]) have been found within the
ABCC8 gene. Usually SNPs linked with disease occur within the coding regions (exons) of the
genes, and they result in a non-synonymous amino acid change. In ABCC8, there are seven such
SNPs (at the time of writing) that cause a switch of amino acids in the mature protein (Figure 2).
However, one of the SNPs of the ABCC8 gene that has been linked with diabetes (R1273R) does
not cause an amino acid change (see below).

Figure 2: SNP positions of ABCC8 mapped to the 3D structure of a multidrug resistance ABC transporter
homolog in Vibrio cholera.
The figure shows the positions of non-synonymous amino acid changes (green residues) caused by SNPs in the coding
sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

ABCC8 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on ABCC8 and diabetes, search PubMed.
3-5

Diabetes

Genetic Factors in Type 2 Diabetes


The two genes that encode the KATP channel, ABBC8 and KCNJ11, reside adjacent to one
another on chromosome 11. A variant of ABCC8, called A1369S, is in almost complete linkage
disequilibrium with a variant of KCNJ11 called E23K. This means that from the genetic evidence,
it is difficult to determine whether it is the A1369S variant or the E23K variant that predisposes to
type 2 diabetes (8).
A mutation in ABCC8 was observed to cause an extremely rare form of diabetes, autosomal
dominant diabetes, in a Finnish family (9). The switch of glutamate to lysine at residue 1506
(E1506K) in the SUR1 protein caused a congenital hyperinsulinemia. The mutation reduced the
activity of KATP channels, increasing insulin secretion. By early adulthood, the ability of the beta
cells to secrete adequate amounts of insulin was exhausted, leading to diabetes (10).
A silent variant in exon 31 of the ABCC8 gene has been associated with high concentrations
of insulin in non-diabetic Mexican Americans. The codon AGG is mutated to AGA, but this still
codes for the residue arginine (R1273R). The normal and mutant alleles were called G and A,
respectively. Among non-diabetics, those who were homozygous for the mutant allele (AA genotype) had higher levels of insulin when fasting, compared with heterozygotes (AG) and normal
wild-type (GG). Because type 2 diabetes is more common in Mexican Americans than in the general US population, it has been proposed that individuals with the AA genotype are at a higher
risk of diabetes because of an over-secretion of insulin (11).
Two common polymorphisms of the ABCC8 gene (exon 16-3t/c and exon 18 T/C) have been
variably associated with type 2 diabetes. However, a recent large case control study in Britain
revealed that these ABCC8 variants did not appear to be associated with diabetes (12).

References
1. Inagaki N, Gonoi T, Clement JP et al. Reconstitution of IKATP: an inward rectifier subunit plus the
sulfonylurea receptor. Science 270:11661170; 1995. (PubMed)
2. Thomas PM, Cote GC, Wohllk N et al. Mutations in the sulfonylurea receptor gene in familial persistent
hyperinsulinemic hypoglycemia of infancy. Science 268:426429; 1995. (PubMed)
3. Nestorowicz A, Inagaki N, Gonoi T et al. A nonsense mutation in the inward rectifier potassium channel
gene, Kir6.2, is associated with familial hyperinsulinism. Diabetes 46:17431748; 1997. (PubMed)
4. Nestorowicz A, Glaser B, Wilson BA et al. Genetic heterogeneity in familial hyperinsulinism. Hum Mol
Genet 7:11191128; 1998. (PubMed)
5. Thomas PM, Cote GJ, Hallman DM et al. Homozygosity mapping, to chromosome 11p, of the gene for
familial persistent hyperinsulinemic hypoglycemia of infancy. Am J Hum Genet 56:416421; 1995.
(PubMed)
6. Thomas PM, Wohllk N, Huang E et al. Inactivation of the first nucleotide-binding fold of the sulfonylurea
receptor, and familial persistent hyperinsulinemic hypoglycemia of infancy. Am J Hum Genet 59:510
518; 1996. (PubMed)
7. Gribble FM, Tucker SJ, Ashcroft FM et al. The essential role of the Walker A motifs of SUR1 in K-ATP
channel activation by Mg-ADP and diazoxide. Embo J 16:11451152; 1997. (PubMed)
8. Florez JC, Burtt N, de Bakker PI et al. Haplotype structure and genotype-phenotype correlations of the
sulfonylurea receptor and the islet ATP-sensitive potassium channel gene region. Diabetes 53:1360
1368; 2004. (PubMed)
9. Huopio H, Reimann F, Ashfield R et al. Dominantly inherited hyperinsulinism caused by a mutation in the
sulfonylurea receptor type 1. Embo J 106:897906; 2000. (PubMed)
3-6

Diabetes

Genetic Factors in Type 2 Diabetes

10. Huopio H, Otonkoski T, Vauhkonen I et al. A new subtype of autosomal dominant diabetes attributable to
a mutation in the gene for sulfonylurea receptor 1. Lancet 361:301307; 2003. (PubMed)
11. Goksel DL, Fischbach K, Duggirala R et al. Variant in sulfonylurea receptor-1 gene is associated with
high insulin concentrations in non-diabetic Mexican Americans: SUR-1 gene variant and
hyperinsulinemia. Hum Genet 103:280285; 1998. (PubMed)
12. Gloyn AL, Weedon MN, Owen KR et al. Large-scale association studies of variants in genes encoding
the pancreatic beta-cell KATP channel subunits Kir6.2 (KCNJ11) and SUR1 (ABCC8) confirm that the
KCNJ11 E23K variant is associated with type 2 diabetes. Hum Genet 52:568572; 2003. (PubMed)

Link Roundup
Live Searches
Diabetes and ABCC8 in PubMed | PubMed Central | Books

Background Information
ABCC8 in OMIM
The Human ATP-Binding Cassette (ABC) Transporter Superfamily [www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..
ShowTOC&rid=mono_001.TOC&depth=2] on the Bookshelf

Molecular Biology
ABCC8 in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_009237.16&gene=ABCC8&graphiconly=TRUE] | Map Viewer | Domains [www.ncbi.nlm.nih.gov/Structure/
cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4507317]: Transmembrane region 1, ATPase 1, Transmembrane domain 2,
ATPase 2 | SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?locusId=6833&view=view&chooseRs=coding&.cgifields=chooseRs] |
BLink [www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4507317&cut=100&org=1] | HomoloGene

3-7

Diabetes

Genetic Factors in Type 2 Diabetes

The Calpain 10 Enzyme (CAPN10)


Summary
Calpain 10 is a calcium-activated enzyme that breaks down proteins. Variation in the non-coding
region of the CAPN10 gene is associated with a threefold increased risk of type 2 diabetes in
Mexican Americans. A genetic variant of CAPN10 may alter insulin secretion, insulin action, and
the production of glucose by the liver.

Nomenclature
Official gene name: calpain 10
Official gene symbol: CAPN10
Alias: calcium-activated neutral protease

Background
The discovery of CAPN10 marks the first time that screening the entire genome led to the identification of a gene linked to a common and genetically complex disease such as type 2 diabetes.
In 1996, a link was made between a region of chromosome 2 and an increased risk of diabetes in Mexican Americans in Texas, USA (1). The region on the chromosome, located near the
end of the long arm of chromosome 2, was named NIDDM1 (Non-Insulin-Dependent Diabetes
Mellitus 1). To pinpoint a gene that conferred risk, new statistical techniques were used, and the
region was sequenced in greater and greater detail. Four years later, a new diabetes susceptibility gene, CAPN10, was discovered (2).
The new gene encoded the enzyme calpain 10, a member of the calpain family of cysteine
proteases. These enzymes are activated by calcium and regulate the functions of other proteins
by cleaving pieces off, leaving the altered protein either more or less active. In this way, they regulate biochemical pathways and are involved in intracellular signaling pathways, cell proliferation,
and differentiation.
For a detailed description of the action of cysteine proteases, visit Stryer's Biochemistry [www.ncbi.nlm.nih.gov/
books/bv.fcgi?call=bv.View..ShowSection&rid=stryer.section.1170#1192]

Calpains are found in all human cells, and 14 members of the calpain family are now known,
many of which are associated with human disease (3). Calpains 1 and 2 are implicated in causing
injury to the brain after a stroke and also have been linked to the pathology seen in Alzheimer's
disease. Calpain 3 is mainly found in the muscle, and mutations cause limb-girdle muscular dystrophy . Mutations of calpains in the worm Caenorhabditis elegans affect sexual development (4),
and mutations of a calpain-like gene in the fly cause a degeneration of parts of the nervous system (5).
Calpain 10 was an unexpected find in the search for a putative diabetes susceptibility gene.
Its link with diabetes is complex; susceptibility is not attributable to a single variation but to several variations of DNA that interact to either increase, decrease, or have no effect on the risk of
developing diabetes. In Mexican Americans, it is thought that the highest risk combination of
3-8

Diabetes

Genetic Factors in Type 2 Diabetes


these variations (termed 112/121, see below) results in a population-attributable risk of 14%, i.e.,
14% of Mexican Americans who have diabetes would not have diabetes if they did not have the
high-risk genetic variant CAPN10.
Calpain 10 is an atypical member of the calpain family, and its biological role is unknown.
Because CAPN10 mRNA is expressed in the pancreas, muscle, and liver, its role in diabetes may
involve insulin secretion, insulin action, and the production of glucose by the liver (2).

Molecular Information
Calpains are found in all human cells and are found throughout the animal kingdom. A BLAST
search [www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=13186302&cut=100&org=1] using human
CAPN10 as a query finds proteins in 18 different species, all of which are multicellular organisms
(metazoans). However, potential true homologous genes have been identified only in the mouse.
The calpains consist of a large catalytic subunit and a small regulatory subunit. The large
subunit contains four domains (IIV), and the catalytic center of the cysteine protease is located
in domain II. Domain I is the N-terminal domain that is processed when the enzyme is activated,
domain III is a linker domain, and the C-terminal domain IV binds calcium and resembles the calcium-binding protein, calmodulin. Domain IV also has characteristic EF-hand motifs (6).
Calpain 10 is an atypical calpain in that it lacks the calmodulin-like domain IV and instead has
a divergent C-terminal domain, domain T. Calpains 5 and 6 also have a domain T, and together
they form a subfamily of calpains (6).
The CAPN10 gene maps to chromosome 2 (Figure 1). It has 13 exons (coding regions) that
span about 30,000 bases (see evidence [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_005416.11&gene=CAPN10&graphiconly=TRUE]) (2). Alternate splicing
of the gene creates at least eight transcripts (named isoform a through isoform h), which in turn
encode proteins ranging from 138 to 672 amino acids in length.

3-9

Diabetes

Genetic Factors in Type 2 Diabetes

Figure 1: Location of CAPN10 on the human genome.


CAPN10 maps to chromosome 2, between approximately 241,840241,880 kilobases (kb). Click

or here for a current

and interactive view of the location of CAPN10 in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of CAPN10 may fluctuate. The live Web site may, therefore, not appear exactly as in
this figure.

Calpain isoform a is the most abdundant isoform. It lacks exons 8, 14, and 15 but remains the
longest transcript and encodes a protein of 672 amino acids that is found in all tissues, with the
highest levels being found in the heart.
Although the structure of CAPN10 has not yet been solved, mapping the CAPN10 sequence
to the crystal structure of human m-caplain (also known as calpain 2 or CAPN2) gives a good
estimate of structure. The domains span from residues 1672 (see domains [http://www.ncbi.nlm.
nih.gov/Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=13186302]).
Many single nucleotide polymorphisms (SNPs) have been found within the CAPN10 gene. A
surprise finding was that the SNPs linked with type 2 diabetes in the Mexican Americans were
located in the non-coding regions (introns) of the gene. Usually, SNPs linked with disease occur
within the coding regions (exons) of the genes, and they result in a non-synonymous amino acid
change. In calpain isoform a, there are four such SNPs (at the time of writing) that cause a switch
of amino acids in the mature protein (Figure 2). However, the SNPs of the CAPN10 gene that

3-10

Diabetes

Genetic Factors in Type 2 Diabetes


have been linked with diabetes are located in introns (introns 3, 6, and 19), and because introns
are not transcribed, they do not directly cause an amino acid change. Instead, it is proposed that
SNPs in CAPN10 introns may alter risk by affecting the transciptional regulation of calpain 10 (2).

Figure 2: SNP positions of CAPN10 mapped to the 3D structure of human m-calpain.


The figure shows the positions of two of the non-synonymous amino acid changes (green residues) caused by SNPs in
the coding sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

CAPN10 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on CAPN10 and diabetes, search PubMed.
Over 10% of Mexican Americans are affected by type 2 diabetes. By studying generations of
Mexican Americans in Star County, Texas, it was found that SNPs in introns of the calpain 10
gene were associated with an increased susceptibility to diabetes.
Intron 3 of the CAPN10 gene contained SNP-43 (also called UCSNP-43 for University of
Chicago SNP-43) in which adenine had been switched to guanine. The high-risk genotype is
SNP-43 G/G. SNPs were also found in intron 6 (SNP-19) and intron 13 (SNP-63). Together,
these three SNPs interact to affect the risk of diabetes.
With two different versions of the gene at three distinct sites, there are eight possible combinations, but only four combinations of alleles are commonly found. At each SNP site, the allele
was labeled "1" and "2". The most common combination in Mexican Americans was 112 on one
chromosome and 121 on the other. This 112/121 comination was associated with a three-fold
increased risk of diabetes. The high-risk combination also increased the risk of developing dia3-11

Diabetes

Genetic Factors in Type 2 Diabetes


betes in Northern Europeans, but because the at-risk genotype is less common, it has less of a
role in determining susceptibility in Europeans. The genotype 112/111 had no effect on the risk of
developing diabetes, and the 112/221 combination actually decreased the risk (2).
Similar to the Mexican Americans, the Pima Indians of Arizona have a high prevalence of
type 2 diabetes. However, in a study among the Pima Indians, no association was found between
the high-risk genotype SNP-43 G/G and an increased prevalence of diabetes, although G/G individuals did have reduced expression of CAPN10 mRNA and showed signs of insulin resistance,
which may increase susceptibility to diabetes (7).
In Europe, CAPN10 appears to contribute less to type 2 diabetes. In Britain, there was no
association between SNP-43, SNP-19, and SNP-63 and diabetes, but it is possible that SNPs at
other sites in the calpain gene may increase the type 2 diabetes risk (8). In a large study in Scandinavians, no association was found between these three SNPs and diabetes (9). In Asia, genetic
variation in CAPN10 did not contribute to diabetes in Japan (10) or in the Samoans of Polynesia
(11), although variations of CAPN10 may play a role in the risk of diabetes in the Chinese (12,
13).
The biological function of calpain 10 has remained unknown for several years; however, a
recent study suggests that calpain 10 may have a critical role in beta cell survival. There are several intracellular stores of calcium; one of these stores contains calcium channels that are sensitive to ryanodine (RyR2). In the beta cell, blocking RyR2 initiates a newly discovered pathway
that activates CAPN10 and triggers cell death (14).

References
1. Hanis CL et al. A genome-wide search for human non-insulin-dependent (type 2) diabetes genes reveals
a major susceptibility locus on chromosome 2. Nat Genet 13:161166; 1996. (PubMed)
2. Horikawa Y et al. Genetic variation in the gene encoding calpain-10 is associated with type 2 diabetes
mellitus. Nat Genet 26:163175; 2000. (PubMed)
3. Huang Y, Wang KK. The calpain family and human disease. Trends Mol Med 7:355362; 2001.
(PubMed)
4. Barnes TM, Hodgkin J. The tra-3 sex determination gene of Caenorhabditis elegans encodes a member
of the calpain regulatory protease family. Embo J 15:44774484; 1996. (PubMed)
5. Delaney SJ et al. Molecular cloning and analysis of small optic lobes, a structural brain gene of
Drosophila melanogaster. Proc Natl Acad Sci U S A 88:72147218; 1991. (PubMed)
6. Dear N et al. A new subfamily of vertebrate calpains lacking a calmodulin-like domain: implications for
calpain regulation and evolution. Genomics 45:175184; 1997. (PubMed)
7. Baier LJ et al. A calpain-10 gene polymorphism is associated with reduced muscle mRNA levels and
insulin resistance. J Clin Invest 106:R69R73; 2000. (PubMed)
8. Evans JC et al. Studies of association between the gene for calpain-10 and type 2 diabetes mellitus in the
United Kingdom. Am J Hum Genet 69:544552; 2001. (PubMed)
9. Rasmussen SK et al. Variants within the calpain-10 gene on chromosome 2q37 (NIDDM1) and
relationships to type 2 diabetes, insulin resistance, and impaired acute insulin secretion among
Scandinavian Caucasians. Diabetes 51:35613567; 2002. (PubMed)
10. Horikawa Y et al. Genetic variations in calpain-10 gene are not a major factor in the occurrence of type 2
diabetes in Japanese. J Clin Endocrinol Metab 88:244247; 2003. (PubMed)

3-12

Diabetes

Genetic Factors in Type 2 Diabetes

11. Tsai HJ et al. Type 2 diabetes and three calpain-10 gene polymorphisms in Samoans: no evidence of
association. Am J Hum Genet 69:12361244; 2001. (PubMed)
12. Hong J et al. Relationship between calpain-10 gene polymorphism, hypertension and plasma glucose.
Zhonghua Nei Ke Za Zhi 41:370373; 2002. (PubMed)
13. Wang Y et al. The UCSNP44 variation of calpain 10 gene on NIDDM1 locus and its impact on plasma
glucose levels in type 2 diabetic patients. Zhonghua Yi Xue Za Zhi 82:613616; 2002. (PubMed)
14. Johnson JD, Han Z, Otani K et al. RyR2 and Calpain-10 delineate a novel apoptosis pathway in
pancreatic islets. J Biol Chem 279:2479424802; 2004. (PubMed)

Link Roundup
Live Searches
Diabetes and CAPN10 in PubMed | PubMed Central | Books

Background Information
Calpain 10 in OMIM

Molecular Biology
Calpain 10 in Entrez Gene | Evidence Viewer [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_005416.11&gene=CAPN10&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=13186302] | SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=11132&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [www.ncbi.nlm.nih.gov/sutils/
blink.cgi?pid=13186302&cut=100&org=1] | HomoloGene

3-13

Diabetes

Genetic Factors in Type 2 Diabetes

The Glucagon Receptor (GCGR)


Summary
Glucagon is a key hormone in the regulation of glucose levels. As such, the GCGR gene which
encodes its receptor is a candidate diabetes susceptibility gene. A mutant glucagon receptor has
been associated with diabetes.

Background
Glucagon, like insulin, is a peptide hormone that is secreted by the pancreas. Unlike insulin,
glucagon acts to promote the mobilization of fuels, particularly glucose. See Figure 1 for a summary of the actions of insulin and glucagon on their three main target tissues, the liver, muscle,
and adipose tissue.

Figure 1: Action of glucagon and insulin on the liver, muscle, and adipose tissue.
Glucagon and insulin have opposing effects on their three main target organs.

Glucagon acts on the liver to stimulate glucose production, either from the breakdown of
glycogen (glycogenolysis) or by the fresh production of glucose from the carbon skeletons of
amino acids (glucogenogenesis).
In addition to raising blood glucose levels, glucagon also stimulates the formation of ketone
bodies. Glucagon does this by stimulating the release of free fatty acids (FFAs) from adipose tissue, and when these enter the liver, glucagon directs their metabolic fate. Instead of being used
to make triglycerides (a storage form of lipid), the FFAs are shunted toward beta-oxidation and
the formation of ketoacids, which can be used by several organs as an alternative fuel to glucose.

3-14

Diabetes

Genetic Factors in Type 2 Diabetes


The glucagon receptor is a member of the 7-transmembrane receptor (7TM) family. These
receptors have also been called the Serpentine receptors because of the way the single polypeptide chain of the receptor "snakes" back and forth across the cell membrane. This diverse family
of receptors not only responds to hormones such as glucagon but also transmits many other signals including light, smell, and neurotransmitters.
Read more about 7TM receptors in Stryer's Biochemistry [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..
ShowSection&rid=stryer.section.2055]

When glucagon outside of the cell binds to the glucagon receptor, it induces a conformational
change in the receptor that can be detected inside the cell. The activated receptor stimulates
intracellular cAMP production, which in turn stimulates protein kinase A (PKA) to phosphorylate
many target proteins, including two enzymes that lead to the breakdown of glycogen and inhibit
further glycogen synthesis.

Molecular Information
The GCGR gene maps to chromosome 17 (Map Viewer). It has 14 exons (coding regions) that
span around 10,700 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_079568.1&gene=GCGR&graphiconly=TRUE]). The 7TM receptor
encoded by this gene is also known as a G protein coupled receptor because a G protein (guanyl
nucleotide-binding protein) conveys an intermediate step in signal transduction.
In the unactivated state, the G protein coupled to the glucagon receptor exists as a heterotrimer consisting of , , and subunits. The subunit (G) binds GDP. Many different G
proteins exist, and for the glucagon receptor, because the activated G subunit stimulates adenylate cyclase, the receptor is classified as a Gs receptor ("s" for stimulatory effect of the subunit).
Binding of glucagon initiates the exchange of Gs-bound GDP for GTP. The G protein then
dissociates into G and a G dimer; the latter transmits the signal that the receptor has bound
its ligand. A single glucagon receptor can activate many G proteins, leading to a greatly amplified
signal.
The signal from an occupied glucagon receptor is "switched off" when the G hydrolyzes the
bound GTP back to GDP, and the GDP-bound G can now reassociate with G to reform the
heterotrimeric protein.
A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4503947&cut=100&org=1]
using human GCGR as a query finds proteins in 35 different species, which are all metazoans
(multicellular organisms). However, potential true homologous genes have thus far been identified only in the mouse, rat, and fruit fly.

GCGR and Diabetes: Digest of Recent Articles


For a more complete list of research articles on GCGR and diabetes, search PubMed.
A missense mutation in the glucagon receptor gene has been associated with decreased tissue sensitivity to glucagon and type 2 diabetes (1, 2).

3-15

Diabetes

Genetic Factors in Type 2 Diabetes


A Gly40Ser mutation of the receptor gene was linked with type 2 diabetes in French and Sardinian patients. The mutation occurs in exon 2 and disrupts an extracellular region of the receptor
(1).
When the signaling properties of the Gly40Ser mutant receptor were examined, it was found
that the mutant receptor bound glucagon with a threefold lower affinity compared with the wildtype receptor. In addition, the production of cAMP in response to glucagon was decreased, and
glucagon-stimulated insulin secretion was also decreased (2). Decreased insulin secretion was
observed in Gly40Ser carriers in a Brazilian population, but the mutant receptor was not associated with type 2 diabetes (3).

References
1. Hager J, Hansen L, Vaisse C et al. A missense mutation in the glucagon receptor gene is associated with
non-insulin-dependent diabetes mellitus. Nat Genet 9:299304; 1995. (PubMed)
2. Hansen LH, Abrahamsen N, Hager J et al. The Gly40Ser mutation in the human glucagon receptor gene
associated with NIDDM results in a receptor with reduced sensitivity to glucagon. Diabetes 45:725730;
1996. (PubMed)
3. Shiota D, Kasamatsu T, Dib SA et al. Role of the Gly40Ser mutation in the glucagon receptor gene in
Brazilian patients with type 2 diabetes mellitus. Pancreas 24:386390; 2002. (PubMed)

Link Roundup
Live Searches
Diabetes and GCGR in PubMed | PubMed Central | Books

Background Information
The glucagon receptor in OMIM

Molecular Biology
Gene name in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_079568.1&gene=GCGR&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4503947]: Hormone receptor, 7 Transmembrane domain | SNPs
[http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?locusId=2642&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic
Variants | BLink [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4503947&cut=100&org=1] | HomoloGene

3-16

Diabetes

Genetic Factors in Type 2 Diabetes

The Enzyme Glucokinase (GCK)


Summary
Glucokinase, encoded by the GCK gene speeds up glucose metabolism and acts as a "glucose
sensor" in the beta cell. Mutant glucokinase causes a rare form of diabetes and may also play a
role in type 2 diabetes.

Nomenclature
Official gene name: Glucokinase
Official gene symbol: GCK
Alias: GK, hexokinase IV, HK4, Maturity Onset Diabetes in the Young type 2, MODY2

Background
The enzyme glucokinase catalyzes glucose metabolism in the liver and in the pancreatic beta
cell. Glucokinase traps glucose inside the cell by catalyzing its phosphorylation to produce glucose-6-phosphate. This is the first and rate-limiting step in glycosis, a pathway that produces
energy in the form of ATP from glucose.
Glucokinase (also called hexokinase IV) differs from the other hexokinases that are found in
other tissues. First, glucokinase has a lower affinity for glucose. This allows other organs such as
the brain and muscles to have first call on glucose when their supply is limited. A second feature
is that glucokinase is not inhibited by its product, glucose-6-phosphate. This lack of negative
feedback inhibition enables hepatic glucokinase to remain active while glucose is abundant,
ensuring that the liver can continue removing glucose from the blood ensuring that no glucose
goes to waste.
Glucokinase is proposed to be an important "glucose sensor" in the following way. The rate of
glucose metabolism is determined by the rate of glucose phosphorylation, which is catalyzed by
glucokinase in the liver and pancreas. The liver and pancreas also express glucose transporter-2
(GLUT2), an insulin-independent cellular protein that mediates the transport of glucose into cells.
The capacity of GLUT2 to transport glucose is very high, facilitating rapid equilibrium between
extracellular and intracellular glucose. Thus, in effect, the extracellular glucose concentrations are
sensed by glucokinase.
By catalyzing the rate-limiting step of glucose metabolism in the liver, glucokinase enables
the liver to buffer the rise in glucose that takes place after a meal. In the pancreas, glucokinase is
the glucose sensor for insulin release. The threshold for glucose-stimulated insulin release is
about 5 mmol/l (1). Mutations of GCK that alter this threshold manifest as three different syndromes and highlight the importance of GCK in glucose homeostasis and diabetes:
1. Activating mutations lower the beta cell threshold for insulin release to as low as 1.5
mmol/l of glucose, leading to an increase in insulin release. This manifests as persistent
hyperinsulinemic hypoglycemia of infancy (PHHI) (2).

3-17

Diabetes

Genetic Factors in Type 2 Diabetes

2. Inactivating mutations raise the beta cell threshold for insulin release. If two alleles
altered by inactivating mutations are inherited, the level of glucose needed to stimulate
insulin release from the pancreas is extremely high. Affected individuals present with diabetes at birth (permanent neonatal diabetes) (3, 4).
3. Maturity onset diabetes in the young, type 2 (MODY2) is caused by inheriting one allele
that has been altered by an inactivating mutation. This partial inactivation leads to an
increase in glucose-stimulated insulin release to about 7 mmol/l. This causes a mild
hyperglycemia that is present at birth but often is only detected in later life (5).
Because of its role as a glucose sensor, GCK is a candidate susceptibility gene for type 2
diabetes.

Molecular Information
The hexokinase family consists of several enzymes that are all evolutionarily related. In vertebrates, there are four hexokinases named I to IV. Glucokinase is a distinct member of this family.
A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4503951&cut=100&org=1]
using human GCK as a query finds proteins in 46 different species, which range from metazoa
(multicellular organisms), fungi, plants, and other eukaryotes. Potential true homologous genes
have thus far been identified in the mouse, rat, fly, mosquito, nematode worm, and the plant,
mouse-ear cress.
The GCK gene maps to chromosome 7 (Figure 1). It has 12 exons (coding regions) that span
around 46,000 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_007819.14&gene=GCK&graphiconly=TRUE]). There are three GCK transcript variants that differ in their first exon, and their expression is tissue specific. One isoform
predominates in the pancreatic beta cells; the other two isoforms are found in the liver. The glucokinase enzyme is found in the outer membrane compartment of mitochondria in these tissues.

3-18

Diabetes

Genetic Factors in Type 2 Diabetes

Figure 1: Location of GCK on the human genome.


GCK maps to chromosome 7, between approximately 43,00044,000 kilobases (kb). Click

or here for a current and

interactive view of the location of GCK in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of GCK may fluctuate. The live Web site may, therefore, not appear exactly as in this
figure.

GCK and Diabetes: Digest of Recent Articles


For a more complete list of research articles on GCK and diabetes, search PubMed.
Mutations known to activate glucokinase include a switch of amino acids at position 455 (Val
to Met) (6) and position 456 (Ala to Val) (7). These mutations are all clustered in one area of the
glucokinase structure that is remote from the substrate binding site, and this region has been
termed the allosteric activator site. Because naturally occurring mutations at this site are associated with an increase in insulin release, the search began for pharmacological allosteric activators that could increase GCK activity, increase the release of insulin, and be used in the treatment
of diabetes.
The metabolic changes that take place in type 2 diabetes include impaired insulin secretion
from the pancreas and increased glucose production by the liver. In turn, studies of gluckokinase
activators (GKAs) appear to show a dual mechanism of action: they increase the release of
insulin from the pancreas and stimulate the use of glucose in the liver, thus making GKAs ideal
candidates for diabetes therapy (8, 9).

References
1. Matschinsky FM. Regulation of pancreatic beta-cell glucokinase: from basics to therapeutics. Diabetes 51
(3):S394S404; 2002. (PubMed)
3-19

Diabetes

Genetic Factors in Type 2 Diabetes

2. Gloyn AL. Glucokinase (GCK) mutations in hyper- and hypoglycemia: maturity-onset diabetes of the
young, permanent neonatal diabetes, and hyperinsulinemia of infancy. Hum Mutat 22:353362; 2003.
(PubMed)
3. Njolstad PR, Sovik O, Cuesta-Munoz A et al. Neonatal diabetes mellitus due to complete glucokinase
deficiency. N Engl J Med 344:15881592; 2001. (PubMed)
4. Njolstad PR, Sagen JV, Bjorkhaug L et al. Permanent neonatal diabetes caused by glucokinase
deficiency: inborn error of the glucose-insulin signaling pathway. Diabetes 52:28542860; 2003.
(PubMed)
5. Katagiri H, Asano T, Ishihara H et al. Nonsense mutation of glucokinase gene in late-onset non-insulindependent diabetes mellitus. Lancet 340:13161317; 1992. (PubMed)
6. Glaser B, Kesavan P, Heyman M et al. Familial hyperinsulinism caused by an activating glucokinase
mutation. N Engl J Med 338:226230; 1998. (PubMed)
7. Christesen HB, Jacobsen BB, Odili S et al. The second activating glucokinase mutation (A456V):
implications for glucose homeostasis and diabetes therapy. Diabetes 51:12401246; 2002. (PubMed)
8. Grimsby J, Sarabu R, Corbett WL et al. Allosteric activators of glucokinase: potential role in diabetes
therapy. Science 301:370373; 2003. (PubMed)
9. Brocklehurst KJ, Payne VA, Davies RA et al. Stimulation of hepatocyte glucose metabolism by novel
small molecule glucokinase activators. Diabetes 53:535541; 2004. (PubMed)

Link Roundup
Live Searches
Diabetes and GCK in PubMed | PubMed Central | Books

Background Information
Glucokinase in OMIM

Molecular Biology
Gene name in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_007819.14&gene=GCK&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=15967159] | SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=2645&view+rs+=view+rs+&chooseRs=all&.cgifields=chooseRs] | Allelic Variants | BLink [http://www.ncbi.nlm.nih.gov/sutils/
blink.cgi?pid=4503951&cut=100&org=1] | HomoloGene

3-20

Diabetes

Genetic Factors in Type 2 Diabetes

The Glucose Transporter GLUT2


Summary
The GLUT2 gene encodes a glucose transporter which regulates the entry of glucose into pancreatic beta cells and is thought to be a "glucose sensor". Mutations of GLUT2 cause a rare
genetic syndrome that disrupts blood glucose regulation. Common variants of GLUT2 may be
linked with type 2 diabetes.

Nomenclature
Official gene name: solute carrier family 2 (facilitated glucose transporter), member 2
Official gene symbol: SLC2A2
Alias: glucose transporter 2, GLUT2

Background
The family of glucose transporter proteins transports glucose across plasma membranes. The
proteins are facilitative transporters carrying glucose across membranes without requiring energy.
Each member of the family has its own specialized function and tissue distribution.
For a summary of the family of glucose transporters, named GLUT1 to GLUT5, read Stryer's Biochemistry [http://
www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..ShowSection&rid=stryer.section.2245#2255].

Beta cells in the pancreas express GLUT2. The transporter has a low affinity for glucose,
requiring a high concentration of glucose to achieve half its maximal rate of transport (high Km).
This enables the beta cell to only begin the insulin secretion cascade during times of plenty. By
regulating the entry of glucose into the beta cell, it has been proposed that GLUT2 is an important
"glucose sensor" (1, 2).
GLUT2 is also found in the liver and transports glucose both into and out of liver cells. It
allows the liver to buffer changes in glucose concentration after a meal. GLUT2 has a high capacity for glucose, so once the glucose levels have risen, the liver can quickly remove large amounts
of glucose from the blood into the liver cells.
Mutations in the gene that encodes GLUT2, called SLC2A2, cause an extremely rare genetic
syndrome called Fanconi-Bickel syndrome (FBS) (3). FBS was named for the scientists who first
described the condition in 1949, and since then over 100 cases have been observed. Affected
individuals may exhibit hyperglycemia in the fed state and/or hypoglycemia during fasting.
Hyperglycemia in FBS may involve inappropriate glucose uptake into the liver and this hyperglycemia could be further enhanced by inappropriate insulin secretion by the pancreas. The rise
in liver intracellular glucose may inhibit the breakdown of glycogen, leading to low levels of glucose in the blood and increased levels of glycogen stored in the liver. A transport defect of the
kidneys also leads to the loss of glucose in the urine which further worsens the hypoglycemia.
Whereas mutations of SLC2A2 cause a rare genetic syndrome that disrupts glucose regulation, it it possible that variations of this gene may also be involved in type 2 diabetes, a common
form of diabetes.
3-21

Diabetes

Genetic Factors in Type 2 Diabetes

Molecular Information
The GLUT2 gene maps to chromosome 3 (Figure 1). It has 11 exons (coding regions) that span
about 31,400 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_005612.14&gene=SLC2A2&graphiconly=TRUE]).
Members of the glucose transporter family each consist of a single polypeptide of about 500
residues long. This peptide crosses the membrane several times, and there are typically 12
transmembrane segments. The translocation of glucose across the membrane is thought to
involve inward-facing and outward-facing glucose binding sites that are located in transmembrane
segments 9, 10, and 11 of the GLUT protein. Mutations of these regions severely impair glucose
transport (4-6).
A highly conservered motif, (R)XGRR, is found not only in the facilitative glucose transporters
but also in the sugar transport superfamily (7). Mutations that disrupt this conserved sequence
result in the Fanconi-Bickel syndrome (3).
Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=6514&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the GLUT2 gene, four (at the time of writing) of which cause non-synonymous amino
acid changes in the mature protein. One of these variants (T110I) may be associated with type 2
diabetes status (8), and different variants have been associated with the Fanconi-Bickel syndrome (view known allelic variants).
The problem of fuel transport appears to have been solved early in evolutionary terms. A
BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4557851&cut=100&org=1] using
human SLC2A2 as a query finds proteins in 55 different species that include sugar transporters in
organisms as diverse as bacteria, fungi, plants, and multicellular species (metazoa). However,
potential true homologous genes have thus far been identified only in the mouse and rat.

3-22

Diabetes

Genetic Factors in Type 2 Diabetes

Figure 1: Location of SLC2A2 on the human genome.


SLC2A2 maps to chromosome 3, between approximately 172,030172,070 kilobases (kb). Click

or here for a current

and interactive view of the location of SLC2A2 in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of SLC2A2 may fluctuate. The live Web site may, therefore, not appear exactly as in
this figure.

SLC2A2 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on SLC2A2 and diabetes, search PubMed.
SLC2A2 is a likely candidate gene for type 2 diabetes because it is a high Km transporter that
regulates entry of glucose in the pancreatic beta cell and triggers insulin secretion.
In one diabetic patient, a mutation that disrupts a highly conserved valine in the 5th membrane spanning domain of GLUT2 was found (3, 9). The valine at position 197 was switched to an
isoleucine residue (V197I). Expression of this mutant protein in Xenopus oocytes abolished the
transporter activity of GLUT2, suggesting that defects in GLUT2 may have a causative role in the
development of diabetes (10).
Many studies have been unsuccessful in finding evidence of association between SNPs in
the SLC2A2 gene and type 2 diabetes (11-13). One SNP that may be significant is a threonine to
isoleucine switch at residue 110 (T110I), which occurs in the second membrane spanning domain
of GLUT2. This SNP is present equally in diabetics and non-diabetics (9), but in one additional
study a modest association between the T110I variant and type 2 diabetes was observed (8). The
functional role of this polymorphism remains unclear as the variant protein appears to have similar levels of expression and does not seem to disrupt glucose transporter activity (10).

References
3-23

Diabetes

Genetic Factors in Type 2 Diabetes

1. Guillam MT, Hummler E, Schaerer E et al. Early diabetes and abnormal postnatal pancreatic islet
development in mice lacking Glut-2. Nat Genet 17:327330; 1997. (PubMed)
2. Efrat S. Making sense of glucose sensing. Nat Genet 17:249250; 1997. (PubMed)
3. Santer R, Groth S, Kinner M et al. The mutation spectrum of the facilitative glucose transporter gene
SLC2A2 (GLUT2) in patients with Fanconi-Bickel syndrome. Hum Genet 110:2129; 2002. (PubMed)
4. Katagiri H, Asano T, Shibasaki Y et al. Substitution of leucine for tryptophan 412 does not abolish
cytochalasin B labeling but markedly decreases the intrinsic activity of GLUT1 glucose transporter. J Biol
Chem 266:77697773; 1991. (PubMed)
5. Garcia JC, Strube M, Leingang K et al. Amino acid substitutions at tryptophan 388 and tryptophan 412 of
the HepG2 (Glut1) glucose transporter inhibit transport activity and targeting to the plasma membrane in
Xenopus oocytes. J Biol Chem 267:77707776; 1992. (PubMed)
6. Ishihara H, Asano T, Katagiri H et al. The glucose transport activity of GLUT1 is markedly decreased by
substitution of a single amino acid with a different charge at residue 415. Biochem Biophys Res
Commun 176:922930; 1991. (PubMed)
7. Maiden MC, Davis EO, Baldwin SA et al. Mammalian and bacterial sugar transport proteins are
homologous. Nature 325:641643; 1987. (PubMed)
8. Barroso I, Luan J, Middelberg RP et al. Candidate gene association study in type 2 diabetes indicates a
tole for genes involved in beta-Cell function as well as insulin action. PLoS Biol 1:E20E20; 2003.
(PubMed)
9. Tanizawa Y, Riggs AC, Chiu KC et al. Variability of the pancreatic islet beta cell/liver (GLUT 2) glucose
transporter gene in NIDDM patients. Diabetologia 37:420427; 1994. (PubMed)
10. Mueckler M, Kruse M, Strube M et al. A mutation in the Glut2 glucose transporter gene of a diabetic
patient abolishes transport activity. J Biol Chem 269:1776517767; 1994. (PubMed)
11. Li SR, Alcolado JC, Stocks J et al. Genetic polymorphisms at the human liver/islet glucose transporter
(GLUT2) gene locus in Caucasian and West Indian subjects with type 2 (non-insulin-dependent) diabetes
mellitus. Biochim Biophys Acta 1097:293298; 1991. (PubMed)
12. Baroni MG, Alcolado JC, Pozzilli P et al. Polymorphisms at the GLUT2 (beta-cell/liver) glucose
transporter gene and non-insulin-dependent diabetes mellitus (NIDDM): analysis in affected pedigree
members. Clin Genet 41:229324; 1992. (PubMed)
13. Moller AM, Jensen NM, Pildal J et al. Studies of genetic variability of the glucose transporter 2 promoter
in patients with type 2 diabetes mellitus. J Clin Endocrinol Metab 86:21812186; 2001. (PubMed)

Link Roundup
Live Searches
Diabetes and GLUT2 in PubMed | PubMed Central | Books

Background Information
GLUT2 in OMIM

Molecular Biology
GLUT2 in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_005612.14&gene=SLC2A2&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4557851] | SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=6514&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [http://www.ncbi.nlm.nih.gov/
sutils/blink.cgi?pid=4557851&cut=100&org=1] | HomoloGene

3-24

Diabetes

Genetic Factors in Type 2 Diabetes

The Transcription Factor HNF4A


Summary
The HNF4A gene encodes a transcription factor that is found in the liver and pancreas. Because
HNF4A maps to a region of chromosome 20 that is linked with type 2 diabetes and because
mutations of this gene cause a rare form of autosomal dominant diabetes, the HNF4A gene is
considered to be a strong candidate for involvement in type 2 diabetes.

Nomenclature
Official gene name: hepatocyte nuclear factor 4 alpha
Official gene symbol: HNF4A
Alias: transcription factor 14, TCF14

Background
The expression of a wide range of genes in the liver is regulated by the transcription factor
HNF4A (hepatocyte nuclear factor 4 alpha). Many functions that the liver carries out may appear
and disappear, depending on whether HNF4A is expressed. In addition, HNF4A controls the
expression of another transcription factor; hepatocyte nuclear factor 1 (HNF1A), which in turn
regulates the expression of several important genes in the liver.
As the name suggests, hepatocyte nuclear factor 4 is found in abundance in the liver, but it is
also found in the beta cells of the pancreas, the kidneys, and intestines. Together with other transcription factors such as HNF1A and HNF1B (encoded by TCF1 and TCF2, respectively), they
make up part of a network of transcription factors that function together to control gene expression in the developing embryo. In particular, HNF4A is thought to play an important role in the
development of the liver, kidney, and intestines.
In pancreatic beta cells, this network of transcription factors regulates the expression of the
insulin gene. In addition, HNF4 and HNF1 regulate the expression of several other genes linked
with insulin secretion, e.g., genes that encode proteins involved in glucose transport and glucose
metabolism (1, 2).
Mutations of HNF4A can cause an extremely rare form of diabetes, maturity onset diabetes in
the young (MODY). Whereas type 2 diabetes is a disorder usually of late onset with significant
genetic basis, MODY by definition occurs in the young (onset at age less than 25 years) and is
inherited in an autosomal dominant fashion.
Patients with MODY caused by HNF4A mutations exhibit a normal response to insulin (they
are not insulin resistant) but do show an impaired response to secreting insulin in the presence of
glucose. Over time, the amount of insulin secreted decreases, leading to diabetes. Affected individuals are usually treated with oral hypoglycemic agents, but up to 40% of patients may require
insulin. This form of diabetes is extremely rare and has only been identified in 13 families (3).
The HNF4A gene is suspected to play a role in type 2 diabetes. Because a mutation of this
gene causes a rare form of diabetes and because the gene maps to an area of chromosome 20
that has been linked with type 2 diabetes, it is speculated that particular HNF4A haplotypes may
be associated with altered insulin secretion.
3-25

Diabetes

Genetic Factors in Type 2 Diabetes

Molecular Information
Hepatocyte nuclear factors (HNFs) are a heterogeneous class of evolutionarily conserved transcription factors that are required for cellular differentiation and metabolism. HNF4A is an orphan
receptor; the ligand(s) that binds to this receptor is unknown (4).
A BLAST search [www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=31077205&cut=100&org=1]
using human HNF4A as a query finds proteins in 47 different species, which are all multicellular
species (metazoans). However, potential true homologous genes have thus far been identified in
the mouse, rat, and nematode worm.
The HNF4A gene maps to chromosome 20 (Figure 1). It has 11 exons (coding regions) that
span over 30,000 bases (see evidence [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_011362.8&gene=HNF4A&graphiconly=TRUE]). There are at least three
different transcript variants of this gene, which encode three different protein isoforms (a, b, and
c). The longest mRNA transcript, NM_000457, encodes the longest HNF4A protein (isoform b),
containing over 450 amino acids.

Figure 1: Location of HNF4A on the human genome.


HNF4A maps to chromosome 20, between approximately 43,70043,750 kilobases (kb). Click

or here for a current and

interactive view of the location of HNF4A in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of HNF4A may fluctuate. Therefore, the live Web site may not appear exactly as in this
figure.

3-26

Diabetes

Genetic Factors in Type 2 Diabetes


Several single nucleotide polymorphisms (SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=3172&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been found
within the HNF4A gene (Figure 2). At the time of writing, three non-synonymous amino acid
changes caused by SNPs have been observed in the longest protein isoform (isoform b). At
present, none of these SNPs have been associated with either type 2 diabetes or MODY (see
known allelic variants). However, several SNPs upstream of the coding region have been associated with both MODY and type 2 diabetes (see below).

Figure 2: SNP positions of HNF4A mapped to the 3D structure of the ligand-binding domain of rat HNF4A.
The figure shows the positions of a non-synonymous amino acid change (Leu148) caused by a SNP in the coding
sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

HNF4A and Diabetes: Digest of Recent Articles


For a more complete list of research articles on HNF4A and diabetes, search PubMed.
It has been proposed that the pancreatic beta cell is sensitive to the amount of HNF4A
present (5). A variety of nonsense and missense mutations in HNF4A cause MODY, where a
decline in insulin secretion occurs. Similarly, SNPs of the gene may have an impact on the beta
cell function, increasing or decreasing insulin secretion.

3-27

Diabetes

Genetic Factors in Type 2 Diabetes


In one British study, a haplotype that was linked with reduced disease risk was identified (5).
Individuals with the "reduced risk" haplotype were significantly associated with increased insulin
secretion. These individuals also showed a trend toward lower fasting glucose levels and 2-hour
plasma glucose levels. This finding led to the speculation that a particular HNF4A haplotype
might be associated with increased insulin secretion capacity and protection against diabetes.
Genetic variants associated with increased diabetes risk have been identified upstream of the
HNF4A coding region, in a recently discovered alternative promoter called P2. P2 lies 46 kb
upstream of the P1 promoter, and while both promoters are involved in the transcription of
HNF4A, they seem to have different roles in different cells. Transcripts from both P1 and P2 have
been found in the pancreas, but the P2 promoter is thought to be the primary transcription site in
the beta cells and a mutation of P2 is a cause of MODY (6).
In the search for genetic variation near the HNF4A gene, two studies, one of the Ashkenazi
Jew population (7), and the other of the Finnish population (8), identified four SNPs that were
associated with type 2 diabetes. These SNPs (named rs4810424, rs1884613, rs1884614, and
rs2144908) flanked the P2 promoter and were associated with type 2 diabetes risk in both study
populations, and the risk of diabetes attributed to each SNP was very similar between the two
populations.
The mechanism by which SNPs near and in the HNF4A gene increases the risk of diabetes is
not yet known. Perhaps because many HNF transcription factors can bind directly to the P2 promoter, any alteration of the binding sites for these factors could disrupt the regulation of HNF4A
expression. Whereas both P1 and P2 promoters are used in liver cells, mainly P2 is used in pancreas cells, leading to SNPs affecting the P2 promoter disproportionately affecting HNF4A
expression in the pancreas, in turn leading to beta cell malfunction and diabetes (9).

References
1. Stoffel M, Duncan SA. The maturity-onset diabetes of the young (MODY1) transcription factor HNF4alpha
regulates expression of genes required for glucose transport and metabolism. Proc Natl Acad Sci USA
94:1320913214; 1997. (PubMed)
2. Wang H, Maechler P, Antinozzi PA et al. Hepatocyte nuclear factor 4alpha regulates the expression of
pancreatic beta-cell genes implicated in glucose metabolism and nutrient-induced insulin secretion. J Biol
Chem 275:3595335959; 2000. (PubMed)
3. Fajans SS, Bell GI, Polonsky KS et al. Molecular mechanisms and clinical pathophysiology of maturityonset diabetes of the young. N Engl J Med 345:971980; 2001. (PubMed)
4. Duncan SA, Navas MA, Dufort D et al. Regulation of a transcription factor network required for
differentiation and metabolism. Science 281:692695; 1998. (PubMed)
5. Barroso J, Luan J, Middelberg RP et al. Candidate gene association study in type 2 diabetes indicates a
role for genes involved in beta-Cell function as well as insulin action. PLoS Biol 1:E20E20; 2003.
(PubMed)
6. Thomas H, Jaschkowitz K, Bulman M et al. A distant upstream promoter of the HNF-4alpha gene
connects the transcription factors involved in maturity-onset diabetes of the young. Hum Mol Genet
10:20892097; 2001. (PubMed)
7. Love-Gregory LD, Wasson J, Ma J et al. A common polymorphism in the upstream promoter region of the
hepatocyte nuclear factor-4 alpha gene on chromosome 20q is associated with type 2 diabetes and
appears to contribute to the evidence for linkage in an Ashkenazi jewish population. Diabetes 53:1134
1140; 2004. (PubMed)
3-28

Diabetes

Genetic Factors in Type 2 Diabetes

8. Silander K, Mohlke KL, Scott LJ et al. Genetic variation near the hepatocyte nuclear factor-4 alpha gene
predicts susceptibility to type 2 diabetes. Diabetes 53:11411149; 2004. (PubMed)
9. Odom DT, Zizlsperger N, Gordon DB et al. Control of pancreas and liver gene expression by HNF
transcription factors. Science 303:13781381; 2004. (PubMed)

Link Roundup
Live Searches
Diabetes and HNF4A in PubMed | PubMed Central | Books

Background Information
HNFA4 in OMIM

Molecular Biology
HNF4A in Entrez Gene | Evidence Viewer [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_011362.8&gene=HNF4A&graphiconly=TRUE] | Map Viewer | Domains [www.ncbi.nlm.nih.gov/Structure/
cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=31077205]: Zinc Finger, Ligand-binding domain | SNPs [www.ncbi.nlm.nih.gov/
SNP/snp_ref.cgi?locusId=3172&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [www.ncbi.
nlm.nih.gov/sutils/blink.cgi?pid=31077205&cut=100&org=1] | HomoloGene

3-29

Diabetes

Genetic Factors in Type 2 Diabetes

The Insulin Hormone (INS)


Summary
The INS gene gene encodes the precursor to the hormone insulin. Genetic variations of the
insulin gene (variable number tandem repeats and SNPs) may play a role in susceptibility to type
1 and type 2 diabetes.

Background
Insulin is a hormone that has a wide range of effects on metabolism. Its overall action is to
encourage the body to store energy rather than use it, e.g., insulin favors the storage of glucose
as glycogen or fat as opposed to breaking down glucose to release ATP. For a summary of the
actions of insulin, see the Physiology and Biochemistry of Sugar Regulation.
Insulin is composed of two distinct polypeptide chains, chain A and chain B, which are linked
by disulfide bonds. Many proteins that contain subunits, such as hemoglobin, are the products of
several genes. However, insulin is the product of one gene, INS.
INS actually encodes an inactive precursor called preproinsulin. Preproinsulin is processed
into proinsulin by removal of a signaling peptide; however, proinsulin is also inactive. The final
processing step involves removal of a C-peptide (a connecting peptide that links chain A to chain
B), and this process produces the mature and active form of insulin. For further information, see
The Story of Insulin.

Molecular Information
Several species, including the rat, mouse, and some species of fish, have two insulin genes. In
contrast, in humans there is a single insulin gene that is located on chromosome 11 (Figure 1). It
has three exons (coding regions) that span about 2,200 bases (see evidence [http://www.ncbi.
nlm.nih.gov/sutils/evv.cgi?taxid=9606&contig=NT_009237.16&gene=INS&graphiconly=TRUE]).
Exon 2 encodes the B chain, along with the signal peptide and part of the C-peptide found in the
insulin precursors. Exon 3 encodes the A chain and the remainder of the C-peptide.
C-peptide is secreted in equal amounts to insulin, but it has long been thought that it has no
biological role. However, in diabetic rats C-peptide has been shown to reduce the dysfunction of
blood vessels and the nervous system that is common in diabetes (1). C-peptide contains the
greatest variation among species, whereas regions of insulin that bind to the insulin receptor are
highly conserved.
Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=3630&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the INS gene, none (at the time of writing) of which cause non-synonymous amino
acid changes in the mature protein (see the allelic variants that are known to be associated with
disease).
A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4557671&cut=100&org=1]
using the human proinsulin precursor as a query finds proteins in 107 different species, which are
all metazoans apart from three plants and one bacterium. However, potential true homologous
genes have thus far been identified only in the mouse and rat.
3-30

Diabetes

Genetic Factors in Type 2 Diabetes

Figure 1: Location of INS on the human genome.


INS maps to chromosome 11, approximately between 2,1442,148 kilobases (kb). Click

on the figure or here for a

current and interactive view of the location of INS in the human genome.
Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of INS may fluctuate; therefore, the live Web site may not appear exactly as in this
figure.

INS and Diabetes: Digest of Recent Articles


For a more complete list of research articles on INS and diabetes, search PubMed.
There is conflicting evidence for the role of the insulin gene in type 2 diabetes predisposition.
The gene contains a variable number of tandem repeats, and it has been proposed that the number of repeats may have an impact on birth weight and diabetes susceptibility (2).
In one study, a DNA variant in the 3' untranslated region of the INS gene was associated with
type 2 diabetes risk (3).
Certain mutations of the insulin gene can result in a rare form of diabetes. One form of mutant
insulin, called Chicago insulin, has been found in individuals who have diabetes that resembles
MODY. This form of diabetes is rare; it is caused by a single gene (is monogenic) and is inherited
in an autosomal dominant fashion.
Mutations of insulin affecting the phenylalanine amino acid at position 24 or 25 of the B chain
(B24 or B25) are associated with diabetes. In Chicago insulin, B25 is mutated to leucine (4-6), or
B24 is mutated to serine (7). The phenylalanine at position 24 is highly conserved, and in crystal
structures of insulin, the aromatic ring of this amino acid is important in anchoring the region of
the insulin that binds to the insulin receptor (8).
Point mutations of INS can also result in increased levels of proinsulin (proinsulinemia) (9). A
switch from arginine to histidine at amino acid position 65 results in proinsulinemia, when it is
probable that the mutation of arginine 65 disrupts the site at which the C-peptide is cleaved by
beta cell proteases to produce insulin from proinsulin. This mutation does not appear to impair
glucose tolerance in heterozygous carriers (10, 11).

3-31

Diabetes

Genetic Factors in Type 2 Diabetes


Other point mutations of arginine 65 have been found in individuals with type 2 diabetes and
a raised level of insulin and/or proinsulin (12, 13). However, one of the individuals had a
Arg65Pro mutation that was not thought to be linked to their diabetes, because the mutation was
present in a non-diabetic daughter and absent in a daughter who had had gestational diabetes
(13).

References
1. Ido Y, Vindigni A, Chang K et al. Prevention of Vascular and Neural Dysfunction in Diabetic Rats by CPeptide. Science 277:563566; 1997. (PubMed)
2. Huxtable SJ, Saker J, Haddad, L et al. Analysis of parent-offspring trios provides evidence for linkage and
association between the insulin gene and type 2 diabetes mediated exclusively through paternally
transmitted class III variable number tandem repeat alleles. Diabetes 49:126130; 2000. (PubMed)
3. Barroso I, Luan J, Middelberg RP et al. Candidate gene association study in type 2 diabetes indicates a
role for genes involved in beta-cell function as well as insulin action. PLoS Biol 1:E20E20; 2003.
(PubMed)
4. Tager H, Given B, Baldwin D et al. A structurally abnormal insulin causing human diabetes. Nature
281:122125; 1979. (PubMed)
5. Given BD, Mako ME, Mako ME et al. Diabetes due to secretion of an abnormal insulin. N Engl J Med
302:129135; 1980. (PubMed)
6. Shoelson S, Haneda M, Blix P et al. Three mutant insulins in man. Nature 302:540543; 1983. (PubMed)
7. Haneda M, Chan SJ, Kwok SC et al. Studies on mutant human insulin genes: identification and sequence
analysis of a gene encoding [SerB24]insulin. Proc Natl Acad Sci U S A 80(20):63666370; 1983.
(PubMed)
8. Hua QX, Shoelson SE, Inouye K et al. Paradoxical structure and function in a mutant human insulin
associated with diabetes mellitus. Proc Natl Acad Sci U S A 90:582586; 1993. (PubMed)
9. Shibasaki Y, Kawakami T, Kanazawa Y et al. Posttranslational cleavage of proinsulin is blocked by a
point mutation in familial hyperproinsulinemia. J Clin Invest 76:378380; 1985. (PubMed)
10. Barbetti F, Raben N, Kadowaki T et al. Two unrelated patients with familial hyperproinsulinemia due to a
mutation substituting histidine for arginine at position 65 in the proinsulin molecule: identification of the
mutation by direct sequencing of genomic deoxyribonucleic acid amplified by polymerase chain reaction.
J Clin Endocrinol Metab 71:164169; 1990. (PubMed)
11. Roder ME, Vissing H, Nauck MA et al. Hyperproinsulinemia in a three-generation Caucasian family due
to mutant proinsulin (Arg65-His) not associated with impaired glucose tolerance: the contribution of
mutant proinsulin to insulin bioactivity. J Clin Endocrinol Metab 81:16341640; 1996. (PubMed)
12. Yano H, Kitano N, Morimoto M et al. A novel point mutation in the human insulin gene giving rise to
hyperproinsulinemia (proinsulin Kyoto). J Clin Invest 89:19021907; 1992. (PubMed)
13. Warren-Perry MG, Manley SE, Ostrega D et al. A novel point mutation in the insulin gene giving rise to
hyperproinsulinemia. J Clin Endocrinol Metab 82:16291631; 1997. (PubMed)

Link Roundup
Live Searches
Diabetes and insulin in PubMed | PubMed Central | Books

Background Information
Insulin in OMIM

Molecular Biology
3-32

Diabetes

Genetic Factors in Type 2 Diabetes

INS in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?


taxid=9606&contig=NT_009237.16&gene=INS&graphiconly=TRUE] | Map Viewer | Insulin Domain [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4557671] | SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=3630&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseR] | Allelic Variants | BLink [http://www.ncbi.nlm.nih.gov/
sutils/blink.cgi?pid=4557671&cut=100&org=1] | HomoloGene

3-33

Diabetes

Genetic Factors in Type 2 Diabetes

The Insulin Receptor (INSR)


Summary
The INSR gene encodes the receptor for insulin. Mutations of the insulin receptor can cause rare
forms of diabetes and may play a role in susceptibility to type 2 diabetes.

Background
Some receptors found on the surfaces of cells are directly linked to enzymes located inside the
cell. The insulin receptor belongs to the largest family of such enzyme-linked receptors, known as
the receptor protein-tyrosine kinase (RTK) family.
More than 50 RTKs have been identified, and they share a common structure: an N-terminal
extracellular domain containing a ligand-binding site, a single transmembrane helix, and a Cterminal cytosolic domain with protein-tyrosine kinase activity.
Receptor tyrosine kinases in Molecular Cell Biology [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..
ShowSection&rid=mcb.section.5803]

Most of the RTKs consist of single polypeptides, and in response to ligand binding, they
dimerize. An exception is the insulin receptor that exists as a dimer in the absence of its ligand.
After the ligand has bound to and activated its receptor, the tyrosine kinase activity of the cystolic
domain of the RTK is activated. Each kinase phosphorylates specific tyrosine residues in the
cytosolic domain of its dimer partner, a process termed autophosphorylation.
Autophosphorylation occurs in two stages. First, tyrosine residues near the catalytic site are
phosphorylated. In the insulin receptor, this results in a conformational change that permits ATP
binding, and it may also increase the receptor protein kinase activity. The receptor kinase then
phosphorylates other tyrosine residues outside the catalytic domain, creating phosphotyrosine
residues that serve as docking ports for additional proteins that transmit intracellular signals
downstream of the activated receptors.
Read more on the signal pathways initiated by insulin binding in Molecular Cell Biology [http://www.ncbi.nlm.nih.gov/
books/bv.fcgi?call=bv.View..ShowSection&rid=mcb.section.5890#5898]

One important docking protein in the insulin receptor is IRS1 (insulin receptor substrate 1).
IRS1 binds to the phosphotyrosine residues in activated receptors via its Src homology 2 (SH2)
domains. Binding of insulin to its receptor can initiate two distinct signaling pathways: one that
includes Ras, and one that does not. The initiation of both pathways appears to involve IRS1.

Molecular Information
The insulin receptor is a dimer, consisting of two polypeptide chains that are linked by disulfide
bonds. Each polypeptide contains an alpha subunit (extracellular) and a beta subunit (transmembrane and intracellular). The alpha and beta subunits are encoded by a single gene, INSR, which
maps to chromosome 19 (Figure 1). It has 23 exons (coding regions) that span about 180,000
3-34

Diabetes

Genetic Factors in Type 2 Diabetes


bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_011255.14&gene=INSR&graphiconly=TRUE]). The INSR gene encodes
a precursor that starts with an 27-amino acid signal sequence, followed by the receptor alpha
subunit, the cleavage site of the processing enzyme, and the receptor beta subunit (1).

Figure 1: Location of INSR on the human genome.


INSR maps to chromosome 19, between approximately 6,7007,500 kilobases (kb). Click

or here for a current and

interactive view of the location of INSR in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of INSR may fluctuate. The live Web site may, therefore, not appear exactly as in this
figure.

Depending upon where the insulin receptor is expressed, there are differences in its composition. The alpha subunit found in INSR in the liver differs from its form in muscle and adipose
tissue in terms of molecular weight, carbohydrate composition, and antigenicity (2). The level of
activity of its tyrosine kinase may also differ, depending on the tissue in which the receptor is
expressed. The mechanism underlying receptor heterogeneity is unknown.
Alternative splicing of exon 11 of the INSR gene generates two insulin receptor isoforms,
INSR type a (without exon 11) and INSR type b (with exon 11). These two transcripts are
expressed in a highly regulated fashion and predominate in different tissues (3). There is evidence that selectivity in the action of insulin may be brought about by the two INSR isoforms (4).

3-35

Diabetes

Genetic Factors in Type 2 Diabetes


Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=3643&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the INSR gene, one (at the time of writing) of which causes non-synonymous amino
acid changes in the mature protein (Figure 2). This protein variant has not been associated with
observed cases of disease (see known allelic variants).

Figure 2: SNP positions of INSR mapped to the 3D structure of a mutant human insulin receptor.
The figure shows the position of a non-synonymous amino acid changes (green residue) caused by a SNP in the coding
sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4557884&cut=100&org=1]


using human INSR as a query finds proteins in 42 different species, which are all metazoans
apart from two viruses. However, potential true homologous genes have thus far been identified
only in the mouse, rat, and mosquito.

INSR and Diabetes: Digest of Recent Articles


For a more complete list of research articles on INSR and diabetes, search PubMed.
Mutations in INSR lead to a variety of rare clinical syndromes ranging in severity from Leprechaunism (usually fatal within the first 2 years of life), Rabson-Mendenhall syndrome (survival
into the second decade), and type A insulin resistance (survival into middle age and beyond).

3-36

Diabetes

Genetic Factors in Type 2 Diabetes


Resistance to insulin is one of the characteristics of type 2 diabetes. However, most diabetics
have a normal sequence of the insulin receptor, indicating that if insulin receptor mutations contribute to the development of type 2 diabetes, they will be present only in a minor fraction of the
diabetic population.
Several variants of the insulin receptor have been associated with hyperglycemia and type 2
diabetes (5, 6). A heterozygous mutation changing Val-985 into methionine has been found to be
more common in type 2 diabetics than in controls; the relative risk for diabetes was over 4% for
Met-985 carriers. The prevalence of the mutation increased with increasing serum glucose levels,
suggesting a role for this receptor variant in hyperglycemia.

References
1. Ullrich A, Bell JR, Chen EY et al. Human insulin receptor and its relationship to the tyrosine kinase family
of oncogenes. Nature 313:756761; 1985. (PubMed)
2. Caro JF, Raju SM, Sinha MK et al. Heterogeneity of human liver, muscle, and adipose tissue insulin
receptor. Biochem Biophys Res Commun 151:123129; 1988. (PubMed)
3. Benecke J, Flier JS, Moller DE et al. Alternatively spliced variants of the insulin receptor protein.
Expression in normal and diabetic human tissues. J Clin Invest 89:20662070; 1992. (PubMed)
4. Leibiger B, Leibiger IB, Moede T et al. Selective insulin signaling through A and B insulin receptors
regulates transcription of insulin and glucokinase genes in pancreatic beta cells. Mol Cell 7:559570;
2001. (PubMed)
5. Hart LM, Stolk RP, Heine RJ et al. Association of the insulin-receptor variant Met-985 with hyperglycemia
and non-insulin-dependent diabetes mellitus in the Netherlands: a population-based study. Am J Hum
Genet 59:11191125; 1996. (PubMed)
6. Hart LM, Stolk RP, Dekker JM et al. Prevalence of variants in candidate genes for type 2 diabetes
mellitus in The Netherlands: the Rotterdam study and the Hoorn study. J Clin Endocrinol Metab 84:1002
1006; 1999. (PubMed)

Link Roundup
Live Searches
Diabetes and INSR in PubMed | PubMed Central | Books

Background Information
The Insulin Receptor in OMIM

Molecular Biology
Insulin receptor in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_011255.14&gene=INSR&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4557884]: Tyrosine Kinase, Furin-like Domain | SNPs [http://www.
ncbi.nlm.nih.gov/SNP/snp_ref.cgi?locusId=3643&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants |
BLink [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4557884&cut=100&org=1] | HomoloGene

3-37

Diabetes

Genetic Factors in Type 2 Diabetes

The Potassium Channel KCNJ11


Summary
Closure of potassium channels in the beta cells of the pancreas triggers the release of insulin.
Drugs that close the channel are used in the treatment of diabetes. Variation of the KCNJ11 gene
which encodes this channel has been linked to both increased and decreased insulin release.

Nomenclature
Official gene name: potassium inwardly-rectifying channel, subfamily J, member 11
Official gene symbol: KCNJ11
Alias: inwardly rectifying potassium channel, KIR6.2; beta-cell inward rectifier subunit, BIR; ATPsensitive inward rectifier potassium channel 11, IKATP

Background
Similar to all cells, pancreatic beta cells maintain a concentration gradient of electolytes (e.g.,
potassium, sodium, and calcium) across their membranes. Ion pumps create the gradient by
pushing sodium (Na+) out of cells and potassium (K+) into cells. This produces a membrane polarity; the inner side of the membrane has a negative charge of around -70 mV. To maintain the
membrane potential and to be able to alter it, specific proteins called ion channels span the
membrane and regulate the passage of ions.
The release of insulin is controlled by ion channels that conduct K+ and are sensitive to ATP
(KATP channels). A rise in glucose and the corresponding rise in ATP shuts the KATP channels,
which increases the membrane potential to -40 mV. This state, known as depolarization, opens
calcium channels, and the entry of Ca2+ into the cell triggers the secretion of insulin (Figure 1).

3-38

Diabetes

Genetic Factors in Type 2 Diabetes

Figure 1: Insulin secretion.


1. Glucose is transported into the beta cell by type 2 glucose transporters (GLUT2). Once inside, the first step in glucose
metabolism is the phosphorylation of glucose to produce glucose-6-phosphate. This step is catalyzed by hexokinase; it is
the rate-limiting step in glycolysis, and it effectively traps glucose inside the cell.
2. As glucose metabolism proceeds, ATP is produced in the mitochondria.
3. The increase in the ATP:ADP ratio closes ATP-gated potassium channels in the beta cell membrane. Positively
charged potassium ions (K+) are now prevented from leaving the beta cell.
4. The rise in positive charge inside the beta cell causes depolarization.
5. Voltage-gated calcium channels open allowing calcium ions (Ca2+) to flood into the cell.
6. The increase in intracellular calcium concentration triggers the secretion of insulin via exocytosis.

Drugs (e.g., sulfonylureas) that close beta cell KATP channels are used to treat type 2 diabetes. They are referred to as "oral hypoglycemic agents" because they are taken by mouth and
stimulate the release of insulin, which lowers blood sugar levels.
The KATP channel consists of two types of subunit: a K+-channel subunit (termed Kir6.2),
and a sulfonylurea receptor subunit (SUR), which is a member of the family of ATP-binding cassette (ABC) transporter proteins. The potassium channel is made from four Kir6.2 subunits and
four SUR subunits (1).
The Kir6.2 subunit is encoded by the KCNJ11 gene. The four Kir6.2 subunits form the pore of
the channel through which K+ passes and also contains the ATP-binding sites. The ABCC8 gene
encodes the SUR subunit. The four SUR subunits modulate the activity of the channel and contain the binding site of the sulfonylurea drugs.
3-39

Diabetes

Genetic Factors in Type 2 Diabetes


Mutations in either KCNJ11 or ABBC8 can reduce KATP channel activity, leading to
increased insulin release and low blood sugar levels, a rare disorder called persistent hyperinsulinemic hypoglycemia of infancy (PHHI) (2-4). In contrast, mice that have targeted overactivity of
the KATP channel have severe diabetes (5), and in humans, activating mutations of KCNJ11
cause permanent neonatal diabetes (6).

Molecular Information
The KCNJ11 gene maps to chromosome 11 (Figure 2). It has only one exon (coding region) that
spans about 2000 bases (2 kilobases, or 2 kb) and does not contain introns (non-coding regions)
(see evidence) [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_009237.16&gene=KCNJ11&graphiconly=TRUE]. The gene encodes a
protein of 390 amino acids and two transmembrane segments.

Figure 2: Location of KCNJ11 on the human genome.


KCNJ11 maps to chromosome 11, between approximately 17,37217,375 kilobases (kb). Click

or here for a current

and interactive view of the location of KCNJ11 in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of KCNJ11 may fluctuate. The live Web site may, therefore, not appear exactly as in
this figure.

Each Kir6.2 subunit is associated with a much larger SUR subunit and also contains a binding site for ATP. The binding of one molecule of ATP causes a conformational change in the
channel protein that closes the channel (7-9).

3-40

Diabetes

Genetic Factors in Type 2 Diabetes


A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=25777632&cut=100&org=1]
using human KCNJ11 as a query finds proteins in 30 different species, which are all metazoans.
However, potential true homologous genes have thus far been identified only in the mouse, rat,
and nematode.
Several single nucleotide polymorphisms (SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=3767&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been found
within the KCNJ11 gene, four (at the time of writing) of which cause non-synonymous amino acid
changes in the mature protein (Figure 3). At least two of these, rs5219 which encodes the E23K
polymorphism and rs5215 which encodes the I337V polymorphism, have been associated with
impaired insulin response and increased risk of type 2 diabetes (see below).

Figure 3: SNP positions of KCNJ11 mapped to the 3D structure of inward rectifier potassium channel 1.
The figure shows the positions of some of the non-synonymous amino acid changes (green residues) caused by SNPs in
the coding sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

KCNJ11 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on KCNJ11 and diabetes, search PubMed.
Three common SNPs (E23K, L270V, and I337V) have been found in the KCNJ11 gene in
Caucasians (10-13).

3-41

Diabetes

Genetic Factors in Type 2 Diabetes


The E23K variant is caused by a switch of guanine to adenine, resulting in a glutamic acid (E)
to lysine (K) substitution at codon 23. Analysis of the E23K variant in various Caucasian populations shows that KK homozygosity has a stronger association with diabetes compared with heterozygous EK or the wild-type EE (11). However, this risk has not been observed in other
Caucasian populations (12).
To investigate the effect of E23K, L270V, and I337V on the functioning of the KATP channel,
human potassium channels containing these SNPs were made. Only one variant, E23K, altered
channel function; it made the channel less sensitive to ATP. Therefore, the E23K channels were
more excitable; they needed greater amounts of ATP before they closed and so were more likely
to remain open. Because the release of insulin is inhibited when the KATP channel is open, E23K
is thought to contribute to the development of diabetes by impairing insulin release (14-16).
The two genes that encode the KATP channel, ABBC8 and KCNJ11, reside adjacent to one
another on chromosome 11. A variant of ABCC8, called A1369S, is in almost complete linkage
disequilibrium with the E23K variant of KCNJ11. This means that from the genetic evidence, it is
difficult to determine whether it is the A1369S variant or the E23K variant that predisposes to type
2 diabetes (17).
In individuals who are obese or have type 2 diabetes, a chronic elevation of free fatty acids is
seen. This leads to the accumulation of long-chain acyl-CoAs in pancreatic beta cells that stimulate KATP channels and inhibit insulin release. The diabetogenic effect of possessing both the
E23K and I337V variants may involve an enhanced stimulatory effect of long-chain acyl-CoAs in
polymorphic KATP channels (18).
The theory that variations in the KCNJ11 gene may inhibit insulin secretion is supported by a
study of adults who did not have diabetes but did have the E23K SNP. It was found that E23K
was associated with a impaired insulin release in response to glucose and also an increased
body mass index (BMI) (19).

References
1. Inagaki N, Gonoi T, Clement JT et al. Reconstitution of IKATP: an inward rectifier subunit plus the
sulfonylurea receptor. Science 270:11661170; 2003. (PubMed)
2. Thomas PM, Cote GC, Wohllk N et al. Mutations in the sulfonylurea receptor gene in familial persistent
hyperinsulinemic hypoglycemia of infancy. Science 268:426429; 1995. (PubMed)
3. Nestorowicz A, Inagaki N, Gonoi T et al. A nonsense mutation in the inward rectifier potassium channel
gene, Kir6.2, is associated with familial hyperinsulinism. Diabetes 46:17431748; 1997. (PubMed)
4. Nestorowicz A, Glaser B, Wilson BA et al. Genetic heterogeneity in familial hyperinsulinism. Hum Mol
Genet 7:11191128; 1998. (PubMed)
5. Koster JC, Marshall BA, Enson N et al. Targeted overactivity of beta cell K(ATP) channels induces
profound neonatal diabetes. Cell 100:645654; 2000. (PubMed)
6. Gloyn AL, Pearson ER, Antcliff JF et al. Activating mutations in the gene encoding the ATP-sensitive
potassium-channel subunit Kir6.2 and permanent neonatal diabetes. N Engl J Med 350:18381849;
2004. (PubMed)
7. Markworth E, Schwanstecher C, Schwanstecher M. ATP4- mediates closure of pancreatic beta-cell ATPsensitive potassium channels by interaction with 1 of 4 identical sites. Diabetes 49:14131418; 2003.
(PubMed)
3-42

Diabetes

Genetic Factors in Type 2 Diabetes

8. Vanoye CG, MacGregor GG, Dong K et al. The carboxyl termini of K(ATP) channels bind nucleotides. J
Biol Chem 277:2326023270; 2002. (PubMed)
9. Trapp S, Haider S, Jones P et al. Identification of residues contributing to the ATP binding site of Kir6.2
Embo J 22:29032912; 2003. (PubMed)
10. Hansen L, Echwald SM, Hansen T et al. Amino acid polymorphisms in the ATP-regulatable inward
rectifier Kir6.2 and their relationships to glucose- and tolbutamide-induced insulin secretion, the insulin
sensitivity index, and NIDDM. Diabetes 46:508512; 1997. (PubMed)
11. Hani EH, Boutin P, Durand E et al. Missense mutations in the pancreatic islet beta cell inwardly rectifying
K+ channel gene (KIR6.2/BIR): a meta-analysis suggests a role in the polygenic basis of Type II diabetes
mellitus in Caucasians. Diabetologia 41:15111515; 1998. (PubMed)
12. Gloyn AL, Weedon MN, Owen KR et al. Large-scale association studies of variants in genes encoding
the pancreatic beta-cell KATP channel subunits Kir6.2 (KCNJ11) and SUR1 (ABCC8) confirm that the
KCNJ11 E23K variant is associated with type 2 diabetes. Diabetes 52:568572; 2003. (PubMed)
13. Hart LM, van Haeften TW, Dekker JM et al. Variations in insulin secretion in carriers of the E23K variant
in the KIR6.2 subunit of the ATP-sensitive K(+) channel in the beta-cell. Diabetes 51:31353138; 2002.
(PubMed)
14. Schwanstecher C, Schwanstecher M. Nucleotide sensitivity of pancreatic ATP-sensitive potassium
channels and type 2 diabetes. Diabetes 51:358362; 2002. (PubMed)
15. Schwanstecher C, Neugebauer B, Schulz M et al. The common single nucleotide polymorphism E23K in
K(IR)6.2 sensitizes pancreatic beta-cell ATP-sensitive potassium channels toward activation through
nucleoside diphosphates. Diabetes 51:363367; 2002. (PubMed)
16. Schwanstecher C, Meyer U, Schwanstecher M. K(IR)6.2 polymorphism predisposes to type 2 diabetes by
inducing overactivity of pancreatic beta-cell ATP-sensitive K(+) channels. Diabetes 51:875879; 2002.
(PubMed)
17. Florez JC, Burtt N, de Bakker PI et al. Haplotype structure and genotype-phenotype correlations of the
sulfonylurea receptor and the islet ATP-sensitive potassium channel gene region. Diabetes 53:1360
1368; 2004. (PubMed)
18. Riedel MJ, Boora P, Steckley D et al. Kir6.2 polymorphisms sensitize beta-cell ATP-sensitive potassium
channels to activation by acyl CoAs: a possible cellular mechanism for increased susceptibility to type 2
diabetes? Diabetes 52:26302635; 2003. (PubMed)
19. Nielsen EM, Hansen L, Carstensen B et al. The E23K variant of Kir6.2 associates with impaired postOGTT serum insulin response and increased risk of type 2 diabetes. Diabetes 52:573577; 2003.
(PubMed)

Link Roundup
Live Searches
Diabetes and KCNJ11 in PubMed | PubMed Central | Books

Background Information
OMIM [http://www.ncbi.nlm.nih.gov/entrez/dispomim.cgi?id=600937]

Molecular Biology
KCNJ11 in Entrez Gene | Evidence Viewer [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_009237.16&gene=KCNJ11&graphiconly=TRUE] | Map Viewer | Domain [www.ncbi.nlm.nih.gov/Structure/
cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=25777632] | SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=3767&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [www.ncbi.nlm.nih.gov/sutils/
blink.cgi?pid=25777632&cut=100&org=1] | HomoloGene

3-43

Diabetes

Genetic Factors in Type 2 Diabetes

The Enzyme Lipoprotein Lipase (LPL)


Summary
Lipoprotein lipase (LPL) is an enzyme that breaks down triglycerides. LPL is functionally impaired
or present at low levels in many type 2 diabetics. Some evidence suggests that insulin plays a
role in regulating LPL synthesis.

Background
Cholesterol and triglycerides circulate in the blood as a complex with proteins, called lipoproteins.
There are several different types of lipoprotein: high-density lipoprotein (HDL), intermediatedensity lipoprotein (IDL), low-density lipoprotein (LDL), and very low-density lipoprotein (VLDL).
VLDL and LDL are sometimes referred to as bad cholesterol because they deposit cholesterol in arteries, which over time contributes to the lipid core of atherosclerotic plaques. LDL
makes up 6070% of the total cholesterol in blood, whereas VLDL makes up 1015%.
IDL is an intermediate between VLDL and LDL. It is richer in cholesterol than VLDL and in
clinical tests gets collected with the LDL fraction.
HDL, sometimes referred to as good cholesterol, makes up about 30% of blood cholesterol.
The level of HDL is inversely correlated with the risk of coronary heart disease, and it seems to
play a protective role. HDL might carry cholesterol away from the arteries to the liver, where it is
broken down.
For more information about blood cholesterol levels, visit the National Heart, Blood, and Lung Institute [www.nhlbi.nih.
gov/health/public/heart/chol/wyntk.htm].

The breakdown and reformation of lipoproteins are interlinked via several steps that often
require the action of enzymes, one of which is lipoprotein lipase (LPL). It is found in heart, muscle, and adipose tissue, where it breaks down triglycerides in VLDL and chylomicrons by hydrolysis into IDL (intermediate-density lipoprotein), fatty acids, and glycerol. Although the fatty acids
can be used as fuel, the IDL, which is richer in cholesterol than VLDL, is taken up by the liver or is
formed into LDL (low-density lipoprotein). LPL activity also indirectly raises HDL levels because
LPL-mediated hydrolysis of VLDL provides surface components that merge with HDL3 to form
HDL2 particles.
For a detailed description of lipoprotein interconversions, visit The University of Manitoba [www.umanitoba.ca/
faculties/medicine/units/biochem/coursenotes/blanchaer_tutorials/LipTutWeb/pages/page1.htm].

As well as the well-established role of LPL in lipid metabolism, it can also bind specific cellsurface proteins, often via heparan sulfate proteoglycans attached to epithelial cells. This action
is independent of any catalytic activity, which means that LPL can "bridge" lipoproteins and the
surface of cells, leading to an increase in the cellular uptake of lipoproteins (1). In the blood vessel wall, this action is thought to be proatherogenic.

3-44

Diabetes

Genetic Factors in Type 2 Diabetes


Severe mutations in LPL cause a deficiency that results in type I hyperlipoproteinemia
( OMIM; emedicine [www.emedicine.com/oph/topic505.htm]), a disorder characterized by high
levels of lipoprotein and triglycerides in the blood.

Molecular Information
Lipoprotein lipase was so difficult to purify that the protein sequence had to be determined from
the nucleotide sequence of its cDNA. The LPL gene maps to chromosome 8 (Figure 1). It has 10
exons (coding regions) that span about 28,800 bases (see evidence [www.ncbi.nlm.nih.gov/sutils/
evv.cgi?taxid=9606&contig=NT_030737.8&gene=LPL&graphiconly=TRUE]) (2-5). The gene
encodes a preprotein of 475 amino acids that contains a signal peptide. Cleavage of the signal
results in a mature protein of 448 amino acids.

Figure 1: Location of LPL on the human genome.


LPL maps to chromosome 8, between approximately 19,60019,830 kilobases (kb). Click

or here for a current and

interactive view of the location of LPL in the human genome.


Note: this figure was created from Build 33 of the human genome. Because the data are recomputed between
genome builds, the exact location of LPL may fluctuate. The live Web site may, therefore, not appear exactly as in this
figure.

Lipoprotein lipase, along with hepatic lipase and pancreatic lipase, makes up a lipase superfamily (6); that is, based on a comparison of the amino acid sequence, all three proteins appear
to have evolved from a common ancestor.
All lipases, including evolutionarily unrelated prokaryotic lipases, share a serine protease-like
catalytic triad of amino acids (Ser-His-Asp). Another class of enzyme, serine proteases,
hydrolyzes peptide bonds in many biological settings, for example, within the blood-clotting cascade. The similar catalytic triad seen in the lipase superfamily suggests that a similar mechanism
of hydrolysis [www.ncbi.nlm.nih.gov/books/bv.fcgi?tool=bookshelf&call=bv.View..

3-45

Diabetes

Genetic Factors in Type 2 Diabetes


ShowSection&searchterm=serine&rid=stryer.section.1170#1178] is used, this time to hydrolyze
fats rather than peptides. Site-directed mutagenesis experiments on LPL support this hypothesis
(7).
Although the structure of LPL has not yet been solved, mapping the LPL sequence to the
crystal structure of pancreatic lipase [www.ncbi.nlm.nih.gov/Structure/mmdb/mmdbsrv.cgi?
form=6&db=t&Dopt=s&uid=21535] (8) gives a good estimate of the structure of LPL [www.ncbi.
nlm.nih.gov/Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4557727]. Although
the enzymatic activity involving the catalytic triad is in the large amino-terminal lipase domain
(residues 1312), the smaller carboxy-terminal domain (the PLAT/LH2 domain; residues 313
448) binds the lipoprotein substrate. In nature, LPL occurs as a homodimer and, when in this
form, binds to cell-surface molecules via several heparin-binding sites, a function that is thought
to increase the cellular uptake of lipoproteins.
Several single nucleotide polymorphisms (SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=4023&view=view&chooseRs=coding&.cgifields=chooseRs]) have been found within the
LPL gene, seven (at the time of writing) of which cause non-synonymous amino acid changes in
the mature protein (Figure 2). At least three of these (rs268, rs328, and rs1801177) have been
associated with observed cases of type I hyperlipoproteinemia (allelic variants [www.ncbi.nlm.nih.
gov/entrez/dispomim.cgi?id=238600&a=238600_AllelicVariant] .0033, .0014, and .0035, respectively).

3-46

Diabetes

Genetic Factors in Type 2 Diabetes

Figure 2: SNP positions of LPL mapped to the 3D structure of lipase.


The figure shows the positions of non-synonymous amino acid changes (green residues) caused by SNPs in the coding
sequence. Disulfide bridges between cysteine residues are shown in orange.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

A BLAST search [www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4557727&cut=100&org=1] using


human LPL as a query finds proteins in 38 different species, which are all metazoans apart from
one bacterium, Trichodesmium erythraeum, an oceanic nitrogen producer. However, potential
true homologous genes have thus far been identified only in the mouse and rat.

LPL and Diabetes: Digest of Recent Articles


For a more complete list of research articles on LPL and diabetes, search PubMed.
A common complication of type 2 diabetes is microalbuminuria, that is, protein being excreted
in the urine because of kidney damage through chronic inflammation. In a study of 386 type 2
diabetic patients, a correlation was found between the presence/severity of microalbuminuria and
genetic variants of LPL (9). Higher concentrations of LDL constituents have been linked previously to progression of microalbuminuria. Furthermore, foam cells (white blood cells that have
"mopped up" excess LDL or VLDL) are observed in many chronic inflammatory diseases, includ3-47

Diabetes

Genetic Factors in Type 2 Diabetes


ing atherosclerosis and kidney disease. These observations point to a possible role for altered
LPL in the progression of inflammatory kidney disease, and therefore the microalbuminuria often
seen in diabetes type 2 patients (9). [More articles like this]
Another study compared the levels of LPL in the blood of 40 type 2 diabetes patients and a
group of healthy individuals prior to heparin injection (10). Because functional LPL is usually
anchored to heparan sulfate proteoglycans on the surface of cells, heparin injection releases the
enzyme into the blood. The amount of LPL in the type 2 diabetes patients was significantly lower
than that of healthy individuals. However, after a course of insulin, the levels of LPL were
increased in the diabetic patients. Because this increase cannot be attributed to released LPL
because no heparin injection was given, it must be caused by the administered insulin (10). This
finding supports previous work showing that insulin plays a role in regulating LPL synthesis (11,
12). [More articles like this]
The H+ allele of the T495G HindIII polymorphism of LPL is associated with coronary heart
disease. This polymorphism was further investigated in 785 Chinese subjects, of which about
60% had been diagnosed with early-onset type 2 diabetes (<40 years old). Within this subset, the
polymorphism was associated with higher plasma triglyceride and lower HDL-cholesterol levels
(13). [More articles like this].
Most recently, SNPs in the 3' end of the LPL gene were found to be associated with insulin
resistance in Mexican Americans (14). The SNPs were analyzed as groups inherited together on
the same chromosome (haplotypes). One haplotype was associated with insulin sensitivity,
another with insulin resistance. Insulin resistance is a pathophysiological determinant of type 2
diabetes. Of note, these same haplotypes were also associated with coronary artery disease,
suggesting that LPL may be a genetic link between diabetes and atherosclerosis (15).

References
1. Mead JR, Irvine SA, Ramji DP. Lipoprotein lipase: structure, function, regulation, and role in disease. J
Mol Med 80:753769; 2002. (PubMed)
2. Deeb SS, Peng RL. Structure of the human lipoprotein lipase gene. Biochemistry 28:41314135; 1989.
(PubMed)
3. Kirchgessner TG, Svenson KL, Lusis AJ, Schotz MC. The sequence of cDNA encoding lipoprotein lipase.
A member of a lipase gene family. J Biol Chem 262:84638466; 1987. (PubMed)
4. Kirchgessner TG et al. Organization of the human lipoprotein lipase gene and evolution of the lipase gene
family. Proc Natl Acad Sci U S A 86:96479651; 1989. (PubMed)
5. Wion KL, Kirchgessner TG, Lusis AJ, Schotz MC, Lawn RM. Human lipoprotein lipase complementary
DNA sequence. Science 235:16381641; 1987. (PubMed)
6. Hide WA, Chan L, Li WH. Structure and evolution of the lipase superfamily. J Lipid Res 33:167178;
1992. (PubMed)
7. Emmerich J et al. Human lipoprotein lipase. Analysis of the catalytic triad by site-directed mutagenesis of
Ser-132, Asp-156, and His-241. J Biol Chem 267:41614165; 1992. (PubMed)
8. Winkler FK, D'Arcy A, Hunziker W. Structure of human pancreatic lipase. Nature 343:771774; 1990.
(PubMed)
9. Mattu RK et al. Lipoprotein lipase gene variants relate to presence and degree of microalbuminuria in
Type II diabetes. Diabetologia 45:905913; 2002. (PubMed)
3-48

Diabetes

Genetic Factors in Type 2 Diabetes

10. Miyashita Y et al. Low lipoprotein lipase mass in preheparin serum of type 2 diabetes mellitus patients
and its recovery with insulin therapy. Diabetes Res Clin Pract 56:181187; 2002. (PubMed)
11. Ong JM, Kirchgessner TG, Schotz MC, Kern PA. Insulin increases the synthetic rate and messenger
RNA level of lipoprotein lipase in isolated rat adipocytes. J Biol Chem 263:1293312938; 1988.
(PubMed)
12. Semenkovich CF et al. Insulin regulation of lipoprotein lipase activity in 3T3-L1 adipocytes is mediated at
posttranscriptional and posttranslational levels. J Biol Chem 264:90309038; 1989. (PubMed)
13. Ma YQ et al. The lipoprotein lipase gene HindIII polymorphism is associated with lipid levels in earlyonset type 2 diabetic patients. Metabolism 52:338343; 2003. (PubMed)
14. Goodarzi MO et al. Lipoprotein lipase is a gene for insulin resistance in Mexican Americans. Diabetes
53:214220; 2004. (PubMed)
15. Goodarzi MO et al. Determination and use of haplotypes: ethnic comparison and association of the
lipoprotein lipase gene and coronary artery disease in Mexican-Americans. Genet Med 5:322327; 2003.
(PubMed)

Link Roundup
Live Searches
Diabetes and LPL in PubMed | PubMed Central | Books

Background Information
Cholesterol levels at the National Heart, Blood, and Lung Institute [www.nhlbi.nih.gov/health/public/heart/chol/wyntk.htm]
Lipoprotein interconversions at the The University of Manitoba [www.umanitoba.ca/faculties/medicine/units/biochem/coursenotes/
blanchaer_tutorials/LipTutWeb/pages/page1.htm]
Type I hyperlipoproteinemia in: OMIM | emedicine [www.emedicine.com/oph/topic505.htm]
Mechanism of hydrolysis [www.ncbi.nlm.nih.gov/books/bv.fcgi?tool=bookshelf&call=bv.View..ShowSection&rid=stryer.
section.1170#1178], foam cells on the Bookshelf

Molecular Biology
Lipoprotein lipase in Entrez Gene | Evidence Viewer [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_030737.8&gene=LPL&graphiconly=TRUE] | Map Viewer | Domains [www.ncbi.nlm.nih.gov/Structure/cdd/
wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4557727]: Lipase, PLAT/LH2 | SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=4023&view=view&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants [www.ncbi.nlm.nih.gov/entrez/dispomim.cgi?
id=238600&a=238600_AllelicVariant] | BLink [www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4557727&cut=100&org=1] | HomoloGene
Hepatic lipase in Entrez Gene
Pancreatic lipase in Entrez Gene | Structure [www.ncbi.nlm.nih.gov/Structure/mmdb/mmdbsrv.cgi?
form=6&db=t&Dopt=s&uid=21535]

3-49

Diabetes

Genetic Factors in Type 2 Diabetes

The Transcription Factor PPARG


Summary
An important diabetes risk factor and drug target is peroxisome proliferator activated receptor
gamma (PPAR). This protein is a member of the nuclear hormone receptor super-family of transcription factors. PPAR is a key regulator of fat cell differentiation. Drugs that activate PPAR
increase sensitivity to insulin and lower blood sugar levels in diabetics. Variants of PPAR influence the risk of developing obesity and type 2 diabetes.

Nomenclature
Official gene name: peroxisome proliferator activated receptor gamma
Official gene symbol: PPARG
Alias: PPARG1, PPARG2

Background
Small can be powerful. Some genetic risk factors may increase the risk of developing a particular
disease by only a small amount. But when the genetic risk factor is common, its effects can tip
the balance toward millions of people developing disease. For type 2 diabetes, one such genetic
risk is the gene that encodes PPAR (peroxisome proliferator activated receptor gamma). Variants of this gene are among the first to be identified as causing a broad impact on the risk of
developing type 2 diabetes (1).
There are three known subtypes of PPARalpha (), delta (), and gamma (). The latter
subtype, PPAR, is abundant in adipose tissue and plays a key role in fat cell differentiation.
PPAR is also the target of the type 2 diabetes drugs called thiazolidinediones (TZDs).
Found in the nucleus of many cells, PPARs are both hormone receptors and transcription
factors. PPARs therefore have two binding sites, one site for ligands (such as fatty acids, hormones, and specific diabetic drugs) and one site for DNA.
To act as a transcription factor, PPAR must first form a complex with another transcription
factor called retinoid X receptor (RXR). When a ligand such as the drug TZD binds to the receptor
PPAR, it activates the PPARRXR complex. The activated complex binds to the promoter
region of specific genes and activates transcription.
To date, several genes have been identified as being direct targets for PPAR, including
lipoprotein lipase (LPL), fatty acid transport protein (FATP), and acetyl CoA-synthase (ACS) (2).
Transcription of these genes influences the metabolism of fatty acids. TZDs such as rosiglitazone
[www.nlm.nih.gov/medlineplus/druginfo/medmaster/a699023.html] can increase insulin-stimulated
glucose disposal by activating the PPARRXR complex primarily in fat cells. In doing so, such
drugs improve sensitivity to insulin and lower blood glucose levels in type 2 diabetes patients.

Molecular Information
Nuclear receptors such as PPAR are one of the largest groups of transcription factors known
today. The members of this family are conserved functionally and structurally and are expressed
in a wide range of multicellular (metazoan) species. A BLAST [http://www.ncbi.nlm.nih.gov/sutils/
3-50

Diabetes

Genetic Factors in Type 2 Diabetes


blink.cgi?pid=20336231&cut=100&org=1] search using human PPARG as a query finds proteins
in 48 different species. However, potential true homologous genes have thus far been identified
only in the mouse and rat.
The PPAR gene maps to chromosome 3 (Figure 1). It has 11 exons (coding regions) that
span more than 140,000 bases (see evidence) [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_022517.16&gene=PPARG&graphiconly=TRUE] (3). There are three isoforms that differ at their 5 ends (4). The isoform type 2 (PPAR2) has an additional 84
nucleotides at its 5 end compared with the isoform type 1 (5) and is also much less abundant
than isoform 1, its highest levels being found in adipose tissue and the colon (4).

Figure 1: Location of PPARG on the human genome.


PPARG maps to chromosome 3, between approximately 12,30012,450 kilobases (kb). Click

or here for a current and

interactive view of the location of PPARG in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of PPARG may fluctuate. The live Web site may, therefore, not appear exactly as in this
figure.

The crystal structure of the ligand-binding domain of PPAR has been solved (6, 7) and
shares many structural features with other nuclear receptors, including a central DNA binding
domain and a ligand-binding domain (LBD) in the C-terminal half of the receptor. The LBD contains 11-13 alpha helices that fold to form a hydrophobic pocket where ligands can bind. The AF2
(activating function-2) helix of PPAR is found here and interacts with RXR and is thought to aid
dimerization and transactivation of the two receptors and also initiates transcription (8).

3-51

Diabetes

Genetic Factors in Type 2 Diabetes


Crystal structures of the heterodimer of PPAR and RXR in complex with their ligands (such
as rosglitazone and 9-cis-retinoic acid, respectively) have also been solved (8, 9). Such structures give us a molecular understanding of how PPAR is able to bind a variety of ligands and aid
the design of drugs that specifically target the gamma receptor.
Several single nucleotide polymorphisms (SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=5468&view=view&chooseRs=coding&.cgifields=chooseRs]) have been found within the
PPARG gene, three (at the time of writing) of which cause non-synonymous amino acid changes
in isoform type 2 of the protein NP_056953. At least one of these (rs1801282) has been associated with obesity (allelic variant .002 [www.ncbi.nlm.nih.gov/entrez/dispomim.cgi?
id=601487&a=601487_AllelicVariant0002]).

PPARG and Diabetes: Digest of Recent Articles


For a more complete list of research articles on PPARG and diabetes, search PubMed.
The variant of the PPAR receptor inherited may affect the risk of obesity or developing type 2
diabetes. Of particular interest is position 12 of the coding region. Most people have the amino
acid proline here; this confers a small risk of developing obesity, about 1.3% (1). For the individual, this 1.3% increase in risk is a small risk, but because 75% of the population have the proline
allele, this translates into a tremendous impact on the number of people developing diabetes (10).
A common variant of the PPAR2 is a single nucleotide polymorphism (SNP) that has alanine
in the place of proline at position 12 (Pro12Ala). Individuals who inherit this variant have a degree
of protection against insulin resistance and obesity (11).
One study screened Pima Indians of Arizona for variations in PPAR2. Type 2 diabetes is
particularly common among this population. Affected individuals are obese and resistant to the
insulin they produce in inadequate amounts. Several new SNPs were identified, many in the promoter region of the gene (12). It is not yet clear which SNPs are the most important in contributing to diabetes in this and in other populations.

References
1. Altshuler D, Hirschhorn JN, Klannemark M et al. The common PPARgamma Pro12Ala polymorphism is
associated with decreased risk of type 2 diabetes. Nature Genetics 26:7680; 2000. (PubMed)
2. Arner PC. The adipocyte in insulin resistance: key molecules and the impact of the thiazolidinediones.
Trends Endocrinol Metab 14:137145; 2003. (PubMed)
3. Fajas L, Auboeuf D, Raspe E et al. The organization, promoter analysis, and expression of the human
PPARgamma gene. J Biol Chem 272:1877918789; 1997. (PubMed)
4. Fajas L, Fruchart JC, Auwerx J. PPARgamma3 mRNA: a distinct PPARgamma mRNA subtype
transcribed from an independent promoter. FEBS Lett 438:5560; 1998. (PubMed)
5. Elbrecht A, Chen Y, Cullinan CA et al. Molecular cloning, expression and characterization of human
peroxisome proliferator activated receptors gamma 1 and gamma 2. Biochem Biophys Res Commun
224:431437; 2003. (PubMed)
6. Nolte RT, Wisely GB, Westin S et al. Ligand binding and co-activator assembly of the peroxisome
proliferator-activated receptor-gamma. Nature 395:137143; 1998. (PubMed)
7. Uppenberg J, Svensson C, Jak M et al. Crystal structure of the ligand binding domain of the human
nuclear receptor PPARgamma. J Biol Chem 273:3110831112; 1998. (PubMed)
3-52

Diabetes

Genetic Factors in Type 2 Diabetes

8. Gampe RT, Montana VG, Lambert MH et al. Asymmetry in the PPARgamma/RXRalpha crystal structure
reveals the molecular basis of heterodimerization among nuclear receptors. Mol Cell 5:545555; 2000.
(PubMed)
9. Cronet P, Cronet P, Petersen JF et al. Structure of the PPARalpha and -gamma ligand binding domain in
complex with AZ 242; ligand selectivity and agonist activation in the PPAR family. Structure (Camb)
9:699706; 2001. (PubMed)
10. Stumvoll M, Haring H. The peroxisome proliferator-activated receptor-gamma2 Pro12Ala polymorphism.
Diabetes 51:23412347; 2002. (PubMed)
11. Li S, Chen W, Srinivasan SR et al. The peroxisome proliferator-activated receptor-gamma2 gene
polymorphism (Pro12Ala) beneficially influences insulin resistance and its tracking from childhood to
adulthood: the Bogalusa Heart Study. Diabetes 52:12651269; 2003. (PubMed)
12. Muller YL, Bogardus C, Beamer BA et al. A functional variant in the peroxisome proliferator-activated
receptor gamma2 promoter is associated with predictors of obesity and type 2 diabetes in Pima Indians.
Diabetes 52:18641871; 2003. (PubMed)

Link Roundup
Live Searches
Diabetes and PPARG in PubMed | PubMed Central | Books

Background Information
PPARG in OMIM

Molecular Biology
PPARG in Entrez Gene | Evidence Viewer [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_022517.16&gene=PPARG&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=20336231]: Ligand-binding domain, Zinc finger | SNPs [www.ncbi.
nlm.nih.gov/SNP/snp_ref.cgi?locusId=5468&view=view&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [http://
www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=20336231&cut=100&org=1] | HomoloGene

3-53

Diabetes

Genetic Factors in Type 2 Diabetes

The Regulatory Subunit of a Phosphorylating Enzyme (PIK3R1)


Summary
The PIK3R1 gene encodes the regulatory subunit of a lipid kinase that has a key role in insulin
signaling. A variant of this regulatory subunit may interfere with the cascade of events that occurs
after insulin binds to its receptor.

Nomenclature
Official gene name: phosphoinositide-3-kinase, regulatory subunit, polypeptide 1 (p85 alpha)
Official gene symbol: PIK3R1
Alias: GRB1, p85-ALPHA

Background
Binding of insulin to its receptor triggers a cascade of phosphorylation and dephosphorylation
reactions beginning at the receptor itself (autophosphorylation) and specific target proteins of the
receptor called insulin receptor substrates (IRSs). The cascade of phosphorylation reactions
spreads to various cellular proteins and is finally terminated when the insulin receptor is dephosphorylated.
Enzymes that catalyze the phosphorylation of proteins are called kinases, and one particular
kinase, phosphatidylinositol 3-kinase (PI3-K), is a major pathway for the metabolic effects of
insulin. It plays a key role in the stimulation of glucose transport, glycogen synthesis, and fat
breakdown.
PI3-K is made up of a catalytic subunit and a regulatory subunit. The regulatory subunit is
encoded by the p85alpha gene, which generates three protein products of 85, 55, and 50 kDa.
Each of the three proteins (p85alpha, p55alpha, and p50alpha) are found in different tissues and
may have specific roles in various tissues (1). The p85alpha protein is the most common protein
isoform in all tissues.
Insulin can initiate multiple signaling pathways, and the PI3-K enzyme is important in the
"RAS-independent pathway". After insulin binding and autophosphorylation of the receptor, IRS-1
binds to a phosphorylated tyrosine residue on the receptor. The regulatory subunit of the PI3-K
enzyme then binds via its Src homology 2 (SH2) domains to the receptor-bound IRS-1. The catalytic subunit phosphorylates specific lipids to produce a second messenger, PIP3 (phosphatidylinositol 3,4,5-phosphate). PIP3 binds to and leads to the activation of protein kinase B,
which in turn promotes the metabolic effects of insulin, promoting glucose uptake (by moving
GLUT4 from intracellular stores to the plasma membrane) and glycogen synthesis (by activating
glycogen kinase 3).
The p85alpha gene is a candidate gene for the development of diabetes because of its role in
insulin signaling. The most predominant protein this gene encodes is the p85alpha protein, but
the p50alpha and p55alpha proteins may also have an important role.

3-54

Diabetes

Genetic Factors in Type 2 Diabetes

Molecular Information
The PIK3R1 gene maps to chromosome 5 (Figure 1). It has 17 exons (coding regions) that span
over 75,000 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_006431.13&gene=PIK3R1&graphiconly=TRUE]).

Figure 1: Location of PIK3R1 on the human genome.


PIK3R1 maps to chromosome 5, between approximately 67,53067,620 kilobases (kb). Click

or here for a current and

interactive view of the location of PIK3R1 in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of PIK3R1 may fluctuate. The live Web site may, therefore, not appear exactly as in this
figure.

Alternative splicing of the p85alpha gene results in three transcript variants that encode three
different protein isoforms. The longest isoform (NP_852664) is 744 amino acids long.
The regulatory subunit of PI3-K contains two SH2 domains that bind to specific phosphorylated tyrosine residues present in motifs possessing the sequence YXXM or YMXM. These motifs
are present on the insulin receptor and in all four of the insulin receptor substrates (IRS-1, -2, -3,

3-55

Diabetes

Genetic Factors in Type 2 Diabetes


and -4). For full activation of the kinase, both SH2 domains of the regulatory subunit need to bind
to the motif present on IRS-1. The resulting conformational change of the regulatory subunit activates the catalytic subunit (2, 3).
Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=5295&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the PIK3R1 gene, and these variants may be involved in the development of diabetes (4). Each of the three protein variants encoded by this gene exhibit two (at the time of
writing) amino acid changes caused by SNPs in the coding region of PIK3R1.
A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=32455248&cut=100&org=1]
using human PIK3R1 as a query finds proteins in 20 different species, which are all metazoans
apart from five proteins found in fungi and one protein found in Archaea (single-cell organisms
that are distinct from bacteria). However, potential true homologous genes have thus far been
identified only in the mouse and rat.

PIK3R1 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on PIK3R1 and diabetes, search PubMed.
A methionine to isoleucine switch at residue 326 (M326I) has been observed in Europeans
and Japanese (5-7). This change occurs just six amino acids from the N-terminal SH2 domain
and appears to affect glucose homeostasis. In the Danish population, the variant was common in
both diabetics and healthy subjects, with homozygous carriers being less effective at glucose
regulation (5). In the Pima Indians, diabetes was less common in homozygous carriers compared
with individuals who were heterozygous or wild type (8).
The M326I variant may lead to a reduced level of protein being expressed (but this may be
offset by an observed increased interaction of the enzyme with IRS-1). The enzyme variant may
also be less efficient in suporting the development of fat cells and in supporting insulin-stimulated
glucose uptake into certain cells (3).
Overall, the M326I variant may only have a minor impact on the signaling events initiated by
the insulin. But if combined with gene variants encoding other signaling proteins or acquired
alterations in protein levels, this enzyme variant may contribute to a functional impact on insulin
signaling and thus contribute to the events that lead to type 2 diabetes.

References
1. Inukai K, Funaki M, Ogihara T et al. p85alpha gene generates three isoforms of regulatory subunit for
phosphatidylinositol 3-kinase (PI 3-Kinase), p50alpha, p55alpha, and p85alpha, with different PI 3-kinase
activity elevating responses to insulin. J Biol Chem 272:78737882; 1997. (PubMed)
2. Rordorf-Nikolic T, Van Horn DJ, Chen D et al. Regulation of phosphatidylinositol 3'-kinase by tyrosyl
phosphoproteins. Full activation requires occupancy of both SH2 domains in the 85-kDa regulatory
subunit. J Biol Chem 270:36623666; 1995. (PubMed)
3. Almind K, Delahaye L, Hansen T et al. Characterization of the Met326Ile variant of phosphatidylinositol 3kinase p85alpha. Proc Natl Acad Sci U S A 99:21242128; 2002. (PubMed)
4. Barroso I, Luan J, Middelberg RP. Candidate gene association study in type 2 diabetes indicates a role
for genes involved in beta-cell function as well as insulin action. PLoS Biol 1:E20E20; 2003. (PubMed)

3-56

Diabetes

Genetic Factors in Type 2 Diabetes

5. Hansen T, Andersen CB, Echwald SM et al. Identification of a common amino acid polymorphism in the
p85alpha regulatory subunit of phosphatidylinositol 3-kinase: effects on glucose disappearance constant,
glucose effectiveness, and the insulin sensitivity index. Diabetes 46:494501; 1997. (PubMed)
6. Hansen L, Zethelius B, Berglund L et al. In vitro and in vivo studies of a naturally occurring variant of the
human p85alpha regulatory subunit of the phosphoinositide 3-kinase: inhibition of protein kinase B and
relationships with type 2 diabetes, insulin secretion, glucose disappearance constant, and insulin
sensitivity. Diabetes 50:690693; 2001. (PubMed)
7. Kawanishi M, Tamori Y, Masugi J et al. Prevalence of a polymorphism of the phosphatidylinositol 3kinase p85 alpha regulatory subunit (codon 326 Met-->Ile) in Japanese NIDDM patients. Diabetes Care
20:10431043; 1997. (PubMed)
8. Baier LJ, Wiedrich C, Hanson RL et al. Variant in the regulatory subunit of phosphatidylinositol 3-kinase
(p85alpha): preliminary evidence indicates a potential role of this variant in the acute insulin response
and type 2 diabetes in Pima women. Diabetes 47:973975; 2003. (PubMed)

Link Roundup
Live Searches
Diabetes and PIK3R1 in PubMed | PubMed Central | Books

Background Information
PIK3R1 in OMIM

Molecular Biology
PIK3R1 in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_006431.13&gene=PIK3R1&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=32455248]: first SH2, second SH2, SH3; GTPase Activator | SNPs
[http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?locusId=5295&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | BLink
[http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=32455248&cut=100&org=1] | HomoloGene

3-57

Diabetes

Other Types of Diabetes

4. Other Types of Diabetes


Created: July 7, 2004

The vast majority of diabetes cases fall into the categories of type 1 and type 2 diabetes. However,
up to 5% of cases have other specific causes and include diabetes that results from the mutation of
a single gene.

4-1

Diabetes

Other Types of Diabetes

Genetic Defects of Beta Cell Function (MODY and Others)


MODY
Maturity onset diabetes in the young (MODY) is an uncommon cause of diabetes that may be
mistaken for type 2 diabetes because treatment of both conditions does not require insulin, at
least in the early stages of the disease.
However, many differences exist between MODY and type 2 diabetes (Table 1). Whereas the
nature of the genetic predispostion of type 2 diabetes is unclear with many susceptibility genes
being implicated, MODY is a monogenic condition (caused by a mutation of a single gene) that is
inherited as an autosomal dominant trait. The onset of diabetes usually occurs in childhood or
adolescence, usually before 25 years of age, although the hyperglycemia is mild in some cases
and may be missed, as with type 2 diabetes. When hyperglycemia is detected in children, MODY
may be misdiagnosed as type 1 diabetes.
Genetic studies have defined a number of subtypes of MODY. Mutations in the genes encoding hepatic nuclear factor 4 (HNF4), glucokinase (GCK), hepatic nuclear factor 1 alpha and 1 beta
(commonly known as HNF1A and HNF1B, but official symbols are TCF1 and TCF2, respectively),
insulin promoter factor 1 (IPF-1), and NEUROD1 are the cause of the six known forms of MODY
(MODY1-6). See Table 2 for a comparison of the MODY-related genes.
MODY2 is caused by a mutant glucokinase enzyme that fails to accurately sense the circulating concentrations of glucose. All of the remaining MODY genes encode transcription factors.
HNF4A, TCF1, TCF2, and IPF-1 form crucial links in the cascade of transcription factors that control the appropriate expression of beta cell genes, such as insulin and the glucose transporter
GLUT2. Mutations of these genes may disrupt the development of beta cells in the embryo and
result in dysfunctioning beta cells in the adult. However, the precise role of these proteins in adult
pancreatic islets is only beginning to be unraveled.
MODY3 and MODY2 are the most common causes of MODY but remain relatively uncommon causes of diabetes.
Table 1. Comparison between type 2 diabetes and MODY
Characteristic

Type 2 diabetes

MODY

Inheritance
Age of onset
Pedigree
Penetrance
Obesity

Polygenic
Usually >40 years of age
Rarely seen across generations
Variable (10-40%)
Usually obese
Usually present

Monogenic, autosomal dominant


Usually <25 years of age
Usually seen across generations
80-90%
Non-obese
Absent

Metabolic syndome1

1 Metabolic syndrome: diabetes, insulin resistance, hypertension, and hypertriglyceridemia.

4-2

Diabetes

Other Types of Diabetes

Table 2. Comparison of MODY-related genes


Type of MODY1

Gene

Molecular basis

MODY1

HNF4A

MODY2

Glucokinase

MODY3

TCF1 (HNF1A)

MODY4

IPF1

MODY5

TCF2 (HNF1B)

MODY6

NeuroD1 (BETA2)

Abnormal regulation of gene transcription in beta


cells causes a defect in the metabolic
signaling of insulin secretion, beta cell mass,
or both.
Reduced phosphorylation of glucose results in a
defect in sensitivity of beta cells to glucose
and a defect in the storage of glucose as
glycogen in the liver.
Abnormal regulation of gene transcription in beta
cells causes a defect in the metabolic
signaling of insulin secretion, beta cell mass,
or both.
Abnormal transcriptional regulation of beta cell
development and function.
Abnormal regulation of gene transcription in beta
cells causes a defect in the metabolic
signaling of insulin secretion, beta cell mass,
or both.
Abnormal transcriptional regulation of beta cell
development and function.

1 MODY, maturity onset diabetes in the young; HNF4A, hepatocyte nuclear factor 4; TCF1, Transcription Factor 1; HNF1A, Hepatocyte
Nuclear Factor 1; IPF1, Insulin Promoter Factor; TCF2, Transcription Factor 2; HNF1B, Hepatocyte Nuclear Factor 1; NeuroD1,
Neurogenic differentiation factor; BETA2, Beta cell E-box transactivator 2.

Other Mutations That Cause Diabetes by Impairing Beta Cell Function


Mutations in mitochondrial DNA are a rare cause of diabetes. Mitochondrial DNA is a circular
molecule that contains 37 genes that are passed on from the mother to her offspring. Paternal
transmission of mitochondrial DNA is thought not to occur, because after fertilization, the fertilized
egg destroys mitochondria derived from the sperm.
Diabetes and hearing loss are associated with a point mutation of mitochondrial DNA. The
mutation occurs in the gene that encodes tRNA leucine, leading to the substitution of guanine for
adenine (AG) at position 3243.
The tRNA Leu 3243 mutation was originally identified in patients with the MELAS syndrome
(mitochondrial myopathy, encephalopathy, lactic acidosis, and stroke-like syndrome); however,
diabetes is not part of this syndrome, suggesting that this mitochondrial mutation may be
expressed as different phenotypes (1).

References
1. Kadowaki T, Kadowaki H, Mori A et al. A subtype of diabetes mellitus associated with a mutation of
mitochondrial DNA. N Engl J Med 330:962968; 1994. (PubMed)

4-3

Diabetes

Other Types of Diabetes

MODY1: Caused by a Mutation in Transcription Factor HNF4A


Summary
The HNF4A gene encodes a transcription factor that is found in the liver and pancreas. HNF4A is
found in a region of chromosome 20 that is linked with type 2 diabetes, and mutations of this
gene cause a rare form of autosomal dominant diabetes (MODY1).

Nomenclature
Official gene name: hepatocyte nuclear factor 4 alpha
Official gene symbol: HNF4A
Alias: transcription factor 14, TCF14, Maturity Onset Diabetes in the Young type 1, MODY1

Background
The expression of a wide range of genes in the liver is regulated by the transcription factor hepatocyte nuclear factor 4 alpha (HNF4A). Many functions that the liver carries out may appear and
disappear, depending on whether HNF4A is expressed. In addition, HNF4A controls the expression of another transcription factor, hepatocyte nuclear factor 1 (HNF1A), which in turn regulates
the expression of several important genes in the liver, including HNF4A.
As the name suggests, HNF4A is found in abundance in the liver, but it is also found in the
beta cells of the pancreas, kidneys, and intestines. Together with other transcription factors such
as HNF1A and HNF1B (encoded by TCF1 and TCF2, respectively), they make up part of a network of transcription factors that functions together to control gene expression in the developing
embryo. In particular, HNF4A is thought to play an important role in the development of the liver,
kidney, and intestines.
In pancreatic beta cells, this network of transcription factors regulates the expression of the
insulin gene. In addition, HNF4 and HNF1 regulate the expression of several other genes linked
with insulin secretion, e.g., genes that encode proteins involved in glucose transport (GLUT2) and
glucose and mitochondrial metabolism (1, 2).
The HNF4A gene is suspected to play a role in type 2 diabetes; inheriting particular HNF4A
variants may alter insulin secretion and predispose toward hyperglycemia. Mutations of HNF4A
can cause an extremely rare form of diabetes, maturity onset diabetes in the young type 1
(MODY1). Whereas type 2 diabetes is a disorder usually of late onset with significant polygenetic
basis, MODY by definition occurs in the young (onset at age less than 25 years) and is a monogenetic disorder inherited in an autosomal dominant fashion.

Molecular Information
Hepatocyte nuclear factors (HNFs) are a heterogeneous class of evolutionarily conserved transcription factors that are required for cellular differentiation and metabolism. HNF4A is an orphan
receptor; the ligand(s) that binds to this receptor is unknown (3).

4-4

Diabetes

Other Types of Diabetes


A BLAST search [www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=31077205&cut=100&org=1]
using human HNF4A as a query finds proteins in 47 different species, which are all multicellular
species (metazoans). However, potential true homologous genes have thus far been identified in
only three species: the mouse, rat, and the nematode Caenorhabditis elegans.
The HNF4A gene maps to chromosome 20 (Figure 1). It has 11 exons (coding regions) that
span over 30,000 bases (see evidence [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_011362.8&gene=HNF4A&graphiconly=TRUE]). There are at least three
different transcript variants of this gene, which encode three different protein isoforms (a, b, and
c). The longest mRNA transcript, NM_000457, encodes the longest HNF4A protein (isoform b),
containing over 450 amino acids.

Figure 1: Location of HNF4A on the human genome.


HNF4A maps to chromosome 20, approximately between 43,70043,750 kilobases (kb). Click

on the figure or here for

a current and interactive view of the location of HNF4A in the human genome.
Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of HNF4A may fluctuate. Therefore, the live Web site may not appear exactly as in this
figure.

Several single nucleotide polymorphisms (SNPs [www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?


locusId=3172&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been found
within the HNF4A gene (Figure 2). At the time of writing, three non-synonymous amino acid
changes caused by SNPs have been observed in the longest protein isoform (isoform b). At
present, none of these SNPs have been associated with either type 2 diabetes or MODY (see
known allelic variants).
4-5

Diabetes

Other Types of Diabetes

Figure 2: SNP positions of HNF4A mapped to the 3D structure of the ligand-binding domain of rat HNF4A.
The figure shows the positions of a non-synonymous amino acid change (Leu-148) caused by a SNP in the coding
sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

HNF4A and MODY1: Digest of Recent Articles


For a more complete list of research articles on HNF4A and MODY1, search PubMed.
Mutations in the HNF4A gene resulting in MODY1 are rare; to date, only 13 families worldwide have been identified as having this form of MODY. Because HNF4A regulates the expression of HNF1A (the cause of MODY3), the mechanisms that underly these forms of MODY are
thought to be similar (4).
Patients with MODY1 or MODY3 primarily have impaired beta cell function (shown by a
defect in glucose-induced insulin secretion), as opposed to a primary defect in insulin activity.
Patients with these forms of MODY present with a mild form of diabetes with a worsening of
hyperglycemia over time, leading to up to 40% of patients requiring insulin. These patients have
the full spectrum of complications of diabetes that is seen in type 1 and type 2 diabetes (4).
In addition to its effects on beta cell function, a deficiency of HNF4A affects the liver. In the
developing embryo, HNF4A is needed for normal liver architecture (5). In the adult, a deficiency
of HNF4A affects lipid synthesis and is associated with reduced serum levels of triglycerides and
lipoproteins (6).
4-6

Diabetes

Other Types of Diabetes

References
1. Stoffel M, Duncan SA. The maturity-onset diabetes of the young (MODY1) transcription factor HNF4alpha
regulates expression of genes required for glucose transport and metabolism. Proc Natl Acad Sci USA
94:1320913214; 1997. (PubMed)
2. Wang H, Maechler P, Antinozzi PA et al. Hepatocyte nuclear factor 4alpha regulates the expression of
pancreatic beta-cell genes implicated in glucose metabolism and nutrient-induced insulin secretion. J Biol
Chem 275:3595335959; 2000. (PubMed)
3. Duncan SA, Navas MA, Dufort D et al. Regulation of a transcription factor network required for
differentiation and metabolism. Science 281:692695; 1998. (PubMed)
4. Fajans SS, Bell GI, Polonsky KS et al. Molecular mechanisms and clinical pathophysiology of maturityonset diabetes of the young. N Engl J Med 345:971980; 2001. (PubMed)
5. Parviz F, Matullo C, Garrison WD et al. Hepatocyte nuclear factor 4alpha controls the development of a
hepatic epithelium and liver morphogenesis. Nat Genet 34:292296; 2003. (PubMed)
6. Shih DQ, Dansky HM, Fleisher M et al. Genotype/phenotype relationships in HNF-4alpha/MODY1:
haploinsufficiency is associated with reduced apolipoprotein (AII), apolipoprotein (CIII), lipoprotein(a), and
triglyceride levels. Diabetes 49:832837; 2000. (PubMed)

Link Roundup
Live Searches
Diabetes and HNF4A in PubMed | PubMed Central | Books

Background Information
HNFA4 in OMIM
MODY1 in OMIM

Molecular Biology
HNF4A in Entrez Gene | Evidence Viewer [www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_011362.8&gene=HNF4A&graphiconly=TRUE] | Map Viewer | Domains [www.ncbi.nlm.nih.gov/Structure/
cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=31077205]: Zinc Finger, Ligand-binding Domain | SNPs [www.ncbi.nlm.nih.gov/
SNP/snp_ref.cgi?locusId=3172&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [www.ncbi.
nlm.nih.gov/sutils/blink.cgi?pid=31077205&cut=100&org=1] | HomoloGene

4-7

Diabetes

Other Types of Diabetes

MODY2: Caused by a Mutation in the Enzyme Glucokinase (GCK)


Summary
Glucokinase, encoded by the GCK gene, catalyzes the first step of glucose metabolism in the
liver. It may also be an important "glucose sensor" in the pancreas. Mutant glucokinase causes a
rare autosomal dominant form of diabetes (MODY2) and may also play a role in type 2 diabetes.

Nomenclature
Official gene name: Glucokinase
Official gene symbol: GCK
Alias: GK, hexokinase 4, HK4, Maturity Onset Diabetes in the Young type 2, MODY2

Background
The enzyme glucokinase catalyzes glucose metabolism in the liver and in the pancreatic beta
cell. Glucokinase traps glucose inside the cell by catalyzing its phosphorylation to produce glucose-6-phosphate. This is the first and rate-limiting step in glycosis, a pathway that produces
energy in the form of ATP from glucose.
Glucokinase (also called hexokinase IV) differs from the other hexokinases that are found in
other tissues. First, glucokinase has a lower affinity for glucose. This allows other organs such as
the brain and muscles to have first call on glucose when their supply is limited. A second feature
is that glucokinase is not inhibited by its product, glucose-6-phosphate. This lack of negative
feedback inhibition enables hepatic glucokinase to remain active while glucose is abundant,
ensuring that the liver can continue removing glucose from the blood ensuring that no glucose
goes to waste.
Glucokinase is proposed to be an important "glucose sensor" in the following way. The rate of
glucose metabolism is determined by the rate of glucose phosphorylation, which is catalyzed by
glucokinase in the liver and pancreas. The liver and pancreas also express glucose transporter-2
(GLUT2), an insulin-independent cellular protein that mediates the transport of glucose into cells.
The capacity of GLUT2 to transport glucose is very high, facilitating rapid equilibrium between
extracellular and intracellular glucose. Thus, in effect, the extracellular glucose concentrations are
sensed by glucokinase.
By catalyzing the rate-limiting step of glucose metabolism in the liver, glucokinase enables
the liver to buffer the rise in glucose that takes place after a meal. In the pancreas, glucokinase is
the glucose sensor for insulin release. The threshold for glucose-stimulated insulin release is
about 5 mmol/l (1). Mutations of GCK that alter this threshold manifest as three different syndromes and highlight the importance of GCK in glucose homeostasis and diabetes:
1. Activating mutations lower the beta cell threshold for insulin release to as low as 1.5
mmol/l of glucose, leading to an increase in insulin release. This manifests as persistent
hyperinsulinemic hypoglycemia of infancy (PHHI) (2).

4-8

Diabetes

Other Types of Diabetes

2. Inactivating mutations raise the beta cell threshold for insulin release. If two alleles
altered by inactivating mutations are inherited, the level of glucose needed to stimulate
insulin release from the pancreas is extremely high. Affected individuals present with diabetes at birth (permanent neonatal diabetes) (3, 4).
3. Maturity onset diabetes in the young, type 2 (MODY2) is caused by inheriting one allele
that has been altered by an inactivating mutation. This partial inactivation leads to an
increase in glucose-stimulated insulin release to about 7 mmol/l. This causes a mild
hyperglycemia that is present at birth but often is only detected in later life (5).

Molecular Information
The hexokinase family consists of several enzymes that are all evolutionarily related. In vertebrates, there are four hexokinases named I to IV. Glucokinase is a distinct member of this family
with a different kinetic profile.
A BLAST search [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4503951&cut=100&org=1]
using human GCK as a query finds proteins in 46 different species, which range from metazoa
(multicellular organisms), fungi, plants, and other eukaryotes. Potential true homologous genes
have thus far been identified in the mouse, rat, fly, mosquito, nematode worm, and the plant
"mouseear cress".
The GCK gene maps to chromosome 7 (Figure 1). It has 12 exons (coding regions) that span
about 46,000 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_007819.14&gene=GCK&graphiconly=TRUE]). There are three GCK transcript variants that differ in their first exons and their expression is tissue specific. One isoform
predominates in the pancreatic beta cells; the other two isoforms are found in the liver. The glucokinase enzyme is found in the outer membrane compartment of mitochondria in these tissues.

4-9

Diabetes

Other Types of Diabetes

Figure 1: Location of GCK on the human genome.


GCK maps to chromosome 7, between approximately 4300044000 kilobases (kb). Click

or here for a current and

interactive view of the location of GCK in the human genome.


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of GCK may fluctuate. The live Web site may, therefore, not appear exactly as in this
figure.

GCK and MODY2: Digest of Recent Articles


For a more complete list of research articles on GCK and MODY2, search PubMed.
Around 200 mutations have been identifed in the GCK gene. No one mutation appears to be
a common cause of MODY2, and the mutations are found throughout the gene (4).
Although MODY is a uncommon cause of diabetes, MODY2 is a common form of this disorder, especially in children with mild hyperglycemia and in women with gestational diabetes and a
family history of diabetes (5). MODY2 has been described in many different populations (6, 7).
Unlike the other forms of diabetes, MODY2 causes a hyperglycemia that is both mild and
nonprogressive. Although present from an early age, hyperglycemia is usually only picked up in
adulthood during screening for other conditions and may be misdiagnosed as type 2 diabetes.
MODY2 can often be treated with diet alone. Only 2% of patients require insulin therapy, and
complications of diabetes are rare (4).

References
1. Matschinsky FM. Regulation of pancreatic beta-cell glucokinase: from basics to therapeutics. Diabetes 51
(3):S394S404; 2002. (PubMed)
2. Froguel P, Vaxillaire M, Sun F et al. Close linkage of glucokinase locus on chromosome 7p to early-onset
non-insulin-dependent diabetes mellitus. Nature 356:162164; 1992. (PubMed)
4-10

Diabetes

Other Types of Diabetes

3. Njolstad PR, Sagen JV, Bjorkhaug L et al. Permanent neonatal diabetes caused by glucokinase
deficiency: inborn error of the glucose-insulin signaling pathway. Diabetes 52:28542860; 2003.
(PubMed)
4. Gloyn AL. Glucokinase (GCK) mutations in hyper- and hypoglycemia: maturity-onset diabetes of the
young, permanent neonatal diabetes, and hyperinsulinemia of infancy. Hum Mutat 22:353362; 2003.
(PubMed)
5. Fajans SS, Bell GI, Polonsky KS. Molecular mechanisms and clinical pathophysiology of maturity-onset
diabetes of the young. N Engl J Med 345:971980; 2001. (PubMed)
6. Cao H, Shorey S, Robinson J et al. GCK and HNF1A mutations in Canadian families with maturity onset
diabetes of the young (MODY). Hum Mutat 20:478479; 2002. (PubMed)
7. Barrio R, Bellanne-Chantelot C, Moreno JC et al. Nine novel mutations in maturity-onset diabetes of the
young (MODY) candidate genes in 22 Spanish families. J Clin Endocrinol Metab 87:25322539; 2002.
(PubMed)

Link Roundup
Live Searches
Diabetes and GCK in PubMed | PubMed Central | Books

Background Information
Glucokinase in OMIM
MODY2 in OMIM

Molecular Biology
Gene name in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_007819.14&gene=GCK&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=15967159] | SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=2645&view+rs+=view+rs+&chooseRs=all&.cgifields=chooseRs] | Allelic Variants | BLink [http://www.ncbi.nlm.nih.gov/sutils/
blink.cgi?pid=4503951&cut=100&org=1] | HomoloGene

4-11

Diabetes

Other Types of Diabetes

MODY3: Caused by a Mutation in Transcription Factor TCF1


Summary
The TCF1 gene encodes a transcription factor that is found in the liver and pancreas and is
important in the development of these and other organs. A mutation of TCF1 causes a rare form
of autosomal dominant diabetes (MODY3).

Nomenclature
Official gene name: Transcription factor 1
Official gene symbol: TCF1
Alias: Hepatic nuclear factor 1 Alpha, HNF1A, Hepatic nuclear factor 1, HNF1, Albumin proximal
factor, LFB1, Maturity Onset Diabetes in the Young type 3, MODY3

Background
TCF1 belongs to a network of transcription factors that co-ordinates the expression of a wide
range of genes in the liver. In the embryo, this network of nuclear proteins guides the development of the liver and continues to be important in the adult. Many functions of the liver appear
and disappear, depending on the expression of these transcription factors.
Although found in highest amounts in the liver, this network of transcription factors is found in
other organs, such as the pancreas and kidney. The transcription factor HNF4A regulates the
expression of TCF1 and may also regulate TCF2.
In pancreatic beta cells, HNF4 and TCF1 regulate the expression of the insulin gene along
with several other genes linked with insulin secretion, e.g., genes that encode proteins involved in
glucose transport (GLUT2) and glucose metabolism (1, 2). Because TCF2 has been found to be
expressed in pancreatic islets, it has been suggested that TCF2 functions with TCF1 to regulate
gene expression in beta cells.
TCF1 and TCF2 share similar domains; they have a similar DNA binding region and dimerization domain. The TCF2 protein is believed to form heterodimers with TCF1, and depending on the
TCF2 isoform, the result may be to activate or inhibit transcription of target genes.
Mutations of HNF4A, TCF1, and TCF2 each cause a distinct form of maturity onset diabetes
in the young (MODY). Mutations of HNF4A cause MODY1, mutations of TCF1 cause MODY3,
and mutations of TCF2 cause MODY5.

Molecular Information
The TCF1 gene is a member of the homeobox family of genes.
In the early 1980s, important developmental genes were identified in the fruit fly. These master control genes lay out the body plan, and if they are mutated, the development of the fly is
disrupted. These genes were called homeobox genes after "homeotic", the description for a shift
in structual development.

4-12

Diabetes

Other Types of Diabetes


Homeobox genes contain a conserved DNA motif of 180 base pairs that encodes a domain (a
homeodomain) of about 60 amino acids. Homeodomains bind to certain regions of DNA and regulate the transcription of many other genes, several of which in turn play a key role in the control of
embryogenesis.
Read more about homeodomains in Gilbert's Developmental Biology [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?
rid=dbio.box.2019]

The TCF1 gene maps to chromosome 12 (Figure 1). It has nine exons (coding regions) that
span about 25,000 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_009775.14&gene=TCF1&graphiconly=TRUE]). The gene encodes a protein of 631 amino acids in length.
Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=6928&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the TCF1 gene, four (at the time of writing) of which cause non-synonymous amino
acid changes in the mature protein (Figure 2). One of these SNPs (rs1169288) has been associated with observed cases of MODY3 (allelic variant .0011).
A BLAST [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=38016909&cut=100&org=1] search
using human TCF1 as a query finds proteins in 31 different species, which are all metazoans
apart from two fungi, one plant, and one bacterium. However, potential true homologous genes
have thus far been identified only in the mouse and rat.

Figure 1: Location of TCF1 on the human genome.


TCF1 maps to chromosome 12, approximately between 119,800119,880 kilobases (kb). Click

on the figure or here for

a current and interactive view of the location of TCF1 in the human genome.
Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of TCF1 may fluctuate. Therefore, the live Web site may not appear exactly as in this
figure.

4-13

Diabetes

Other Types of Diabetes

Figure 2: SNP positions of TCF1 mapped to the 3D structure of chain A of human HNF1A (TCF1) bound to DNA.
The figure shows the position of a non-synonymous amino acid change (Ala 14, the green residue) caused by a SNP in
the coding sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

TCF1 and MODY3: Digest of Recent Articles


For a more complete list of research articles on TCF1 and MODY3, search PubMed.
In many populations, mutations in the TCF1 gene resulting in MODY3 are the most common
cause of MODY. Because the expression of TCF1 is regulated by HNF4A (the cause of MODY1),
the mechanisms that underly these forms of MODY are thought to be similar (3).
Patients with MODY1 or MODY3 primarily have impaired beta cell function (shown by a
defect in glucose-induced insulin secretion), as opposed to a primary defect in insulin activity (3).
Patients with these forms of MODY present with a mild form of diabetes with a worsening of
hyperglycemia over time. Up to 40% of patients require insulin, and the full spectrum of diabetes
complications that are seen in type 1 and type 2 diabetes may develop.
In animal models, a deficiency of the TCF1 gene causes hyperglycemia because of defective
beta cell signaling, leading to defective insulin secretion (4). The beta cell dysfunction in MODY3
may be caused by loss-of-function mechanisms, such as reduced DNA binding and impaired
transcriptional activation (5).
In addition to its effects on beta cell function, a deficiency of TCF1 affects the kidneys and
genital system. Patients with TCF1 mutations have decreased renal reabsorption of glucose and
excrete glucose in the urine (glycosuria) (6).
4-14

Diabetes

Other Types of Diabetes


Several of the mRNA transcripts encoded by the mutant HNF1A genes are unstable, suggesting that haploinsufficiency (deficient amounts of gene product) of HNF1A is responsible for the
pathogenesis of MODY3 (7, 8). Haploinsufficiency for HNF1A may also cause diabetes that may
be misdiagnosed as type 1 diabetes but is not caused by an autoimmune attack (idiopathic type 1
diabetes) (9).

References
1. Stoffel M, Duncan SA. The maturity-onset diabetes of the young (MODY1) transcription factor HNF4alpha
regulates expression of genes required for glucose transport and metabolism. Proc Natl Acad Sci USA
94:1320913214; 1997. (PubMed)
2. Wang H, Maechler P, Antinozzi PA et al. Hepatocyte nuclear factor 4alpha regulates the expression of
pancreatic beta-cell genes implicated in glucose metabolism and nutrient-induced insulin secretion. J Biol
Chem 275:3595335959; 2000. (PubMed)
3. Fajans SS, Bell GI, Polonsky KS. Molecular mechanisms and clinical pathophysiology of maturity-onset
diabetes of the young. N Engl J Med 345(13):971980; 2001. (PubMed)
4. Dukes ID, Sreenan S, Roe MW et al. Defective pancreatic beta-cell glycolytic signaling in hepatocyte
nuclear factor-1alpha-deficient mice. J Biol Chem 273:2445724464; 1998. (PubMed)
5. Bjorkhaug L, Sagen JV, Thorsby P et al. Hepatocyte nuclear factor-1 alpha gene mutations and diabetes
in Norway. J Clin Endocrinol Metab 88:920931; 2003. (PubMed)
6. Pontoglio M, Prie D, Cheret C et al. HNF1alpha controls renal glucose reabsorption in mouse and man.
EMBO Rep 1:359365; 2000. (PubMed)
7. Thomas H, Badenberg B, Bulman M et al. Evidence for haploinsufficiency of the human HNF1alpha gene
revealed by functional characterization of MODY3-associated mutations. Biol Chem 383:16911700;
2002. (PubMed)
8. Harries LW, Hattersley AT, Ellard S. Messenger RNA transcripts of the hepatocyte nuclear factor-1alpha
gene containing premature termination codons are subject to nonsense-mediated decay. Diabetes
53:500504; 2004. (PubMed)
9. Yoshiuchi I, Yamagata K, Yoshimoto M et al. Analysis of a non-functional HNF-1alpha (TCF1) mutation in
Japanese subjects with familial type 1 diabetes. Hum Mutat 18:345351; 2001. (PubMed)

Link Roundup
Live Searches
Diabetes and HNF1A in PubMed | PubMed Central | Books

Background Information
TCF1 in OMIM
MODY3 in OMIM

Molecular Biology
TCF1 in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_009775.14&gene=TCF1&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=38016909]: N terminus domain, Homeodomain, C terminus domain
(beta isoform), C terminus domain (alpha isoform) | SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?locusId=6927&view+rs
+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?
pid=38016909&cut=100&org=1] | HomoloGene

4-15

Diabetes

Other Types of Diabetes

MODY4: Caused by a Mutation in Transcription Factor IPF1


Summary
Insulin promotor factor-1 (IPF1) is responsible for the development of the pancreas in the embryo
and is also a key regulator of insulin gene expression. Mutations of IPF1 cause a rare form of
autosomal dominant diabetes (MODY4) and may play a role in susceptibility to type 2 diabetes.

Nomenclature
Official gene name: Insulin promoter factor 1
Official gene symbol: IPF1
Alias: Homeodomain transcription factor, Islet/Duodenum homeobox-1, IDX-1, Somatostatin transcription factor 1, STF-1, Pancreas duodenum homeobox-1, PDX-1, Maturity Onset Diabetes in
the Young type 4, MODY4

Background
In the developing embryo, the pancreas is formed from two buds of the primitive gut that eventually fuse together to form the pancreas gland. The exocrine part of the pancreas consists of cells
that produce digestive enzymes, such as proteases and lipases, that are delivered to the gut via
pancreatic ducts. The endocrine pancreas is much smaller and mainly consists of three cell types
alpha, beta, and deltathat produce glucagon, insulin, and somatostatin, respectively.
The development of the pancreas has been well studied, and many transcription factors can
be used to identify pancreatic cells at different stages of development. Insulin promoter factor-1
(IPF1) is one such transcription factor and is an early pancreatic marker that is also found in adult
beta cells. At a slightly later stage in pancreas development, TCF1 is expressed and is also found
in adult beta cells.
In the embryo, the presence of IPF1 is vital to ensure the correct development of the pancreas. Loss of both copies of the gene can cause the pancreas not to form (pancreas agenesis).
Without IPF1, the proliferation and differentiation of precursor cells into the endocrine and
exocrine parts of the pancreas are blocked (1, 2).
IPF1 continues to be essential for normal pancreatic function in the adult. IPF1 regulates the
expression of several pancreatic genes, most notably insulin (INS), glucose transporter type 2
(GLUT2), glucokinase (GCK), and somatostatin. Loss of one copy of IPF1 has been linked to
MODY4 and may play a role in susceptibility to type 2 diabetes.
The dual role of IPF1 during the development of the pancreas in the embryo and as a regulator of pancreatic genes in the adult underscores the importance of IPF1 in glucose homeostasis.

Molecular Information
The IPF1 gene is a member of the homeobox family of genes.
In the early 1980s, important developmental genes were identified in the fruit fly. These master control genes lay out the body plan, and if they are mutated, the development of the fly is
disrupted. These genes were called homeobox genes after "homeotic", the description for a shift
4-16

Diabetes

Other Types of Diabetes


in structural development. Homeobox genes such as IPF1 are important in determining cell fates;
in the embryo, the presence of IPF1 ensures that pancreatic precursor cells develop into their
destined mature pancreatic cells.
Homeobox genes contain a conserved DNA motif of 180 base pairs that encodes a domain (a
homeodomain) of about 60 amino acids. Homeodomains bind to certain regions of DNA and regulate the transcription of many other genes, several of which in turn play a key role in the control of
embryogenesis.
Read more about homeodomains in Gilbert's Developmental Biology [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?
rid=dbio.box.2019]

The IPF1 gene maps to chromosome 13 (Figure 1). It has two exons (coding regions) that
span about 6,000 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_024524.13&gene=IPF1&graphiconly=TRUE]). The gene encodes a protein of 283 amino acids in length.

Figure 1: Location of IPF1 on the human genome. The IPF1 gene maps to chromosome 13, approximately between
26,28526,300 kilobases (kb). Click

on the figure or here for a current and interactive view of the location of IPF1 in the

human genome.

4-17

Diabetes

Other Types of Diabetes


Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of IPF1 may fluctuate. Therefore, the live Web site may not appear exactly as in this
figure.

The IPF1 gene sequence is highly conserved throughout evolution. Even a species that is
distant in evolution from humans, such as the zebrafish, has a homeodomain that shares 95%
homology with animal IPF1 homeodomains (3). A BLAST search [http://www.ncbi.nlm.nih.gov/
sutils/blink.cgi?pid=4557673&cut=100&org=1] using human IPF1 as a query finds proteins in 47
different species, which are all metazoans (multicellular organisms). However, potential true
homologous genes have thus far been identified only in the mouse and the plant Arabidopsis
thaliana (the first plant for which the complete genome has been sequenced).

IPF1 and Diabetes: Digest of Recent Articles


For a more complete list of research articles on IPF1 and diabetes, search PubMed.
Mutations in the gene that encodes IPF1 are a rare cause of MODY; in fact, the current
understanding of MODY 4 is based on studies of a single family (4).
In this family, an infant was born with agenesis of the pancreas that resulted in permanent
neonatal diabetes and lack of pancreatic digestive enzymes. The infant was found to be homozygous for a deletion mutation in IPF1 that caused a frameshift. The resulting truncated protein did
not contain the homeodomain that is essential for DNA binding (5).
Family members who were heterozygous for the same mutation had a mild form of diabetes
(now called MODY4) that was being treated with either diet alone or oral hypoglycemic agents.
Being heterozygous carriers for the IPF1 mutation was linked with severely impaired insulin
secretion. Affected family members could be traced back to six generations. Compared with other
forms of MODY, the expression of this form of diabetes may occur at later ages (2).
Further investigations into the role of IPF1 in the developing pancreas and in the functioning
of the adult pancreas will improve our understanding of how beta cell dysfunction arises and
leads to diabetes. The nature of IPF1 is also important when considering how to produce functional beta cells that can be transplanted in the hope that such cells can continue secreting
insulin, thus providing a cure for diabetes (6).

References
1. Jonsson J, Carlsson L, Edlund T et al. Insulin-promoter-factor 1 is required for pancreas development in
mice. Nature 371:606609; 1994. (PubMed)
2. Stoffers DA, Zinkin NT, Stanojevic V et al. Pancreatic agenesis attributable to a single nucleotide deletion
in the human IPF1 gene coding sequence. Nat Genet 15:106110; 1997. (PubMed)
3. Milewski WM, Duguay SJ, Chan SJ et al. Conservation of PDX-1 structure, function, and expression in
zebrafish. Endocrinology 139:14401449; 1998. (PubMed)
4. Fajans SS, Bell GI, Polonsky KS. Molecular mechanisms and clinical pathophysiology of maturity-onset
diabetes of the young. N Engl J Med 345:971980; 2001. (PubMed)
5. Stoffers DA, Stanojevic V, Habener JF. Insulin promoter factor-1 gene mutation linked to early-onset type
2 diabetes mellitus directs expression of a dominant negative isoprotein. J Clin Invest 102:232241;
1998. (PubMed)
4-18

Diabetes

Other Types of Diabetes

6. Edlund H. Pancreatic organogenesis--developmental mechanisms and implications for therapy. Nat Rev
Genet 3:524532; 2002. (PubMed)

Link Roundup
Live Searches
Diabetes and IPF1 in PubMed | PubMed Central | Books

Background Information
IPF1 in OMIM
MODY4 in OMIM

Molecular Biology
IPF1 in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_024524.13&gene=IPF1&graphiconly=TRUE] | MapViewer | Homeobox domain [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4557673] | SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?
locusId=3651] | Allelic Variants | BLink [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4557673&cut=100&org=1] | Homologene

4-19

Diabetes

Other Types of Diabetes

MODY5: Caused by a Mutation in Transcription Factor TCF2


Summary
TCF2 encodes a transcription factor that is found in the liver and pancreas and is important in the
development of these and other organs. A mutation of TCF2 causes a rare form of autosomal
dominant diabetes (MODY5).

Nomenclature
Official gene name: Transcription factor 2
Official gene symbol: TCF2
Alias: Hepatic nuclear factor 1 Beta, HNF1B, Hepatic nuclear factor 2, HNF2, Transcription factor,
liver-specific; Variant hepatic nuclear factor, VHNF1, LFb3, Maturity Onset Diabetes in the Young
type 5, MODY5

Background
TCF2 belongs to a network of transcription factors that co-ordinates the expression of a wide
range of genes in the liver. In the embryo, this network of nuclear proteins guides the development of the liver and continues to be important in the adult. Many functions of the liver appear
and disappear, depending on the expression of these transcription factors.
Although found in highest amounts in the liver, this network of transcription factors is found in
other organs, such as the pancreas and kidney. The transcription factor HNF4A regulates the
expression of TCF1 and may also regulate TCF2.
In pancreatic beta cells, HNF4 and TCF1 regulate the expression of the insulin gene along
with several other genes linked with insulin secretion, e.g., genes that encode proteins involved in
glucose transport and glucose metabolism (1, 2). Because TCF2 has been found to be expressed
in pancreatic islets, it has been suggested that TCF2 functions with TCF1 to regulate gene
expression in the beta cells.
TCF1 and TCF2 share similar domains; they have a similar DNA binding region and dimerization domain. The TCF2 protein is believed to form heterodimers with TCF1, and depending on the
TCF2 isoform, the result may be to activate or inhibit transcription of target genes.
Mutations of HNF4A, TCF1, and TCF2 each cause a distinct form of maturity onset diabetes
in the young (MODY). Mutations of HNF4A cause MODY1, mutations of TCF1 cause MODY3,
and mutations of TCF2 cause MODY5.

Molecular Information
The TCF2 gene is a member of the homeobox family of genes.
In the early 1980s, important developmental genes were identified in the fruit fly. These master control genes lay out the body plan, and if they are mutated, the development of the fly is
disrupted. These genes were called homeobox genes after "homeotic", the description for a shift
in structural development. In the embryo, the presence of TCF2 is needed for the correct development of the kidneys and genital system.
4-20

Diabetes

Other Types of Diabetes


Homeobox genes contain a conserved DNA motif of 180 base pairs that encodes a domain (a
homeodomain) of about 60 amino acids. Homeodomains bind to certain regions of DNA and regulate the transcription of many other genes, several of which in turn play a key role in the control of
embryogenesis.
Read more about homeodomains in Gilbert's Developmental Biology [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?
rid=dbio.box.2019]

The TCF2 gene maps to chromosome 17 (Figure 1). It has nine exons (coding regions) that
span about 60,000 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_078100.1&gene=TCF2&graphiconly=TRUE]). The gene encodes a protein of 557 amino acids in length.
There are at least two transcript variants of TCF2; a third variant has been identified in the rat
but not yet in man. Transcript variant a encodes protein isoform a, which predominates in the liver
and stimulates transcription. Transcript variant b contains an intron that is spliced out of other
variants, and as a result the encoded protein, isoform b, has a distinct C terminus. Isoform b
appears to be unable to stimulate transcription and instead inhibits the transactivation activity of
TCF1 (3).
Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=6928&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the TCF2 gene, two (at the time of writing) of which cause non-synonymous amino
acid changes in the mature protein (Figure 2). One of these SNPs (rs1800575) has been associated with observed cases of MODY5 (allelic variant .0001).
A BLAST [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4507397&cut=100&org=1] search
using human TCF2 isoform a as a query finds proteins in 23 different species, which are all metazoans apart from two fungi and one bacterium. However, potential true homologous genes have
thus far been identified only in the mouse and rat.

4-21

Diabetes

Other Types of Diabetes

Figure 1: Location of TCF2 on the human genome.


TCF2 maps to chromosome 17, approximately between 36,22036,320 kilobases (kb). Click

on the figure or here for a

current and interactive view of the location of TCF2 in the human genome.
Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of TCF2 may fluctuate. Therefore, the live Web site may not appear exactly as in this
figure.

4-22

Diabetes

Other Types of Diabetes

Figure 2: SNP positions of TCF2 mapped to the 3D structure of human TCF1 bound to DNA.
The figure shows the position of a non-synonymous amino acid change (Arg87, green residue) caused by a SNP in the
coding sequence.

Click on the figure or this Cn3D icon for a dynamic view (you will need to download the Cn3D viewer [www.ncbi.
nlm.nih.gov/Structure/CN3D/cn3d.shtml] to do this)

TCF2 and MODY5: Digest of Recent Articles


For a more complete list of research articles on TCF2 and MODY5, search PubMed.
Mutations in TCF2 are associated with MODY5, severe renal defects, and genital malformations (4).
Although renal disease is a common complication of diabetes, the renal disease that is associated with TCF2 mutations and MODY5 appears to be a direct result of the TCF2 mutation rather
than a complication of hyperglycemia (5, 6).
Several mutations of TCF2 have been identified, many of which involve a deletion that disrupts the DNA binding domain (7). One such mutation was found in a Norwegian family in which a
deletion in exon 2 resulted in a protein that lacked amino acids Arg-137 to Lys-161. Affected
members had mild diabetes and non-diabetic renal disease that was worsening; both are features
of MODY5. In addition, two affected female carriers of this mutation had an undeveloped vagina
and uterus (8).
This mutant TCF2 protein was unable to bind to TCF1 and could not stimulate transcription of
a target gene, indicating that this was a loss-of-function mutation.

4-23

Diabetes

Other Types of Diabetes


A deletion mutation of TCF2 that spared the DNA binding domain was found to encode a protein with increased transactivation potential, suggesting that this was a gain-of-function mutation
(9).
When both types of human mutant TCF2 were overexpressed in the embryo of the developing frog, they both interrupted the proper development of the kidney. This reflects the different
types of renal disease that are seen in individuals with different types of TCF2 mutations. These
findings imply that TCF2 has a central role in normal kidney development (9, 10).

References
1. Stoffel M, Duncan SA. The maturity-onset diabetes of the young (MODY1) transcription factor HNF4alpha
regulates expression of genes required for glucose transport and metabolism. Proc Natl Acad Sci
94:1320913214; 1997. (PubMed)
2. Wang H, Maechler P, Antinozzi PA et al. Hepatocyte nuclear factor 4alpha regulates the expression of
pancreatic beta-cell genes implicated in glucose metabolism and nutrient-induced insulin secretion. J Biol
Chem 275:3595335959; 2000. (PubMed)
3. Bach I, Mattei MG, Cereghini S et al. Two members of an HNF1 homeoprotein family are expressed in
human liver. Nucleic Acids Res 19:35533559; 1991. (PubMed)
4. Fajans SS, Bell GI, Polonsky KS. Molecular mechanisms and clinical pathophysiology of maturity-onset
diabetes of the young. N Engl J Med 345:971980; 2001. (PubMed)
5. Nishigori H, Yamada S, Kohama T et al. Frameshift mutation, A263fsinsGG, in the hepatocyte nuclear
factor-1beta gene associated with diabetes and renal dysfunction. Diabetes 47:13541355; 1998.
(PubMed)
6. Bingham C, Bulman MP, Ellard S et al. Mutations in the hepatocyte nuclear factor-1beta gene are
associated with familial hypoplastic glomerulocystic kidney disease. Am J Hum Genet 68:219224; 2001.
(PubMed)
7. Horikawa Y, Iwasaki N, Hara M et al. Mutation in hepatocyte nuclear factor-1 beta gene (TCF2)
associated with MODY. Nat Genet 17:384385; 1997. (PubMed)
8. Lindner TH, Njolstad PR, Horikawa Y et al. A novel syndrome of diabetes mellitus, renal dysfunction and
genital malformation associated with a partial deletion of the pseudo-POU domain of hepatocyte nuclear
factor-1beta. Hum Mol Genet 8:20012008; 1999. (PubMed)
9. Wild W, Pogge von Strandmann E, Nastos A et al. The mutated human gene encoding hepatocyte
nuclear factor 1beta inhibits kidney formation in developing Xenopus embryos. Proc Natl Acad Sci U S A
97:46954700; 2000. (PubMed)
10. Bingham C, Ellard S, Allen L et al. Abnormal nephron development associated with a frameshift mutation
in the transcription factor hepatocyte nuclear factor-1 beta. Kidney Int 57:898907; 2000. (PubMed)

Link Roundup
Live Searches
Diabetes and TCF2 in PubMed | PubMed Central | Books

Background Information
TCF2 in OMIM
MODY5 in OMIM

Molecular Biology

4-24

Diabetes

Other Types of Diabetes

TCF2 name in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?


taxid=9606&contig=NT_078100.1&gene=TCF2&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/Structure/
cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=6031205]: N terminus domain, Homeodomain | SNPs [http://www.ncbi.nlm.nih.
gov/SNP/snp_ref.cgi?locusId=6928&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink [http://
www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4507397&cut=100&org=1] | HomoloGene

4-25

Diabetes

Other Types of Diabetes

MODY6: Caused by a Mutation in Transcription Factor NEUROD1


Summary
The transcription factor NEUROD1 can directly activate the transcription of the insulin gene. It is
also needed in the development of the pancreas beta cells and the nervous system. Mutations of
this gene cause MODY6, the most recently discovered form of autosomal dominant diabetes, and
may also play a role in type 1 and type 2 diabetes.

Nomenclature
Official gene name: Neurogenic differentiation 1
Official gene symbol: NEUROD1
Alias: NEUROD, NeuroD, BETA2, BHF-1, Maturity Onset Diabetes in the Young type 6, MODY6

Background
The development and the normal functioning of the endocrine pancreas are dependent on a network of transcription factors. These proteins influence the transcription of genes in a negative or
positive way. The NEUROD1 gene encodes a transcription factor that is a positive regulator for
the transcription of the insulin gene.
NEUROD1 (for "neurogenic differentiation") is a protein that was first discovered to be important in the development of the embyronic nervous system. The expression of NEUROD1 stimulates neurons to mature, or differentiate, and it has the potential to convert undifferentiated cells
into neurones.
In animal models, mutations of the NEUROD1 gene disrupts the normal development of the
pancreas, leading to diabetes (1). Certain structures in the brain, such as the cerebellum and hippocampus, also fail to develop properly, resulting in seizures (2, 3).
The link between NEUROD1 and diabetes was first suggested when it was discovered that
the NEUROD1 gene is located in the same region of a chromosome that is linked with type 1 diabetes susceptibility (4). This region is called IDDM7 and is found on the short arm of chromosome
2.
In addition to the link with type 1 diabetes, variants of NEUROD1 have also been linked with
susceptibility to type 2 diabetes (5), and a mutation of NEUROD1 causes the most recently discovered form of autosomal dominant diabetes, maturity onset diabetes in the young type 6
(MODY6) (5, 6).

Molecular Information
NEUROD1 belongs to a group of transcription factors called basic helix-loop-helix (bHLH) proteins. The bHLH proteins contain a conserved sequence of amino acids that binds to DNA. This
sequence is also known as a DNA-binding motif, and the HLH motif consists of a short alpha helix
connected by a flexible loop to a second, longer alpha helix (see the HLH domain [http://www.
ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4505377]).

4-26

Diabetes

Other Types of Diabetes


bHLH proteins are classified into two groups based on how they bind to DNA and in what
tissues they are found. Class A members tend to be expressed in all tissues, whereas class B
members, such as NEUROD1, are found only in specific tissues, mainly in the nervous system
and the pancreas.
bHLH proteins can function as transcription factors only when two bHLH monomers complex
to form a dimer. The two-helix structure of HLH binds both to DNA and to the HLH motif of a second HLH protein. The second HLH protein can be the same (resulting in a homodimer) or different (resulting in a heterodimer), and alpha helices extending from the dimerization interface make
specific contacts with DNA.
The NEUROD1 gene maps to chromosome 2 (Figure 1). It has two exons (coding regions)
that span about 4,860 bases (see evidence [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_005403.14&gene=NEUROD1&graphiconly=TRUE]); only exon 2 is translated (7). The gene encodes a protein of 356 amino acids.
Several single nucleotide polymorphisms (SNPs [http://www.ncbi.nlm.nih.gov/SNP/snp_ref.
cgi?locusId=4760&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs]) have been
found within the NEUROD1 gene, three (at the time of writing) of which cause non-synonymous
amino acid changes in the mature proteins.
A BLAST [http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4505377&cut=100&org=1] search
using human NEUROD1 as a query finds proteins in 24 different species, which are all metazoans (multicellular). However, potential true homologous genes have thus far been identified
only in the mouse, rat, and roundworm.

Figure 1: Location of NEUROD1 on the human genome.


NEUROD1 maps to chromosome 2, approximately between 18,273018,2760 kilobases (kb). Click

on the figure or

here for a current and interactive view of the location of NEUROD1 in the human genome.
Note: this figure was created from Build 34 of the human genome. Because the data are recomputed between
genome builds, the exact location of NEUROD1 may fluctuate. Therefore, the live Web site may not appear exactly as in
this figure.

4-27

Diabetes

Other Types of Diabetes

NEUROD1 and MODY6: Digest of Recent Articles


For a more complete list of research articles on NEUROD1 and MODY6, search PubMed.
NEUROD1, after its heterodimerization with the HLH protein E37, regulates transcription of
the insulin gene. NEUROD1 binds to the E-box motif of the insulin gene promoter. It is proposed
that deficient binding of NEUROD1 or binding of transcriptionally inactive NEUROD1 to target
promoters in pancreatic islets leads to the development of diabetes (5).
Mutations in NEUROD1 have been found in three families to date, and these mutations are
associated with type 2 diabetes and MODY (8).
In one family, a G T substitution in codon 111 caused a switch in amino acids at this position from arginine to leucine (Arg111Leu). The Arg-111 residue is found in the DNA-binding
domain of NEUROD1 and has been evolutionarily conserved from the fruit fly to mammals and is
found in all members of the HLH family of transcription factors (1). In this family, the Arg111Leu
mutation was associated with type 2 diabetes; of the six carriers of the mutation, four were diagnosed with diabetes in their mid-40s, and two had impaired glucose tolerance.
A second family had an insertion of a cytosine residue in codon 206 (206+C), resulting in a
frameshift mutation. The truncated protein that was synthesized lacked the C-terminal third of the
protein, which includes the transactivation domain (9). Of the nine carriers of the 206+C mutation,
seven had diabetes. The nature of the diabetes observed in the second family was different in
several ways: the diabetes was diagnosed at an earlier age and was more severe (2 of the 206
+C carriers required treatment with insulin), and the affected individuals were not obese and had
low insulin levels. The early onset and severity of diabetes resemble MODY rather than type 2
diabetes. Thus, mutations in NEUROD1 are proposed to be the cause of a new subtype of
MODY, designated MODY6.

References
1. Naya FJ, Huang HP, Qiu Y. Diabetes, defective pancreatic morphogenesis, and abnormal
enteroendocrine differentiation in BETA2/neuroD-deficient mice. Genes Dev 11:23232334; 1997.
(PubMed)
2. Miyata T, Maeda T, Lee JE. NeuroD is required for differentiation of the granule cells in the cerebellum
and hippocampus. Genes Dev 13:16471652; 1999. (PubMed)
3. Liu M, Pereira FA, Price SD. Essential role of BETA2/NeuroD1 in development of the vestibular and
auditory systems. Genes Dev 14:28392854; 2000. (PubMed)
4. Tamimi R, Steingrimsson E, Copeland NG. The NEUROD gene maps to human chromosome 2q32 and
mouse chromosome 2. Genomics 34:418421; 1996. (PubMed)
5. Malecki MT, Jhala US, Antonellis A. Mutations in NEUROD1 are associated with the development of type
2 diabetes mellitus. Nat Genet 23:323328; 1999. (PubMed)
6. Fajans SS, Bell GI, Polonsky KS. Molecular mechanisms and clinical pathophysiology of maturity-onset
diabetes of the young. N Engl J Med 345:971980; 2001. (PubMed)
7. Furuta H, Horikawa Y, Iwasaki N. Beta-cell transcription factors and diabetes: mutations in the coding
region of the BETA2/NeuroD1 (NEUROD1) and Nkx2.2 (NKX2B) genes are not associated with maturityonset diabetes of the young in Japanese. Diabetes 47:13561358; 1998. (PubMed)
8. Kristinsson SY, Thorolfsdottir ET, Talseth B. MODY in Iceland is associated with mutations in HNF1alpha and a novel mutation in NeuroD1. Diabetologia 44:20982103; 2001. (PubMed)
4-28

Diabetes

Other Types of Diabetes

9. Qiu Y, Sharma A, Stein R. p300 mediates transcriptional stimulation by the basic helix-loop-helix
activators of the insulin gene. Mol Cell Biol 18:29572964; 1998. (PubMed)

Link Roundup
Live Searches
Diabetes and NEUROD1 in PubMed | PubMed Central | Books

Background Information
NEUROD1 in OMIM
MODY6 in OMIM

Molecular Biology
NEUROD1 name in Entrez Gene | Evidence Viewer [http://www.ncbi.nlm.nih.gov/sutils/evv.cgi?
taxid=9606&contig=NT_005403.14&gene=NEUROD1&graphiconly=TRUE] | Map Viewer | Domains [http://www.ncbi.nlm.nih.gov/
Structure/cdd/wrpsb.cgi?INPUT_TYPE=precalc&SEQUENCE=4505377]: Helix-Loop-Helix (HLH) Domain | SNPs [http://www.ncbi.
nlm.nih.gov/SNP/snp_ref.cgi?locusId=4760&view+rs+=view+rs+&chooseRs=coding&.cgifields=chooseRs] | Allelic Variants | BLink
[http://www.ncbi.nlm.nih.gov/sutils/blink.cgi?pid=4505377&cut=100&org=1] | HomoloGene

4-29

Diabetes

Other Types of Diabetes

Genetic Defects in Insulin Action


A defect in the action of insulin, or the body being resistant to insulin, occurs when a given
amount of insulin produces a subnormal biological response. Obesity is by far the most common
cause of insulin resistance; however, genetic defects that disrupt the action of insulin may also
cause insulin resistance (Table 1).
One symptom of severe insulin resistance is a skin disorder called acanthosis nigricans.
Areas of the skin have abnormally increased coloration (hyperpigmentation) and "velvety" thickening (hyperkeratosis). This is particularly noticable in skin fold regions, such as of the neck and
groin and under the arms. Acanthosis nigricans is commonly seen in inherited syndromes of
severe insulin resistance.

Type A Insulin Resistance


A previous classification of severe insulin resistance recognized two syndromes, type A and type
B. Type A insulin resistance is inherited in a dominant manner (1), and in a minority of cases, a
mutation in the insulin receptor gene can be isolated (2). Type B insulin resistance is not inherited; instead, it is caused by anti-insulin receptor antibodies and is often seen in older females
with signs of autoimmune disease.
Type A Insulin Resistance in OMIM

Individuals with type A insulin resistance often have signs of polycystic ovarian syndrome:
increased virilization caused by high levels of androgen hormones (hyperandrogenemia), a disruption of the menstrual cycle (oligomenorrhea), and an increase in body hair (hirsutism).

Leprechaunism
Leprechaunism is an autosomal recessive disorder attributable to a defect in the insulin receptor
(INSR). In 1954, the first cases were described by Donohue and Uchida, hence the alternative
name for this condition, Donahue's syndrome (2).
The clinical features of Leprechaunism include an "elfin-like" facial appearance with protuberant ears and relatively large hands and feet. A decreased amount of subcutaneous fat and muscle mass is seen, and the skin is abnormal with increased hair growth. Acanthosis nigricans is
also often present. This condition is usually fatal within the first couple of years of life.
Leprechaunism in OMIM

A number of mutations of the INSR have been found to cause leprechaunism. Among these
are mutations that: cause a premature chain termination in the alpha subunit, thereby deleting the
transmembrane and tyrosine kinase domains of the receptor (4); impair INSR dimerization and
transport to the cell surface (5); and impair autophosphorylation of INSR (6).

4-30

Diabetes

Other Types of Diabetes

Rabson-Mendenhall Syndrome
In 1956, a pathologist, Dr. Rabson, and a family physician, Dr. Mendenhall, described the case of
three young siblings who initially presented with skin and teeth abnormalities. The children, two
girls and one boy, had a distinct appearance with coarse skin, acanthosis nigricans, and a senile
facies. It was later discovered that the pineal gland, a gland at the base of the brain that secretes
melatonin, was increased in size (pineal hyperplasia).
The constellation of pineal hyperplasia, insulin resistance, and other somatic abnormalities is
called Rabson-Mendenhall syndrome. This rare syndrome is caused by a mutation of the insulin
receptor gene, which leads to severe insulin resistance.
Rabson-Mendenhall syndrome in OMIM

Additional symptoms that appear from the first year of life include abdominal swelling and
abnormal enlargement of the clitoris in females and penis in boys. Deficiency or absence of adipose tissue may also be present.

Lipoatrophic Diabetes
Lipoatrophy is the wasting away (atrophy) of fat tissue. In the syndrome Berardinelli-Seip, lipoatrophy is so severe that from birth adipose tissue is almost absent. From early infancy severe
insulin resistance causes diabetes. Other features include acanthosis nigricans, increased production of androgen hormones, an enlarged liver, and increased muscle mass.
Berardinelli-Seip Congenital Lipodystrophy syndrome in OMIM

At least two mutations on different chromosomes have been identified as a cause of Berardinelli syndrome, mutations in AGPAT2 on chromosome 9 (7) and mutations in BSCL2 mutation
on chromosome 11 (8, 9). In many cases, disruption of the structure and the function of the
insulin receptor cannot be found. For this reason, it is assumed that the problem lies at the postreceptor level, involving signal transduction.
Table 1. Syndromes of severe insulin resistance
Syndrome
Disruption of insulin receptor
Type A insulin resistance
Leprechaunism
Rabson-Mendenhall syndrome
Lipoatrophic syndromes
e.g., Berardinelli-Seip (congenital generalized
lipodystrophy)
Acquired syndromes
Type B insulin resistance

Cause

Mode of inheritance

Mutation of insulin receptor in up to 10%


of cases
Mutation of insulin receptor
Mutation of insulin receptor

Usually dominant
Recessive
Recessive

Involves at least two loci

Recessive

Anti-insulin receptor antibodies

N/A

4-31

Diabetes

Other Types of Diabetes

References
1. Kahn CR, Flier JS, Bar RS et al. The syndromes of insulin resistance and acanthosis nigricans. Insulinreceptor disorders in man. N Engl J Med 294:739745; 1976. (PubMed)
2. Moller DE, Cohen O, Yamaguchi Y et al. Prevalence of mutations in the insulin receptor gene in subjects
with features of the type A syndrome of insulin resistance. Diabetes 43:247255; 1994. (PubMed)
3. Donohue WL, Uchida I. Leprechaunism: a euphemism for a rare familial disorder. J Pediatr 45:505519;
1954. (PubMed)
4. Kadowaki T, Bevins CL, Cama A et al. Two mutant alleles of the insulin receptor gene in a patient with
extreme insulin resistance. Science 240:787790; 1988. (PubMed)
5. Kadowaki T, Kadowaki H, Accili D et al. Substitution of arginine for histidine at position 209 in the alphasubunit of the human insulin receptor. A mutation that impairs receptor dimerization and transport of
receptors to the cell surface. J Biol Chem 266:2122421231; 1991. (PubMed)
6. van der Vorm ER, Kuipers A, Kielkopf-Renner S et al. A mutation in the insulin receptor that impairs
proreceptor processing but not insulin binding. J Biol Chem 269:1429714302; 1994. (PubMed)
7. Garg A, Wilson R, Barnes R et al. A gene for congenital generalized lipodystrophy maps to human
chromosome 9q34. J Clin Endocrinol Metab 84:33903394; 1999. (PubMed)
8. Magre J, Delepine M, Khallouf E et al. Identification of the gene altered in Berardinelli-Seip congenital
lipodystrophy on chromosome 11q13. Nat Genet 28:365370; 2001. (PubMed)
9. Agarwal AK, Simha V, Oral EA et al. Phenotypic and genetic heterogeneity in congenital generalized
lipodystrophy. J Clin Endocrinol Metab 88:48404847; 2003. (PubMed)

4-32

Diabetes

Other Types of Diabetes

Diseases in the Exocrine Pancreas


The pancreas gland lies across the posterior abdominal wall, behind the stomach. It has two main
portions, the endocrine portion that secretes hormones and the exocrine portion that secretes
enzymes.
The exocrine portion of the pancreas makes up more than 95% of its total cell mass. Here,
powerful digestive enzymes are produced and are delivered to the duodenum (a part of the small
intestine) via the pancreatic duct. These enzymes break down carbohydrates, proteins, and fats
into smaller molecules that can be absorbed across the gut wall.
Any processes that diffusely injure the pancreas can result in diabetes and include:

pancreatitis
trauma or surgical removal of the pancreas
cancer of the pancreas
cystic fibrosis
fibrocalculous pancreatopathy
the iron storage disease hemochromatosis

4-33

Diabetes

Other Types of Diabetes

Diseases of the Endocrine System


Most cases of diabetes are caused by a combination of loss of beta cell function and insulin resistance. However, diabetes may also be caused by endocrine disorders that produce excess hormones that antagonize the action and secretion of insulin e.g., cortisol, growth hormone, and
glucagon.
Patients with these endocrine diseases frequently develop diabetes secondary to a hormoneinduced hyperglycemia that causes either insulin loss or an increase in insulin resistance. Treatment of the underlying disorder often leads to the normalization of blood glucose levels.

Cushing's Syndrome
Cortisol increases blood sugar by increasing the liver's production of glucose while at the same
time increasing insulin resistance in peripheral tissues. Cushing's syndrome is caused by an
excess of cortisol, and hyperglycemia or diabetes is commonly observed in affected individuals.

Acromegaly
Growth hormone is synthesized in the pituitary gland in the brain. A tumor in this gland is the
main cause of acromegaly, an endocrine disorder characterized by excess growth hormone.
Growth hormone increases insulin resistance, resulting in over half the patients showing signs of
glucose intolerance and hyperinsulinemia.

Pheochromocytoma
Epinephrine, as part of the "fight or flight" response, mobilizes glucose to ensure a readily accessible source of fuel in an emergency. By acting on alpha adrenoreceptors, epinephrine inhibits
insulin secretion, increases the breakdown of glycogen to glucose in the liver and muscle, and
also stimulates the breakdown of fat. By acting on beta adrenoreceptors, epinephrine increases
peripheral insulin resistance.
Excess epinephrine can be the result of a pheochromocytoma, a tumor originating from the
adrenal gland. The tumor increases epinephrine synthesis, and the resulting increased circulating
levels of epinephrine can lead to diabetes.

Glucagonoma
Glucagon opposes many of insulin's actions. Tumors of the pancreatic alpha cells are rare, but
they may cause an increase in glucagon levels, resulting in impaired glucose regulation.

Somatostatinoma
Somatostatin is secreted by a range of tissues, including the delta cells of the pancreas, and
somatostatin inhibits the secretion of growth hormone. Diabetes is associated with somatostatinoma, a rare endocrine pancreatic tumor that secretes excess somatostatin, which inhibits insulin
secretion.

4-34

Diabetes

Other Types of Diabetes

Drug- or Chemical-induced Diabetes


Many medications can impair insulin secretion. Such drugs may not directly cause diabetes but
rather precipitate diabetes in individuals with pre-exisitng insulin resistance and deficiency. In
addition, certain hormones, when in excess or when given as a therapy, can impair the action of
insulin, e.g., glucocorticoids, thyroid hormone.
Although rare, particular toxins such as rat poison and specific drugs can permanently
destroy the beta cells of the pancreas. This results in the abrupt onset of diabetes that requires
insulin treatment, e.g., Vacor, Pentamidine.

Beta-Adrenergic Agonists
Beta-Adrenergic agonists such as salbutamol are most commonly used in the treatment of
asthma. One of the possible side effects of beta agonists is hyperglycemia, which is caused by a
decrease in insulin sensitivity.

Diazoxide
Diazoxide is used to treat hypoglycemia, and it works by preventing the pancreas from releasing
insulin. Diazoxide is a potassium channel opener; it activates the pancreatic beta cell ATPsensitive K+ (KATP) channel, hyperpolarizes the beta cell, and prevents insulin release. It is used
in the treatment of insulinomas (insulin-secreting tumors), persistant hyperinsulimic hypoglycemia
of infancy, and because diazoxide also dilates blood vessels, it can be used to lower high blood
pressure.

Glucocorticoids
Glucocorticoid drugs are synthetic copies of the body's steroid hormones. Steroids raise blood
glucose levels by counteracting many of the actions of insulin, favoring the break down of carbohydrates, fat, and even protein, releasing raw materials from which glucose can be made. Excess
steroid hormone (e.g., in Cushing's disease) or prolonged use of steroid drugs (e.g., prednisolone) can lead to glucose intolerance or diabetes.

Interferon-alpha Therapy
The body makes interferon alpha (IFN) as part of the immune response. It is produced in particular types of white blood cells in response to infection or cancer. IFN can be given as a treatment
for certain types of cancer and long-standing infections and inflammatory conditions.
When IFN therapy is used to treat chronic hepatitis C, a rare side effect is that some
patients develop diabetes. IFN appears to trigger an autoimmune attack against several
endocrine organs, including the pancreas islet cells. Similar to type 1 diabetes, this new-onset
diabetes requires treatment with insulin.

4-35

Diabetes

Other Types of Diabetes

Nicotinic Acid
Nicotinic acid is a B vitamin that is found in meat, poultry, fish, wholemeal cereals, pulses, and
coffee. It is also taken as a drug to lower lipid levels (serum cholesterol and triglycerides). There
are several side effects of taking nicotinic acid, including liver toxicity and deranged blood glucose
levels.

Pentamidine
Pentamidine is an antiprotozoal agent used to treat trypanosomiasis, leishmaniasis, and some
fungal infections. A more common use of this drug in the United States is in the treatment of
pneumocystis pneumoniae, which can cause pneumonia in immunocomprised patients. Pentamidine can cause irreversible beta cell damage, leading to loss of insulin and resulting in diabetes. It
is also toxic to the central nervous system.

Phenytoin
Phenytoin is an anticonvulsant drug that is effective in controlling a wide variety of seizure disorders. It is thought to suppress seizures by blocking sodium ion channels in neurons, preventing
overexcitation. One side effect of phenytoin use is hyperglycemia. This may be because phenytoin blocks calcium ion channels in the pancreatic beta cells, inhibiting insulin release.

Thiazides
High doses of thiazides (a type of diuretic) can worsen hyperglycemia in type 2 diabetes. Thiazides appear to impair insulin secretion as a consequence of causing K+ depletion (a known
side effect of thiazides). Thiazides may also increase insulin resistance. The effect on glucose
intolerance is less when the dose of thiazide is decreased.

Vacor
Accidental ingestion of the rat poison Vacor can be fatal. Vacor is toxic to pancreatic beta cells,
rapidly depleting insulin production and causing acute diabetic ketoacidosis.

Link Roundup
MedlinePlus Drug Information
Pentamidine [http://www.nlm.nih.gov/medlineplus/druginfo/uspdi/202449.html] | Nicotinic acid [http://www.nlm.nih.gov/medlineplus/
druginfo/uspdi/202405.html] | Glucocorticoids [http://www.nlm.nih.gov/medlineplus/druginfo/uspdi/202018.html] | Diazoxide [http://
www.nlm.nih.gov/medlineplus/druginfo/uspdi/202191.html] | Beta adrenergic agonists: inhaled [http://www.nlm.nih.gov/medlineplus/
druginfo/uspdi/202095.html], oral [http://www.nlm.nih.gov/medlineplus/druginfo/uspdi/202096.html] | Thiazides [http://www.nlm.nih.
gov/medlineplus/druginfo/uspdi/202208.html] | Phenytoin [http://www.nlm.nih.gov/medlineplus/druginfo/medmaster/a682022.html] |
Interferon-alpha therapy [http://www.nlm.nih.gov/medlineplus/druginfo/uspdi/202299.html]

4-36

Diabetes

Other Types of Diabetes

Infections
A genetic predisposition to type 1 diabetes has been well established. However, many lines of
evidence also point to the existence of environmental risk factors that may act as the trigger for
the autoimmune attack on the pancreas (1).
Viruses have been suspected to contribute to the onset of type 1 diabetes because new
cases of diabetes occur more frequently at certain times of the year (2). More recently, virusspecific IgM antibodies have been isolated from patients with new-onset diabetes (3), and pancreatic extracts from patients who died from new-onset diabetes cause diabetes in animals by the
destruction of beta cells (4).
Several viruses have been associated with inducing certain cases of diabetes and include the
following:

rubella virus
Coxsackie B virus
mumps virus
cytomegalovirus
Epstein-Barr virus
adenovirus
rotavirus
References
1. Yoon JW, Jun HS, et al. Role of viruses in the pathogenesis of type 1 diabetes mellitus. In: Diabetes
Mellitus: a fundamental, clinical text. Philadelphia: Lippincott-Raven; 2000. p. 419430.
2. Gamble DR, Taylor KW. Seasonal incidence of diabetes mellitus. Br Med 13:631633; 1969. (PubMed)
3. Banatvala JE, Bryant J, Schernthaner G et al. Coxsackie B, mumps, rubella and cytomegalovirus specific
IgM responses in patients with juvenile-onset insulin-dependent diabetes mellitus in Britain, Austria and
Australia. Lancet 1:14091412; 1985. (PubMed)
4. Yoon JW, Austin M, Onodera T et al. Isolation of a virus from the pancreas of a child with diabetic
ketoacidosis. N Engl J Med 300:11731179; 1979. (PubMed)

4-37

Diabetes

Other Types of Diabetes

Uncommon Forms of Immune-mediated Diabetes


Type 1 diabetes is caused by an autoimmune attack of the pancreatic islets. Immune-mediated
attacks are also responsible for rarer syndromes of which diabetes is a feature, including those
discussed below.

Antibodies to Insulin
Rarely, the formation of insulin autoantibodies can deplete levels of insulin to such an extent that
diabetes develops.

Antibodies to the Insulin Receptor


Autoantibodies directed against the insulin receptor are occasionally found in patients who have
co-existing autoimmune diseases, such as systemic lupus erythematosus (SLE). As in other
states of severe insulin resistance, the skin disorder acanthosis nigricans is often found. The syndrome of severe insulin resistance with circulating antibodies to the insulin receptor is known as
type B insulin resistance.
Anti-insulin receptor antibodies can cause hyperglycemia by binding to the insulin receptor
and blocking the binding of insulin to its receptor in target tissues. In rare cases, these autoantibodies can have the opposite effect, causing hypoglycemia by mimicking the action of insulin.

"Stiff Man" Syndrome


The Stiff Man syndrome is a rare autoimmune disorder of the central nervous system that is
characterized by stiffness of the axial muscles. Individuals have painful muscle spasms that may
be precipitated by unexpected events or physical contact. As the disease progresses, there is
increasing stiffness of the muscles supporting the spine and in the arms and legs.
Stiff Man Syndrome in OMIM

The autoantibody anti-glutamic acid decarboxylase (GAD) is found in high levels in classical
Stiff Man syndrome (1). The antibody is directed against an enzyme found in the nerve tissue and
may play a role in the abnormal muscle activity of these patients (electrical studies show that the
muscle is unable to relax). In addition, anti-GAD antibodies may attack the pancreas, which also
contains the enzyme. This may be the cause of one in three patients with Stiff Man syndrome
developing a form of insulin-dependent diabetes.

References
1. Solimena M, Folli F, Aparisi R et al. Autoantibodies to GABA-ergic neurons and pancreatic beta cells in
stiff-man syndrome. N Engl J Med 322:15551560; 1990. (PubMed)

4-38

Diabetes

Other Types of Diabetes

Other Genetic Syndromes Sometimes Associated with Diabetes


Many genetic syndromes are accompanied by an increased incidence of diabetes mellitus and
include syndromes caused by a single gene mutation and syndromes caused by a chromosomal
abnormality.

Gene Mutations
Friedreich ataxia [Genes and Disease [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.
View..ShowSection&rid=gnd.section.205]]

Huntington's chorea [Genes and Disease [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.


View..ShowSection&rid=gnd.section.207]]

Lawrence-Moon-Biedel syndrome
Myotonic dystrophy [Genes and Disease [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.
View..ShowSection&rid=gnd.section.164]]

Porphyria [Genes and Disease [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.View..


ShowSection&rid=gnd.section.267]]

Prader-Willi syndrome [Genes and Disease [http://www.ncbi.nlm.nih.gov/books/bv.fcgi?call=bv.


View..ShowSection&rid=gnd.section.182]]

Wolfram's syndrome

Chromosomal Abnormalities
Down's syndrome
Klinefelter's syndrome
Turner's syndrome

4-39

Diabetes

Gestational Diabetes

5. Gestational Diabetes
Created: July 7, 2004

During pregnancy there are many changes that take place in the mother's metabolisma rise in
insulin resistance is one of these changes.
The placenta supplies a growing fetus with nutrients and produces a variety of hormones to
maintain the pregnancy. Some of these hormones, such as human placental lactogen, have a
blocking effect on insulin that usually begins 20 to 24 weeks into the pregnancy. The contra-insulin
effect of placental hormones leads to higher levels of maternal blood glucose after eating (postprandial levels) that may aid fetal growth.
Normally, the mother's beta cells can produce additional insulin to overcome the insulin
resistance of pregnancy. As the placenta grows, more hormones are produced, and insulin
resistance becomes greater. When the mother's production of insulin is not enough to overcome the
effect of the placental hormones, gestational diabetes mellitus (GDM) results. GDM is defined as
"carbohydrate intolerance of varying degrees of severity with onset or first recognition during
pregnancy" (1). GDM complicates 7% of all pregnancies in the United States (2) and is more
common in populations with a higher rate of type 2 diabetes mellitus, such as African Americans,
Asian Americans, Hispanic Americans, and Native Americans (3, 4).
The main complications of GDM are increased fetal size, which may complicate delivery, and
hypoglycemia in the baby immediately after delivery. Women with GDM generally have normal
blood sugar levels during the critical first trimester (before the 13th week) of pregnancy. This is in
contrast to patients with type 1 diabetes, where hyperglycemia in this period may cause congenital
birth defects.
After a positive screening test, the diagnosis of GDM is made by a glucose tolerance test. In this
test, a sugary drink is given, and a series of blood tests are taken at set time intervals (Table 1). If
hyperglycemia is detected, treatment begins with a change in diet and an increase in exercise. If
these lifestyle changes fail to control blood glucose levels, insulin therapy is started.
Women with pre-existing diabetes require higher doses of insulin during pregnancy because of
the increase in insulin resistance. If their diabetes is usually controlled using oral hypoglycemic
agents, they are usually transferred to insulin to enable better glucose control and because the
safety of most hypoglycemic agents has not been studied in pregnancy.
GDM can disappear within hours of giving birth, depending on individual factors such as beta
cell function and predisposing factors such as obesity. However, a significant portion of women go
on to develop type 2 diabetes. Because GDM and type 2 diabetes both feature insulin resistance
and share risk factors such as obesity, it is possible that these two conditions may also share
diabetes susceptibility genes.

5-1

Diabetes

Gestational Diabetes

References
1. Metzger BE, Coustan DR et al. Summary and recommendations of the Fourth International WorkshopConference on gestational diabetes mellitus. Diabetes Care 21 (suppl2):B161B167; 1998. (PubMed)
2. Gabbe SG, Graves CR et al. Management of diabetes mellitus complicating pregnancy. Obstet Gynecol
102:857868; 2003. (PubMed)
3. Engelgau MM, Herman WH, Smith PJ et al. The epidemiology of diabetes and pregnancy in the U.S.,
1988. Diabetes Care 18:10291033; 1995. (PubMed)
4. Gestational diabetes mellitus. Diabetes Care 26 Suppl 1:S103S105; 2003. (PubMed)
Table 1. Diagnosis of gestational diabetes.

Glucose load, 100 g


Fasting
1 hour
2 hours
3 hours
Glucose load, 75 g
Fasting
1 hour
2 hours

mg/dl

mmol/l

95
180
155
140

5.3
10.0
8.6
7.8

95
180
155

5.3
10.0
8.6

Gestational diabetes can be diagnosed using either a 100-g or 75-g oral glucose load. Two or more of the venous plasma glucose
concentrations must be met or exceeded for a positive diagnosis. The test should be done in the morning after an overnight fast.

5-2

You might also like