Genetic Basis of DM

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

NIH Public Access

Author Manuscript
Cellscience. Author manuscript; available in PMC 2006 August 3.
Published in final edited form as:
NIH-PA Author Manuscript

Cellscience. 2006 April 30; 2(4): 100–131.

The Genetic Basis of Type 2 Diabetes

Swapan Kumar Das and Steven C Elbein


University of Arkansas for Medical Sciences and Central Arkansas Veterans Healthcare System,
Little Rock, Arkansas

Abstract
Type 2 Diabetes results from a complex physiologic process that includes the pancreatic beta cells,
peripheral glucose uptake in muscle, the secretion of multiple cytokines and hormone-like molecules
from adipocytes, hepatic glucose production, and likely the central nervous system. Consistent with
the complex web of physiologic defects, the emerging picture of the genetics will involve a large
number of risk susceptibility genes, each individually with relatively small effect (odds ratios below
1.2 in most cases). The challenge for the future will include cataloging and confirming the genetic
risk factors, and understanding how these risk factors interact with each other and with the known
environmental and lifestyle risk factors that increase the propensity to type 2 diabetes.
NIH-PA Author Manuscript

Type 2 Diabetes: A Prologue


Type 2 Diabetes Mellitus (T2DM) is a complex heterogeneous group of metabolic condition
characterized by elevated levels of serum glucose, caused mainly by impairment in both insulin
action and insulin secretion. T2DM exerts a huge toll in human suffering and economy. A
recent evaluation using a computerized generic formal disease model (DisMod II) revealed
that excess global mortality due to diabetes in the year 2000 was equivalent to 5.2% of all
deaths and diabetes is likely to be the fifth leading cause of death, similar in magnitude to
numbers reported for HIV/AIDS (1). The prevalence of diabetes for all age-groups worldwide
was estimated to be 2.8% in 2000 and is projected to be 4.4% in 2030 (most of which will be
T2DM). The total number of people with diabetes is projected to rise from 171 million in 2000
to 366 million in 2030, with India, China and USA being the top 3 countries estimated to have
the highest numbers of people with diabetes (2). T2DM accounts for 5–10% of the total health
care budget in many countries. In the United States, the American Diabetes Association
estimated the total cost of diabetes management in 2002 at $132 billion, which is about 1.3%
of US gross domestic product (3,4).
NIH-PA Author Manuscript

The recent global epidemic of T2DM almost certainly indicates the importance of
environmental triggers such as sedentary lifestyle and dietary changes over last several
decades. Nonetheless, T2DM is among many complex diseases for which a genetic
contribution is well accepted. Despite the diverse phenotypic nature of T2DM, twin and family
studies, and the wide spectrum of diabetes prevalence across populations provide convincing
evidence for an important role of

Identification of the genetic etiology of T2DM has proved challenging. Although many
laboratories have chosen a genetic approach to define its pathophysiology, the molecular
mechanisms for this strong genetic predisposition remain elusive. Based on epidemiological
data, the total sibling relative risk (λs) for T2DM has been estimated at 3–4, although the
number of loci that contribute to this risk is uncertain. Multiple genome scans and a plethora
of candidate gene association studies of T2DM in conglomerate lead to the inescapable
conclusion that early proposals of an oligogenic model of T2DM were overly optimistic. The
increasingly complex physiologic picture reflects this multifaceted genetic landscape. Early
views of T2DM as primarily a defect in insulin mediated muscle glucose uptake have been
tempered by a new appreciation of the key roles of the pancreatic β-cell, the adipocyte as a
Das and Elbein Page 2

secretory organ, a primary role of the liver, and possibly a key role of the central nervous
system. Identification of the genetic components of type 2 diabetes is one of the most important
areas of diabetes research because elucidation of the diabetes genes (alleles) will influence all
NIH-PA Author Manuscript

efforts toward a mechanistic understanding of the disease, its complications, and its treatment,
cure, and prevention. In this review we will discuss our current understanding of the genetic
basis of the pathophysiology of Type 2 diabetes.

T2DM Is a Genetic Disease: Classical Evidence


Multiple lines of evidence support the view that genetic components plays an important role
in the pathogenesis of T2DM.
1. The spectrum of T2DM prevalence in different ethnic groups’ The prevalence of
T2DM varies widely among populations, from 1% in Chile Mapuche Indian, 2%
among Caucasians in Europe to as high as 41% in the Nauru (Pacific Island) and 50%
among Pima Indians in Arizona (5). Part of this observed ethnic variability can be
attributed to non-genetic environmental and cultural factors; however, the observation
that the disease prevalence varies substantially among ethnic groups that share a
similar environment supports the idea that genetic factors contribute to disease
predisposition.
2. Familial aggregation: Other than genes, families share environments, culture and
NIH-PA Author Manuscript

habits, yet familial aggregation of the disease is another source of evidence for a
genetic contribution to the disease. Evidence for a genetic role includes the nearly 4-
fold increased risk for T2DM in siblings of a diabetic proband compared with the
general population (λS of 3.5 to 4), the odds ratio (OR) of 3.4–3.5 with only a single
affected parent, and the increase in the OR to 6.1 if both parents are affected (6).
3. Twin studies: Multiple studies of twin concordance rates have been undertaken in
T2DM. Estimates for concordance rates have ranged from 0.29 to 1.00 in monozygotic
(MZ) twins, while in dizygotic (DZ) twins the range was 0.10–0.43 (7–10).
Concordance among both MZ and DZ twins increases with the duration of follow up
period (10). In spite of several caveats in twin studies, the high concordance in MZ
twins and the 50% fall in DZ twins provides compelling evidence for a genetic
component of T2DM.
4. Heritability of intermediate phenotypes: Insulin sensitivity and insulin secretion
deteriorate in parallel in most human T2DM. Both defects predicted subsequent
T2DM in several studies and both defects are shown to be present in nondiabetic but
genetically identical co-twins of a diabetic proband (11). Data from multiple
laboratories, including ours, support a genetic basis for measures of both insulin
NIH-PA Author Manuscript

sensitivity and insulin secretion (12–14).

Pathophysiology of T2DM: Current status


The early pathophysiology of T2DM remains uncertain and controversial. Three key defects
mark the onset of hyperglycemia in T2DM: increased hepatic glucose production, diminished
insulin secretion, and impaired insulin action (15,16). Unfortunately, at the time of
hyperglycemia, glucose and possibly lipid toxicity obscure the primary effects. Prospective
and cross-sectional analyses of euglycemic individuals at risk (relatives of T2DM individuals)
circumvent this dilemma, and suggest a key early and predictive role of reduced insulin
sensitivity in T2DM pathogenesis. Nonetheless, few individuals with genetic insulin resistance
develop diabetes, and 25% of nondiabetic individuals may have insulin sensitivity as low as
that seen in T2DM. The increase in hepatic glucose production was formerly viewed as a late
event, occurring only at the onset of hyperglycemia and glucose intolerance. However, knock
out murine models suggest a more primary role for hepatic insulin resistance.

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 3

The importance and timing of a β-cell defect have been hotly debated (17). Obese individuals,
those with a family history of diabetes, and individuals with impaired glucose tolerance (IGT)
are all characterized by absolute hyperinsulinemia until the development of overt
NIH-PA Author Manuscript

hyperglycemia. However, when β-cell function is viewed in the context of reduced insulin
sensitivity, considerable data support the early failure of insulin secretion in T2DM
pathogenesis (18,19). Animal models also support this concept. Despite much study, the signals
that cause normal β-cell compensation and hyperinsulinemia, the mechanisms of this
compensation, the point in the pathogenesis of T2DM where this compensatory mechanism
fails, and the etiology of this failure all remain subjects of speculation. Insulin sensitivity and
insulin secretion deteriorate in parallel in most human T2DM. The remarkable difficulty in
uncoupling these two defects may suggest a common mechanism.

Extensive investigation of humans and animal models has failed to clarify the key or early
events that lead to the inexorable rise in plasma glucose and eventual insulin deficiency. Current
models, which propose a role for multiple pathways affecting multiple organs, suggest a
complex genetic substructure. Most glucose after a meal is taken up by skeletal muscle, with
adipose playing a much smaller role. Traditionally, resistance to insulin mediated, non-
oxidative glucose metabolism was viewed as the early and primary defect leading to T2DM.
More recently, the progressive failure of the β-cell to compensate for insulin resistance has
received considerable attention (19), but the role of liver and adipose had been considered late
or relatively insignificant factors, respectively. Animal models have challenged even this
NIH-PA Author Manuscript

paradigm. Muscle-specific insulin receptor (INSR) inactivation did not lead to IGT despite
insulin resistance, whereas muscle specific inactivation of glucose transporter GLUT4 caused
both profound insulin resistance and T2DM (20). Unexpectedly, GLUT4 inactivation limited
to white adipose resulted in insulin resistance, glucose intolerance, and even T2DM with an
impairment in whole body glucose uptake markedly out of proportion to the expected
contribution of white adipose (20). This isolated defect in white adipose resulted in defective
insulin action in both muscle and liver. Paradoxically, inactivation of the adipocyte insulin
receptor gene resulted in improved insulin sensitivity and protection against obesity. These
studies suggest a central and previously under-appreciated role for adipocytes in the
pathogenesis of T2DM. However, insulin receptor inactivation in the liver also resulted in
hyperinsulinemia, hepatic insulin resistance, and peripheral insulin resistance (21–24). These
studies suggest considerable crosstalk among organ systems, as well as an important role for
insulin signaling pathways in the β-cell (25). Thus, T2DM is the final outcome of a multiorgan
disease which is characterized by early direct or indirect defects in muscle, adipocytes,
hepatocytes, β-cells, and possibly the central nervous system (CNS).

Cytokines, Adipokines and Inflammation


Insulin stimulated glucose uptake into muscle cells and adipocytes depends mostly on the
NIH-PA Author Manuscript

glucose transporter GLUT4, skeletal muscle being the principal site for this glucose uptake.
Downregulation of GLUT4 expression in adipose tissue is a common feature of insulin-
resistant states, including obesity, T2DM, and the metabolic syndrome (26). Given that skeletal
muscle is the major site for glucose disposal, the hyperglycemia associated with obesity and
T2DM cannot be explained by the decreased uptake of glucose into adipose tissue due to
downregulation of GLUT4 in adipocytes. Furthermore, adipocyte-selective knockout of
GLUT4 (adipose-Glut4 −/−) in mice resulted in systemic insulin resistance similar to that
induced by muscle-selective GLUT4 knockout mice (muscle-Glut4 −/−, 27,28). These studies
suggest that adipocyte GLUT4 deficiency results in circulating factors that are responsible for
cross-organ communication (“crosstalk”). Yang et. al (29) recently identified retinol binding
protein-4 (RBP4) as one candidate for that crosstalk in the adipose tissue of adipose-specific
Glut4 knockout mice. Furthermore, serum RBP4 protein levels were elevated in insulin-
resistant mice and obese and diabetic humans. Transgenic overexpression of the human

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 4

RBP4 gene or injection of recombinant RBP4 protein in normal mice caused insulin resistance,
whereas Rbp4 knockout mice had enhanced insulin sensitivity (29). Increased serum RBP4
protein levels induced hepatic expression of the gluconeogenic enzyme phosphoenolpyruvate
NIH-PA Author Manuscript

carboxykinase (PEPCK) and impaired insulin signaling in muscle.

Investigators have proposed other potential mediators of inter-organ crosstalk to explain T2DM
pathogenesis. Obesity predisposes to insulin resistance and T2DM. Adipocytes, both visceral
and peripheral, secrete a plethora of factors which may alter systemic insulin action and hepatic
glucose production, including adiponectin, resistin, leptin, cytokines IL-6 and TNFα, visfatin,
retinol binding protein 4 (RBP4), as well as free fatty acids (FFA) (30–32). In recent years, the
concept that obesity and elevated cytokine secretion results in an inflammatory state that is
itself the cause of insulin resistance has emerged (32–34). The inhibition of signaling
downstream of the insulin receptor may be a primary mechanism through which this
inflammatory state causes insulin resistance. Exposure of cells to TNF-α or elevated levels of
free fatty acids stimulates phosphorylation of serine residues of IRS-1. This phosphorylation
reduces tyrosine phosphorylation of IRS-1 in response to insulin, and thus signaling
downstream of IRS-1 (35–38). The inflammatory state is increased by the accumulation of
macrophages in adipose tissue with obesity (39;40). Thus, both adipose tissue and the
macrophages within adipose tissue serve in both endocrine and paracrine fashion to promote
inflammation and decrease insulin sensitivity (32,33).
NIH-PA Author Manuscript

Endoplasmic Reticulum Stress Response and Diabetes


Metabolic and inflammatory stresses of obesity may also disrupt cell metabolism, particularly
the processing of cellular proteins in the secretory cells of the liver, pancreas, and adipocyte.
The response to this increased demand is the endoplasmic reticulum stress response pathway,
which activates an extensive cascade which shuts down further translation, increases the
transcription of chaperone proteins, and if not successful, causes cellular apoptosis (41–43).
Among the activated proteins are IRE1 and XBP1. Özcan and colleagues (42) showed recently
that the endoplasmic reticulum stress pathway is active in adipose tissue and in the liver from
obese mice, leading to increased activation of c-jun N terminal kinase (JNK), which in turn
phosphorylates IRS-1 on serine residues, thus suppressing insulin action and downstream
signaling pathways. Similar data also support a role for endoplasmic reticulum stress pathways
in β-cell dysfunction and apoptosis (44).

Endoplasmic reticulum stress is a molecular link between obesity, decreased insulin sensitivity
and type 2 diabetes. Additionally, increased glucose metabolism can lead to a rise in
mitochondrial production of reactive oxygen species (ROS). Inflammatory pathways can be
initiated by extracellular mediators such as cytokines and lipids or by intracellular stresses such
as endoplasmic reticulum stress or excess ROS production by mitochondria. Signals from all
NIH-PA Author Manuscript

of these mediators converge on inflammatory signaling pathways, including the kinases JNK
and IKK. These pathways lead to the production of additional inflammatory mediators through
transcriptional regulation as well as to the direct inhibition of insulin signaling (45,46).

Mitochondria and Reactive Oxygen Species


Reactive Oxygen Species (ROS) are another factor suggested to have a role in insulin
resistance. ROS production is elevated in obesity, which in turn causes enhanced activation of
inflammatory pathways (47,48). Several lines of evidence support the association of oxidative
stress markers with diabetes (47–49). Houstis et al (50) recently connected these mechanisms
by showing that treatment of 3T3-L1 adipocyte with either TNFα (an inflammatory cytokine
that is elevated in obesity) or dexamethasone (an anti-inflammatory agent that causes insulin
resistance) increased the ROS level and resulted in decreased insulin action. Antioxidant
molecules or transgenes encoding ROS scavenging enzymes both ameliorated the insulin

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 5

resistance of TNFα or dexamethasone treated 3T3-L1 adipocytes to varying degrees.


Furthermore, antioxidant molecules improved insulin sensitivity and glucose homeostasis in
obese, insulin resistant ob/ob mice (50). Oxidative stress is also likely to be involved in
NIH-PA Author Manuscript

progression of pancreatic β-cell dysfunction in T2DM.

Reduced mitochondrial oxidative phosphorylation has recently been demonstrated in several


studies, using both microarrays to identify downregulated genes involved in oxidative
phosphorylation (OXPHOS genes). PGC1 appears to be at least partially responsible for this
defect (51,52). Supportive studies in humans have suggested a role for defective mitochondrial
fatty acid oxidation, mitochondrial dysfunction, and reduced numbers of skeletal muscle
mitochondria in T2DM pathogenesis, and have suggested that increased intramyocellular lipid
content was associated with defects in mitochondrial activity (53–56). Morino et al (56) found
38% less mitochondrial density in muscle of insulin resistant (IR) individuals compared to
controls, which was associated with a 50% increase in serine phosphorylation of IRS-1 and a
60% reduction in insulin stimulated Akt activation (a measure of insulin signaling). Decreased
mitochondrial fatty acid oxidation, caused by either mitochondrial dysfunction and/or reduced
mitochondrial numbers, produces increased levels of intracellular fatty acyl CoA and
diacylglycerol. These molecules activate a novel protein kinase C, which in turn activates a
serine kinase cascade, again leading to increased serine phosphorylation of IRS-1. As described
above, serine phosphorylation blocks IRS-1 tyrosine phosphorylation and inhibits downstream
signaling, including recruitment of GLUT4 to the plasma membrane and insulin-mediated
NIH-PA Author Manuscript

glucose uptake in skeletal muscle (53).

Oxidative Metabolism and the Pancreatic β-Cell


Insulin secretion by the pancreatic β-cell is modulated by multiple stimuli. Oxidative
mitochondrial metabolism and adenosine triphosphate (ATP) generation is essential to glucose
stimulated insulin secretion. The increased ratio of ATP to adenosine diphosphate (ADP) in
the β-cell triggers a series of events: inhibition of the cell’s ATP/ADP-regulated potassium
channel (KATP, encoded by genes KCNJ11 and ABCC8), plasma membrane depolarization,
opening of a voltage-gated calcium channel, calcium influx, and transport and binding of
insulin granules to the cell surface (53). The ATP/ADP ratio is in turn altered by UCP2, an
integral mitochondrial membrane protein that permits protons to leak across the mitochondrial
inner membrane, thus uncoupling of glucose oxidative metabolism from ATP production
(57). By decreasing the amount of ATP generated from glucose, UCP2 expression negatively
regulates glucose-stimulated insulin secretion. Over-expression of UCP2 in β-cells in vitro
decreases glucose-stimulated insulin secretion (58), whereas targeted inactivation of the UCP2
gene in mice has the opposite effect (59). Importantly, heterozygosity for a null UCP2 allele
produces an effect that is intermediate between those observed in wild-type and homozygous
mice, indicating that even relatively small changes in UCP2 expression may have a substantial
NIH-PA Author Manuscript

impact on glucose-stimulated insulin secretion.

Glucose Homeostasis: The Central Nervous System


Levels of glucose in the blood are regulated by a complex interplay between the appearance
of glucose from both intestinal absorption and hepatic production and its disappearance through
insulin-dependent and insulin independent glucose uptake in a variety of tissues. After the
overnight fast, glucose is largely produced by glycogen breakdown and gluconeogenesis.
About 80% of this glucose released by liver is metabolized independent of insulin by brain and
other insulin–independent tissues (gut, red cells). The medial hypothalamus, a major integrator
of nutritional and hormonal signals, plays a pivotal role not only in the regulation of energy
balance but also in the modulation of liver glucose output. Diminished responses of first-order
neurons, such as those in the arcuate nucleus, may contribute to the pathogenesis of T2DM
(60–62). The enzyme carnitine palmitoyltransferase-1 (CPT1) regulates long-chain fatty acid

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 6

(LCFA) entry into mitochondria, where they undergo β-oxidation. Obici et al (63) observed
that the selective reduction of lipid oxidation in the hypothalamus by inhibiting or decreasing
the activity of CPT1, substantially diminished food intake and endogenous glucose production.
NIH-PA Author Manuscript

These results indicated that changes in the rate of lipid oxidation in selective hypothalamic
neurons signaled nutrient availability to the hypothalamus, which in turn modulated the
exogenous and endogenous inputs of nutrients into the circulation.

A recent study by Pocai et al (64) revealed an important role for circulating LCFA and insulin
levels, and a sensing role of hypothalamic (arcuate nucleus) KATP channels. Insulin infused into
the region of the arcuate nucleus potently reduced hepatic glucose production from the liver.
This effect was mediated by changing the level of activity of the hepatic branch of the vagus
nerve that provided the major neural input into the liver. KATP channel inhibitors infused into
the arcuate nucleus completely blocked the effects of CNS-infused insulin to reduce hepatic
glucose production. Consistent with these results, mice lacking the SUR1 (ABCC8) subunit
of the KATP channel were resistant to the inhibitory action of insulin on gluconeogenesis. Thus,
activation of hypothalamic KATP channels normally restrains hepatic gluconeogenesis.
Alterations within this central nervous system/hepatic circuit may contribute to diabetic
hyperglycaemia (64). Other peripheral signals to the hypothalamic centers include leptin and
insulin (61).

Glucose Homeostasis: Liver


NIH-PA Author Manuscript

Glycogenolysis and gluconeogenesis are the two known source of hepatic glucose production
(HGP). HGP is controlled by the integrated metabolite fluxes to and from the liver, hormonal
cues, and as reviewed above, neurotransmitter release. Fasting promotes glucose production
through cAMP-dependent mechanisms, whereas feeding inhibits glucose production through
insulin. In the absorptive state, insulin silences hepatic PGC-1α transcription, and suppresses
gluconoegenic enzymes PEPCK and G6Pase through the insulin signaling pathway (IRS2-Akt-
mediated phosphorylation, 65). Glucagon opposes these actions by stimulating the
transcription of gluconeogenic genes via the cyclic AMP-inducible factor CREB (cAMP
response element–binding protein), thus increasing HGP. Recent findings (66–69) have begun
to elucidate the role of additional hormone-regulated transcription factors that control
gluconeogenesis. LKB1, a tumor-suppressor gene which encodes a protein-threonine kinase,
phosphorylates and activates AMPK (adenosine monophosphate activated protein kinase) and
SIK2 (Salt-inducible kinase) in liver (67). The deletion of LKB1 in the liver of adult mice
resulted in a nearly complete loss of AMPK activity and hyperglycemia with increased
gluconeogenic and lipogenic gene expression. LKB1 acts through TORC2 (transducer of
regulated CREB activity 2), a transcriptional coactivator of CREB which in turn acts on
PGC-1α to increase gluconeogenesis. Downregulation of TORC2 reduced PGC-1α expression
and normalized blood glucose levels in mice with liver specific LKB1 inactivation, indicating
NIH-PA Author Manuscript

that TORC2 is a critical target of LKB1/AMPK signaling (67).

The β-Cell and Type 2 Diabetes


Impairment of insulin secretion from pancreatic β-cell is also a major component of T2DM
pathogenesis. Analysis of mutations involved in six different maturity onset diabetes of the
young (MODY) genes have revealed the important role of transcription factors in the insulin
secretion. In a recent gene expression analysis of pancreatic islets isolated from T2DM and
NGT individuals, Gunton et al (70) found a 90% decrease of the transcription factor,
arylhydrocarbon nuclear receptor translocator (ARNT) in T2DM islets compared with
nondiabetic individuals. Mice lacking ARNT in β-cells exhibited abnormal glucose tolerance,
impaired insulin secretion and changes in islet gene expression similar to human T2DM (70).
ARNT and its partner proteins are members of the basic helix-loop-helix Per/AhR/ARNT/Sim
(bHLH-PAS) family of transcription factors. ARNT interacting partners include HIF1α,

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 7

HIF2α and AhR, which together regulate the transcription of genes required for optimal
glucose-responsive insulin secretion by β-cells. ARNT-HIF dimers may directly regulate the
expression of several glycolytic enzymes. Oxygen and NO may also regulate insulin secretion
NIH-PA Author Manuscript

through this ARNT-related pathway. ARNT is an obligate partner of several transcription


factors involved in the response to toxins and hypoxic stress, and may integrate the genetic and
environmental insults leading to T2DM pathogenesis (71).

Many mechanisms contributing to T2DM may trigger β-cell apoptosis and reduced β-cell mass
or ability to compensate for insulin resistance (72). Mechanisms include endoplasmic reticulum
stress (44), chronic hyperglycemia (glucotoxicity, 73,74), chronic hyperlipidemia
(lipotoxicity) (75;76), oxidative stress (77), and inflammatory cytokines (78). Decreased IRS-2
expression, an essential β-cell growth factor, may lead to spontaneous β-cell apoptosis (79).
Because several mechanisms relevant to pathogenesis of T2DM may increase IRS-2 serine/
threonine phosphorylation (80,81) with resultant IRS-2 ubiqutination and proteosomal
degradation, defects in insulin signaling and insulin secretion may be coupled.

Multiple pathways may lead to the T2DM phenotype. This review of the pathophysiology has
covered only a fraction of the genes and pathways that may be implicated in T2DM
pathogenesis, and that may harbor genetic variants. Based on the intertwined and complex
physiologic pathway, a complicated genetic picture can be well anticipated. Indeed, such a
picture is now emerging and is reviewed below.
NIH-PA Author Manuscript

The Quest for T2DM Susceptibility Loci: A Linkage Approach


As we reviewed above, decades of careful research have failed to clarify a unified model of
the pathogenesis of T2DM. hence, using pathophysiology to predict susceptibility loci is
difficult. The complex genetic pattern of T2DM, including the variable age of onset, reduced
penetrance (not everyone with the genes will get T2DM), locus and allelic heterogeneity, and
the likelihood of phenocopies (multiple different genetic causes and nongenetic causes), as
well as the likelihood that many genes interacting with the environment contribute to T2DM
have conspired to confound the search for T2DM susceptibility loci. Thus, T2DM remains as
geneticist’s “Gordian knot”.

Two broad approaches have been used to define the genetic predisposition of T2DM. First,
molecular events in diabetes pathogenesis have been examined directly by testing the role of
sequence variants of specific candidate genes. The candidate gene approach focuses on the
search for an association between T2DM and sequence variants in or near biologically defined
candidate genes which have been chosen based on their known physiological function. The
importance of these variants or other nearby variants is tested by comparing the frequency in
T2DM subjects and control individuals. Despite of the apparent simplicity of study design and
NIH-PA Author Manuscript

power of the study, candidate gene studies are necessarily limited to the investigation of known
genes that are already considered important in glucose homeostasis. Our partial knowledge of
the mechanisms contributing to T2DM and the complicated pathways involved in those known
mechanisms make the candidate gene approach even more daunting. Although the wealth of
information on pathways involved in T2DM pathophysiology is accumulating, and new
bioinformatics approaches may guide the selection of potential susceptibility genes, novel
genes and pathways cannot be identified by this approach.

To overcome the shortcomings of the candidate gene studies, investigators have applied a
genome-wide linkage scan strategy in which regularly spaced markers are traced in families
and sibling pairs for segregation with T2DM. No prior knowledge of gene or gene effects is
necessary, but the genetic locus must have sufficient impact on the disease susceptibility to be
detectable. The susceptibility locus is first localized to a chromosomal region. The specific
gene and sequence variant are then identified within this region by a positional candidate gene

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 8

approach in which genes are selected within the region for a physiologic impact. Alternatively,
an unbiased approach of testing a dense map of markers in cases and controls across the region
of linkage may be used. Most recently, technological advances have made a new approach
NIH-PA Author Manuscript

possible which combines the power to detect small effects that characterizes candidate gene
studies with the unbiased approach of the linkage study. This approach, known as the genome-
wide association, has been applied in a limited fashion. For example, a recent study (82)
identified a variant in the insulin induced gene 2 (INSIG2) that was associated with obesity in
children and adults in several populations using the genome-wide association approach.

Over 27 genome wide linkage scans have been published in multiple ethnic groups including
Mexican-American, Caucasian, Native American, Chinese, Japanese, Indian and African based
populations, and in both founder and out-bred populations. Family configurations have
included extended pedigrees, nuclear families, and simple affected (diabetic) sib pairs, and
both definitions of “affected” and the analytical approaches have varied. Other scans have
examined intermediate traits related to T2DM. The scans and best replicated regions are
summarized in Table 1. Despite such differences, at least 8 susceptibility regions show at least
some evidence for replication in our opinions and appear likely to harbor one or more
susceptibility genes. Many other regions with some evidence of linkage can be defined (see
Table 1), with much of the genome potentially identified. In consideration of brevity, we restrict
our review to those regions with the best evidence for replication.
NIH-PA Author Manuscript

Chromosome 2q
The first locus for T2DM was mapped by Hanis et. al. in 1996 (83) to the telomeric region of
chromosome 2q (2q33–q37) in 330 Mexican American sib pairs and called NIDDM1. Further
fine mapping placed NIDDM1 near D2S140 (263.6 cM; multipoint lod score=4.03) with a 1-
lod support interval from 257 to 269 cM, a 12-cM region that included marker D2S125. In
subsequent studies, Cox and colleagues (84) used a two–locus conditional search strategy to
show that NIDDM1 interacted with a chromosome 15 locus near CYP19, with an increase in
the evidence for linkage from a logarithm of odds ratio of linkage to chance segregation (LOD)
of 1.3 to 4.0 when NIDDM1 was taken into account. Similarly, evidence for linkage at
NIDDM1 increased from 4.03 to 7.28 by conditioning on chromosome 15 linkage and
decreased the 1-lod support interval (the size of the probable region harboring the susceptibility
gene) from 12 cM to 7 cM (from 257–269 cM to 259–266 cM). Studies to identify a
susceptibility gene in this region (calpain 10) are reviewed below. Other studies, including
those from our laboratory (85) and the original Mexican American linkage study (83), have
found suggestive evidence for linkage of T2DM to chromosome 2q in a second, more
centromeric region from the CAPN10 gene. Other studies with some evidence for linkage have
included Han Chinese (86) at marker D2S126 (221 cM), Japanese families selected for a BMI<
30 kg/m2 at 2q34 (223 cM) (87), and indigenous Australian families at 2q24.3 (177 cM; LOD
NIH-PA Author Manuscript

score 3.9, 88). The gene or geness responsible for these observations are unknown.

Chromosome 20
Chromosome 20 was among the first regions with evidence for linkage in multiple studies
(89–91). A large genome scan in 716 affected sib pairs from 477 Finnish families (The Finland–
U.S. Investigation of Non-Insulin- Dependent Diabetes Mellitus Genetics or FUSION study)
provided considerable additional evidence (92), and subsequent studies have added to the
replication in Ashkenazy Jewish families (93) and in African sib pairs (94). In the large
FUSION study, the evidence for linkage was observed on both the long and short arms, with
multipoint lod scores of 3.08 on 20p (19.5cM) and 2.06 on 20q (57cM). Further high resolution
mapping and conditioning on a chromosome 2p locus (8.5cM) provided further support for a
locus on 20q at 69 cM (lod 5.50, 95). Linkage in the 472 sib pairs of Ashkenazy Jewish descent
was similarly in the region form 51 cM to 62 cM, with a much weaker signal on 20p (2 cM,

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 9

93). A genome wide scan on Han Chinse families also found evidence for linkage near marker
D20S196 on 20q in the subset of families with low BMI (86), and a similar subset analysis in
Japanese families with BMI <30kg/cm2 provided additional evidence for T2DM linkage at
NIH-PA Author Manuscript

20q12–q13 (D20S119 at 61.77 cM; MLS 2.32, 87). In 343 sibling pairs from West Africa,
highly suggestive linkage was found at 96 cM (D20S171, lod 2.63) (94). At least two strong
candidate genes (PTPN1 and HNF4A) show evidence for an association with T2DM, and are
reviewed below. However, additional susceptibility genes seem likely to explain this well
replicated linkage across a broad region of chromosome 20.

Chromosome12
A T2DM susceptibility loci on chromosome 12q near HNF1α was reported first by Mahtani
et al (96) in a subset analysis of families from Western Finland that had the lowest mean insulin
levels. In an extension of this initial study by using 32 additional families, for a total of 54
multiplex families, Lindgren et al replicated the initial findings at chromosome 12q24 (97),
and support a locus on 12q near 130–150cM. In the West African study of affected sibling
pairs, a T2DM susceptibility locus was also reported at 12q24 (D12S2070 at 125.31cM;
LOD=1.92, 94), and Ehm et al reported a locus on chromosome 12q in the Caucasian subset
of families collected under the American Diabetes Association’s Genetics of NIDDM
(GENNID) study, albeit somewhat centromeric to the earlier findings at 95 cM (lod 2.81, 98).
Other reports of linkage provide additional support, albeit many findings are to a more
centromeric region on chromosome 12 (50–100cM, 89,99–102). These findings suggest
NIH-PA Author Manuscript

substantial replication to chromosome 12, but the gene or genes responsible remain unknown
and the actual number of loci is uncertain.

Chromosome 1q
The best replicated locus at 1q21–q23 was initially identified independently by our laboratory
in North European Caucasians (85) and in Pima Indians with onset before age 25 years (103).
Subsequent replications of this region included Amish (104), French (105), United Kingdom
(106), US Caucasians (Framingham, 107), two Chinese populations (108,109) and Arkansas
African-American families (unpublished). In contrast to chromosomes 20 and 12, the
chromosome 1 linkage peaks overlap in a relatively narrow region extending from 165 cM to
200 cM (110). This well replicated region of T2DM linkage is characterized by an extraordinary
gene density, including a plethora of strong candidate genes. We recently reported that this
region encompasses at least two and possibly three linkage peaks (111). We have reported that
two genes in this region, liver pyruvate kinase (PKLR, 112) and calsequestrin1 (CASQ1, 113),
were associated with T2DM. Both have been observed in other populations (114), and support
a model of multiple susceptibility loci accounting for the chromosome 1 linkage signal. Work
is ongoing to identify additional susceptibility genes in this region as an international
NIH-PA Author Manuscript

collaborative effort.

Chromosome 18p
A locus on 18p (0–20cM) is likewise now well replicated. This region was suggestive in a scan
from our laboratory (85). A subgroup analysis of the 20% of the most obese subjects from a
study of 480 Finnish-Swedish Caucasian affected sib-pairs showed linkage to 18p11 (lod= 3.82
at D18S976 – D18S391, 115). Two subsequent studies from the Netherlands also suggested a
locus that interacted with obesity. Aulchenko et al (116) reported a lod score of 2.3 at marker
D18S63. When association was tested, the most significant findings were for subjects with
high BMI. In a study from van Tilburg and colleagues (117), the linkage was again for the 20%
of the most obese pedigrees at 18p11, between markers D18S471 and D18S843. Hence, the
putative T2DM susceptibility locus at 18p11 is replicating well, but may be of most importance
in interaction with obesity.

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 10

Chromosomes 3q, 11q and 8p


Several other loci show reasonable evidence for replication, and many cross ethnic groups. A
locus at 3q22–q27 near the adiponectin locus (165–220cM) has replicated well, and
NIH-PA Author Manuscript

adiponectin itself may be partially responsible (88,105,118). A broad region of 11q has been
implicated in several studies, and obesity also maps to this region (95,103,119,120). In Pima
Indians, the strongest 11q linkage was for a bivariate trait of obesity (BMI) and T2DM (103).
Finally, studies in British families, Utah families of European descent, and indigenous
Australian populations suggest an 8th possible locus at 8p21–p23 (85,88,106). With the
exception of adiponectin, genes accounting for these linkage signals have not been identified.

Established T2DM Susceptibility Genes


Genes Identified by Linkage Studies
Horikawa Y and colleagues (121) used a relatively sparse linkage disequilibrium map of SNPs
in the region of linkage on chromosome 2q to identify 3 common intronic variants of a
previously unknown gene. The gene, calpain 10 (CAPN10), belonged to the family of calcium-
activated, non-lysosomal thiol proteinases. Both a single intronic variant (UCSNP-43: G to A)
and a specific haplotype combination defined by three polymorphisms (UCSNP-43, -19, and
-63) were associated with T2DM in Mexican Americans, with lesser evidence of an association
in a Northern European population from the Botnia region of Finland (121). Surprisingly,
individuals with a combination of two different haplotypes (121/112) were at the highest risk
NIH-PA Author Manuscript

of T2DM. In further support of CAPN10 as a susceptibility gene, Baier and colleagues (122)
showed altered gene transcription and reduced muscle mRNA levels in muscle biopsies from
Pima Indians with T2DM. Recent studies provide further support for a role of CAPN10 in
T2DM, even among European derived populations (123). SNP-44 (designated “CAPN10-
g4841TtoC”; minor allele frequency 16%), located in intron 3 and 11 bp from SNP-43, was
independently associated with T2DM in several populations, and was in linkage disequilibrium
with the missense mutation Thr504Ala and two 5′-UTR variants (SNP-134 and SNP-135). An
initial meta-analysis also supported an association with T2DM (124), but a subsequent meta-
analysis of over 5000 cases and 5000 controls suggested an effect size that was clearly less
than suggested in the first studies (125) and raised questions about the role in T2DM.
Nonetheless, we and others have found support for a role of CAPN10 in the association with
intermediate, quantitative traits (126;127). Furthermore, CAPN10 variants were valuable along
with traditional risk factors in predicting the onset of T2DM in a prospective study of Botnian
Finnish individuals (128).

Although initial studies of the HNF4α gene, an obvious candidate on chromosome 20, failed
to find any association with T2DM (92), subsequent studies based on the discovery of an
upstream, β-cell specific promoter and first exon prompted a reevaluation. Two studies found
NIH-PA Author Manuscript

evidence for association with T2DM that could partially explain the linkage signal (129,130).
HNF4A has a complex expression pattern, which includes elaborate alternative splicing, and
is expressed in many tissues, including the liver and pancreas. Three of the isoforms are
transcribed by an alternative P2 promoter, located ~46 kb upstream of the P1 promoter and the
rest of the coding exons. Transcripts from both the P1 and P2 promoters have been detected
in pancreatic β-cells, but the P2 promoter is suggested to be the primary transcription start site
in these cells. In total, 4 SNPs flanking the P2 promoter were associated with T2DM in both
populations. Subsequent replication has been inconsistent. Vaxillaire et al failed to replicate
in a French case control study (131). Hansen and colleagues found only modest evidence for
one of 4 SNPs originally reported (p=0.02, OR=1.15) in a large Danish study (132), and
Weedon et al found a similar effect in a large British study (5256 subjects; p=0.02, OR=1.15,
133). Damcott et al reported replication in the Old Order Amish (134), and Muller and
colleagues reported replication in Pima Indians with OR 1.3 (135). Winckler and colleagues,

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 11

on the other hand, found no replication in 7883 Caucasians from several populations (136).
Although these SNPs appear to have a role in T2DM susceptibility, the effect is likely much
smaller than originally reported, and likely similar to that of CAPN10.
NIH-PA Author Manuscript

The PTPN1 gene (chromosome 20q13) codes for protein tyrosine phosphatase 1B (PTP1B),
which negatively regulates insulin signaling by dephosphorylating the phosphotyrosine
residues of the insulin receptor kinase activation domain. Multiple noncoding SNPs in the
PTPN1 gene (PTP1B) on chromosome 20q13 were recently implicated in T2DM in Caucasian
and Mexican-American populations (137,138). The association was observed between
multiple SNPs and T2DM, with the most consistent evidence for association occurring with
SNPs spanning the 3′ end of intron 1 of PTPN1 through intron 8 (p values ranging from 0.043
to 0.004 in one case-control set and 0.038–0.002 in a second case-control set). All of the
associated SNPs were present in a single 100-kb haplotype block that encompassed the
PTPN1 gene. Haplotype frequencies were significantly different between T2DM case and
control subjects (p=0.0035–0.0056), with a single common haplotype (36%) contributing
strongly to the evidence for association, with OR 1.3. Furthermore, the same haplotypes were
associated with glucose homeostasis measures in 811 Hispanic subjects (138). However, Florez
et al failed to replicate the findings in a large Caucasian study (139). Further study is needed
to determine the role of the strong candidate in a region of replicated linkage.

The search for diabetes susceptibility genes on most other chromosomes is ongoing. On
NIH-PA Author Manuscript

chromosome 1, we initially described an association of polymorphisms of the liver pyruvate


kinase gene with T2DM (112); subsequent preliminary reports have confirmed that association
in Older Order Amish (140) and in other populations with chromosome 1 linkage (141). The
region has extended linkage disequilibrium, such that several potential candidate genes are
included within the two associated haplotypes. Work is ongoing in our laboratory and by others
to identify the responsible variants. Closer to the most prominent chromosome 1 linkage peak,
we and Fu et al independently showed an association of noncoding variants of calsequestrin
gene (CASQ1) with T2DM in populations of European ancestry (113,114). The effect size is
likely similar to that of CAPN10 or HNF4A. The role of this muscle gene, which is not an
obvious candidate for T2DM, must await replication in other populations, however.
Adiponectin is a strong candidate for T2DM given the clear role of plasma adiponectin levels
in insulin sensitivity and the fall in adiponectin levels with obesity and T2DM. Adiponectin
maps to the region of replicated linkage on chromosome 3q. A very large number of studies
have examined adiponectin variants with metabolic traits, T2DM, and risk of conversion of
impaired glucose tolerance to T2DM. Multiple studies found evidence that promoter variants
in the APM1 gene were associated with T2DM in French and Swedish Caucasian and Japanese
populations, and some data support a role of the APM1 locus in controlling adiponectin levels
(142). However, APM1 variants did not appear to account for the 3q27 linkage in French
NIH-PA Author Manuscript

families (142). Zacharova and colleagues showed that APM1 variation at positions 45 (G allele)
in exon 2 and 276 in intron 2 (T allele) were associated with a 4.5 fold increased risk of
converting from impaired glucose tolerance to T2DM in the STOP-NIDDM trial (143). On the
other hand, no association was observed in Pima Indians (144). Larger studies are needed to
determine the magnitude of risk imparted by APM1 variation.

Grant and colleagues (145) searched for a T2DM susceptibility gene under the suggestive
linkage peak on chromosome 10q in an Icelandic population. They identified a 6 allele simple
tandem repeat (STR) polymorphism in intron 3 of the transcription factor 7 like 2 gene
(TCF7L2, formerly TCF4). Impressive associations were found in a Danish population and a
United States Caucasian population, with a relative risk of 1.45 for heterozygotes and 2.41 in
homozygotes. Although the exact role of TCF7L2 is unknown, it appears to act through the
Wnt pathway to regulate proglucagon gene expression in enteroendocrine cells. The authors
calculated a relatively high 27% population attributable risk (the proportion of T2DM that

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 12

would not be present if the gene could be removed from the population). Two additional
noncoding SNPs were closely correlated with the simple tandem repeat composite risk allele.
Additional replication is needed to determine the real role of this novel gene.
NIH-PA Author Manuscript

Despite initial identification through a linkage signal, which would be expected to select
relatively strong effects, each of these variants reviewed above has only modest individual
effect (Odds Ratio, OR<1.4). Indeed, each with further study has shown much smaller effect
than in the initial report – a phenomenon named the “winner’s curse”. Available data suggest
that 1) the effects of individual variants are likely small; 2) multiple variants probably
contribute to replicated linkage signals, as we have demonstrated on chromosome 1q; and 3)
most variants are noncoding and regulatory rather than altering protein structure. We predict
that each linkage region will in fact harbor multiple disease genes of small effect, and that no
single variant will explain the linkage peak.

Genes Discovered by the Candidate Gene Approach


Extensive studies have been conducted of genes identified through physiologic pathways.
Consistent with the very large numbers of studies, many have shown modest associations in
single populations, but these associations have not yet been replicated. We make no attempt
to provide a comprehensive review of the hundreds of genes examined by investigators around
the world, and instead focus on genes for which reasonable replication has been achieved
already. In each case, like those genes identified through linkage, the effect of these variants
NIH-PA Author Manuscript

is relatively small, with OR<1.4 and in most cases approaching 1.2 or less. Perhaps as a bias
of candidate gene studies, these variants are more often coding.

The peroxisome proliferator activated receptor gamma (PPARγ) is a member of nuclear


hormone receptor superfamily of transcription factors, and the target of the widely used class
of insulin sensitizers, the thiazolidinediones. Multiple studies have examined a common amino
acid change in the PPARγ2 isoform in which alanine is substituted for proline at position 12
(P12A). Although smaller studies provided inconsistent results, larger studies have shown a
consistently protective effect of the rare, alanine allele (A/A) when compared with the common
proline allele. Hence, the common P12 allele consistently increases T2DM risk by an OR of
1.2–1.3 in cross sectional studies (146). Furthermore, in the Nurses Health Study, the OR of
carrying the Ala12 allele was 0.72, again suggesting a protective role (147). Several, but not
all studies suggest that the rare “A” allele protects against insulin resistance and obesity, but
the association with obesity has been inconsistent. In studies of families from Utah, we found
an association with several “metabolic syndrome” traits, including BMI, serum total cholesterol
levels, triglyceride levels, systolic and diastolic blood pressures, and glucose concentrations.
Gene-environment interactions have been suggested, and may explain the many inconsistent
results. Because the common allele is also the risk allele for T2DM, and this allele has a
NIH-PA Author Manuscript

frequency of at least 75% in most populations, the population attributable risk is again
extremely high at 25%. Hence, the small individual effect of this variant translates into a large
effect in the context of the population.

The β-cell potassium channel comprises two subunits, the potassium channel encoded by the
gene KCNJ11, and the regulatory subunit (SUR1 encoded by ABCC8) that binds sulfonylureas
and ATP. Variants in both subunits have been associated with T2DM. Two variants in
ABCC8 were initially associated in smaller studies (148–150), and larger subsequent studies
appeared to confirm the association (151,152). Other studies have not confirmed the
association, however (153,154). We (155) and others (156) have shown altered insulin
secretion associated with these variants. The adjacent KCNJ11 gene was initially noted to have
multiple nonsynonymous coding variants (157), but none were associated with T2DM in this
or other early studies (146,158). However, a coding glutamine to lysine at position 23 in the
single exon (E23K) was shown to lower the sensitivity of the potassium channel to ATP and

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 13

thus reduce insulin secretion in vitro (159). Gloyn et al subsequently showed a modest
association of the E23K variant with T2DM in nearly 2500 British case-control subjects
(153), and in a meta-analysis confirmed association with an OR of 1.16. More recently Florez
NIH-PA Author Manuscript

and colleagues (154) thoroughly examined the two subunits (KCNJ11, ABCC8), and tested
sufficient variants to completely tag all variation in this region. Again, the E23K variant was
associated with T2DM in 3,413 subjects, and the association was confirmed in a meta-analysis
that included over 5000 T2DM subjects and 4747 controls (p<10−5, OR 1.15). The E23K
variant is in nearly complete linkage disequilibrium (completely correlated) with a Serine to
Alanine change at position 1369 in exon 33 of the sulfonylurea receptor, and thus S1369A and
E23K cannot be distinguished genetically. Nonetheless, E23K appears to have a role
comparable to the PPARG P12A polymorphism. The appropriate genetic model is close to
recessive, in which the relatively rare K23 homozygotes have the highest risk, but E23K
heterozygotes are at some increased risk. Some (160) but not all (161) studies have found an
effect of the K allele on insulin secretion.

Other variants in candidate genes have been more difficult to establish. The insulin gene
variable tandem repeat polymorphism, which has clearly been associated with type 1 diabetes,
and been controversial for over 20 years in T2DM (162). Whereas considerable evidence
supports an association with obesity and birth weight (163–165), the association with T2DM
is inconsistent (165–167). Similarly, both physiologic and genetic data are inconsistent for the
glycine to arginine (G972R) variant of the insulin receptor substrate 1(IRS1) gene (168,169),
NIH-PA Author Manuscript

now more correctly referred to as G971R. Despite evidence that this variant alters insulin action
on IRS1 (169), the association with T2DM has been inconsistently replicated. Two recent large
studies have revisited this variant. Zeggini and colleagues (170) examined 971 cases and 1257
controls from the United Kingdom, and found no association. However, a meta-analysis of 31
studies that included 5104 cases and 7,418 controls showed a modest association (p=0.025)
with an OR of 1.15, suggesting that the effect size was similar to or smaller than that of
KCNJ11 and PPARG. Florez et al (171) three large Caucasian samples with 9000 total
individuals, including 4279 cases and 3532 controls, and found no evidence for an association
with T2DM. If the IRS1 G972R (G971R) variant has any role in T2DM, clearly the role is quite
small and of minimal clinical significance. Similarly, the association of a glycine to serine
variant at position 482 of the PGC1α?gene has been reported by several groups, but not yet
well replicated. PGC1α??as reviewed above, plays key roles in liver gluconeogenesis and was
implicated in array studies of diabetic and control muscle. Positive associations were reported
in JapaneseT2DM (172) and Danish cases and controls (173), and subsequently some
replication was seen in populations from India (174) and Korea (175). Other associations were
reported with insulin resistance (176) elevated glucose in nondiabetic offspring of diabetic
parents (177), and reduced insulin secretion (178), but not with metabolic syndrome (179).
Additional studies are required to determine the real role of this variant in diabetes
NIH-PA Author Manuscript

susceptibility.

Variants in the gene ectonucleotide pyrophosphatase phosphodiesterase 1 (ENPP1), also


known as plasma cell glycoprotein 1, were shown recently to be associated with both childhood
and adult obesity in a French population. A moderate excess of the risk haplotype was also
seen in adults with T2DM compared with controls (10.7% vs 7.1%, OR 1.44), and the
association was replicated in an Austrian cohort with similar frequencies. A Mantel-Haenszel
analysis of 2,569 European subjects supported the association, with OR 1.56 and p=0.00002
(180). The nonsynonymous variant lysine 212 to glutamine (K121Q) has been best studied,
and also shown to be associated with obesity but not diabetes in Caucasian and African
American subjects ascertained from the New York Cancer Project (181). Other investigators
found an association of K121Q with earlier T2DM onset and coronary disease among T2DM
patients (182). The association of K121Q with T2DM was replicated in three relatively small
populations, two of South Asian ancestry and one Caucasian (183). Although the K121Q

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 14

variant appears to associate with T2DM and obesity, additional work is needed to determine
the magnitude of the risk, the populations affected, and to confirm a role comparable to
KCNJ11 or PPARG.
NIH-PA Author Manuscript

Conclusions: Placing Genetic Susceptibility in Context


In this review, we have restricted discussion to those genes with the best evidence for a role in
T2DM, and have not considered a multitude of other associations, many reported by our
laboratory and others (151), which suggest that the eventual genetic landscape of T2DM will
integrate a large number of genes, many with multiple variants. The completed work and the
few susceptibility genes that are now well validated suggest that these susceptibility variants
will have small effects, most with odds ratios below 1.4 and likely below 1.2. As these
susceptibility genes inhabit different pathways in the complex physiologic network that causes
T2DM, we expect many of the variants will interact. We propose that much as very small
changes in gene expression in the oxidative phosphorylation pathways appear to result in
decreased oxidative phosphorylation and mitochondrial function (52), multiple defects of
individually small effect will combine in a single pathway to have large physiologic effect. We
propose that broad linkage peaks, as found on chromosome 20 and chromosome 1q, in fact
represent multiple genes, many in either the same pathway or in interacting pathways. Sorting
out these complex gene-gene interactions will require sizeable populations, however.
NIH-PA Author Manuscript

Although we have focused on gene variation as the cause of T2DM, the role of environment
in the current diabetes pandemic is undeniable. Different environmental stresses, population
differences in activity, and different diets clearly cause some genes to manifest as a disease
phenotype. For example, fish oil may stimulate PPARγ in much the same fashion as the
thiazolidinedione class of drugs; studies of the interaction of the P12A variant with dietary
composition have not been performed. The spectacular rise in rates of diabetes among Pima
Indians and other populations as they adopt Western diets and life styles dramatically
demonstrates the key role of environment. Consequently, the effect of a common gene variant
between populations that have very different diets and exercise habits might be expected to be
totally different, thus explaining some instances of lack of replication.

In a recent study, Ioannidis et al (184), examined the genetic effects for 43 validated gene-
disease associations across 697 studies of varied ancestry. They argued that proposed racial
differences in genetic risk should be scrutinized cautiously, and that genetic effects are usually
consistent across human populations. Instead, they argued, small sample sizes, flawed study
designs, or other biases may be more common reasons than true ‘racial’ heterogeneity for the
observed discrepancies. Nonetheless, marked differences in allele frequencies are clearly
evident in known susceptibility genes, and ethnic differences in disease prevalence are also
NIH-PA Author Manuscript

common. We predict that ethnic “specific” risk factors will indeed be present, although
differences in the degree of risk, allele frequency, and gene-gene or gene-environment
interactions. Clearly much remains to be learned regarding these issues for complex disease.
Our tacit application of models derived from our understanding of monogenic disease may
retard our understanding of common complex diseases such as T2DM.

Whereas the focus on adult diseases such as T2DM is on the role of post-natal or post-pubertal
environment, considerable data that cannot be reviewed here suggest a role for genes that
control fetal growth and intrauterine environment in T2DM. For example, Beck-Nielsen
(185) demonstrated in a group of middle aged, monozygotic twins that the twin with T2DM
was consistently the lighter of the pair at birth. Hence, even in genetically identical individuals,
the intrauterine environment may influence fetal growth differentially with implications for
subsequent adult metabolic health. When considered in the context of numerous studies of birth
weight and the risk for diabetes and metabolic syndrome, this study suggests that gene-

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 15

environment interactions may need to consider intrauterine environment in addition to the


external environment.
NIH-PA Author Manuscript

With the human genome sequence completed in 2003, new data are rapidly expanding on the
noncoding aspects of our genome. T2DM and most other complex diseases appear to be caused
primarily by noncoding rather than coding variants. Thus, KCNJ11, PPARG, IRS1, and
ENPP1 coding variants discussed above are likely the exception, and their prominence
reflecting a strong publication bias to report and validate changes in protein structure while
ignoring noncoding variants as silent. This bias is likely unfounded. Noncoding SNPs or
microsatellites are clearly the only explanation for a multitude of complex disease genes,
including HNF4A, PTPN1, CASQ1, PKLR, and CAPN10 reviewed above among many others.
Whereas not long ago such findings were viewed as implausible, recent studies (186–188) have
revolutionized our understanding of the genome, and suggest that much noncoding DNA is
indeed functional as long range enhancer elements controlling gene transcription, noncoding
RNA that may have regulatory functions, and small RNA species that have widespread
regulatory properties and are often derived from intronic sequences of coding genes.
Furthermore, over 70% of genes likely undergo alternative splicing and generate multiple
protein forms (186). The balance between those splice forms often correlates with noncoding
SNPs that may have profound effects on phenotype. Given the likely extensive role of intronic
and intergenic DNA in determining phenotype, a major role of sequence variants in noncoding
regions in T2DM pathogenesis should be anticipated. Ignoring this wealth of knowledge will
NIH-PA Author Manuscript

impede our continued quest to understand the role of genes in T2DM susceptibility. Our
increased understanding of such phenomena will open new doors to understanding how
common variants can alter disease susceptibility, and will be essential in understanding the
physiologic importance of the genetic associations that we uncover.

Clinical Implications
The molecular genetic definition of T2DM pathophysiology will certainly satisfy our
intellectual curiosity about this common disease, but will it impact health care? Our knowledge
to predict T2DM is currently limited. Pharmacological intervention rather than prevention
remains the primary approach to this disease. In a recent study, Lyssenko et.al.(128) combined
variants in only two genes, PPARG and CAPN10, both reviewed above, with traditional risk
factors. These investigators achieved a remarkable 21.2 –fold increased risk (estimated by
hazard ratio-HR) in obese carriers of the PPARG P12/P12 genotype and CAPN10 SNP43/44
GG/TT genotypes along with elevated fasting plasma glucose (FPG). This study is a landmark
in combining individually weak genetic risk factors with traditional risk factors to improve the
prediction of future T2DM. Although the combination of genetic risk factors in this model has
been challenged (189), the methods of Lyssenko and colleagues model the was in which
NIH-PA Author Manuscript

multiple small genetic risk factors can be combined to predict risk. Future predictive testing
might include testing a battery of common genetic variants, followed by computational joint
assessment of traditional environmental and lifestyle risk factors, which may assist physicians
and patient in choosing a preventive strategy. If personalized medicine is to be realized in the
future for complex diseases, such models will be required. The future challenges for the
genetics of T2DM will include not only the cataloging of what can now be recognized as likely
100 or more genetic susceptibility variants, but more importantly the synthesis of this catalog
with environmental risk factors, gene-gene, and gene-environment interactions. The utility of
genetic approaches will depend on a holistic understanding of the interactions of among genetic
variants, and between genetic variants and the environmental, lifestyle factors, treatments, and
perhaps even intrauterine environment. This integrated approach must replace the current
methods of using a reductionist approach to understand each component separately.

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 16

References
1. Roglic G, Unwin N, Bennett PH, Mathers C, Tuomilehto J, Nag S, Connolly V, King H. The burden
NIH-PA Author Manuscript

of mortality attributable to diabetes: realistic estimates for the year 2000. Diabetes Care 2005;28:2130–
2135. [PubMed: 16123478]
2. Wild S, Roglic G, Green A, Sicree R, King H. Global prevalence of diabetes: estimates for the year
2000 and projections for 2030. Diabetes Care 2004;27:1047–1053. [PubMed: 15111519]
3. Smyth S, Heron A. Diabetes and obesity: the twin epidemics. NatMed 2006;12:75–80.
4. Yach D, Stuckler D, Brownell KD. Epidemiologic and economic consequences of the global epidemics
of obesity and diabetes. NatMed 2006;12:62–66.
5. Diamond J. The double puzzle of diabetes. Nature 2003;423:599–602. [PubMed: 12789325]
6. Meigs JB, Cupples LA, Wilson PW. Parental transmission of type 2 diabetes: the Framingham
Offspring Study. Diabetes 2000;49:2201–2207. [PubMed: 11118026]
7. Barnett AH, Eff C, Leslie RD, Pyke DA. Diabetes in identical twins. A study of 200 pairs Diabetologia
1981;20:87–93.
8. Newman B, Selby JV, King MC, Slemenda C, Fabsitz R, Friedman GD. Concordance for type 2 (non-
insulin-dependent) diabetes mellitus in male twins. Diabetologia 1987;30:763–768. [PubMed:
3428496]
9. Poulsen P, Kyvik KO, Vaag A, Beck-Nielsen H. Heritability of type II (non-insulin-dependent) diabetes
mellitus and abnormal glucose tolerance--a population-based twin study. Diabetologia 1999;42:139–
145. [PubMed: 10064092]
NIH-PA Author Manuscript

10. Medici F, Hawa M, Ianari A, Pyke DA, Leslie RD. Concordance rate for type II diabetes mellitus in
monozygotic twins: actuarial analysis. Diabetologia 1999;42:146–150. [PubMed: 10064093]
11. Vaag A, Henriksen JE, Madsbad S, Holm N, Beck-Nielsen H. Insulin secretion, insulin action, and
hepatic glucose production in identical twins discordant for non-insulin-dependent diabetes mellitus.
JClinInvest 1995;95:690–698.
12. Elbein SC, Hasstedt SJ, Wegner K, Kahn SE. Heritability of pancreatic beta-cell function among
nondiabetic members of Caucasian familial type 2 diabetic kindreds. JClinEndocrinolMetab
1999;84:1398–1403.
13. Elbein SC, Wegner K, Kahn SE. Reduced beta-cell compensation to the insulin resistance associated
with obesity in members of caucasian familial type 2 diabetic kindreds. Diabetes Care 2000;23:221–
227. [PubMed: 10868835]
14. Gerich JE. The genetic basis of type 2 diabetes mellitus: impaired insulin secretion versus impaired
insulin sensitivity. EndocrRev 1998;19:491–503.
15. DeFronzo RA, Bonadonna RC, Ferrannini E. Pathogenesis of NIDDM. A balanced overview Diabetes
Care 1992;15:318–368.
16. Stumvoll M, Goldstein BJ, van Haeften TW. Type 2 diabetes: principles of pathogenesis and therapy.
Lancet 2005;365:1333–1346. [PubMed: 15823385]
17. Weyer C, Bogardus C, Mott DM, Pratley RE. The natural history of insulin secretory dysfunction
NIH-PA Author Manuscript

and insulin resistance in the pathogenesis of type 2 diabetes mellitus. J ClinInvest 1999;104:787–
794.
18. Ferrannini E. Insulin resistance versus insulin deficiency in non-insulin- dependent diabetes mellitus:
problems and prospects. EndocrRev 1998;19:477–490.
19. Kahn SE. The relative contributions of insulin resistance and beta-cell dysfunction to the
pathophysiology of Type 2 diabetes. Diabetologia 2003;46:3–19. [PubMed: 12637977]
20. Minokoshi Y, Kahn CR, Kahn BB. Tissue-specific ablation of the GLUT4 glucose transporter or the
insulin receptor challenges assumptions about insulin action and glucose homeostasis. JBiolChem
2003;278:33609–33612.
21. Fisher SJ, Kahn CR. Insulin signaling is required for insulin’s direct and indirect action on hepatic
glucose production. JClinInvest 2003;111:463–468.
22. Michael MD, Kulkarni RN, Postic C, Previs SF, Shulman GI, Magnuson MA, Kahn CR. Loss of
insulin signaling in hepatocytes leads to severe insulin resistance and progressive hepatic dysfunction.
MolCell 2000;6:87–97.

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 17

23. Mauvais-Jarvis F, Kulkarni RN, Kahn CR. Knockout models are useful tools to dissect the
pathophysiology and genetics of insulin resistance. ClinEndocrinol(Oxf) 2002;57:1–9.
24. Baudry A, Leroux L, Jackerott M, Joshi RL. Genetic manipulation of insulin signaling, action and
NIH-PA Author Manuscript

secretion in mice. Insights into glucose homeostasis and pathogenesis of type 2 diabetes. EMBO Rep
2002;3:323–328. [PubMed: 11943762]
25. Accili D. Lilly lecture 2003: the struggle for mastery in insulin action: from triumvirate to republic.
Diabetes 2004;53:1633–1642. [PubMed: 15220184]
26. Shepherd PR, Kahn BB. Glucose transporters and insulin action--implications for insulin resistance
and diabetes mellitus. NEnglJMed 1999;341:248–257.
27. Abel ED, Peroni O, Kim JK, Kim Y-B, Boss O, Hadro E, Minnemann T, Shulman GI, Kahn BB.
Adipose-selective targeting of the GLUT4 gene impairs insulin action in muscle and liver. Nature
2001;409:729–733. [PubMed: 11217863]
28. Zisman A, Peroni OD, Abel ED, Michael MD, Mauvais-Jarvis F, Lowell BB, Wojtaszewski JF,
Hirshman MF, Virkamaki A, Goodyear LJ, Kahn CR, Kahn BB. Targeted disruption of the glucose
transporter 4 selectively in muscle causes insulin resistance and glucose intolerance. NatMed
2000;6:924–928.
29. Yang Q, Graham TE, Mody N, Preitner F, Peroni OD, Zabolotny JM, Kotani K, Quadro L, Kahn BB.
Serum retinol binding protein 4 contributes to insulin resistance in obesity and type 2 diabetes. Nature
2005;436:356–362. [PubMed: 16034410]
30. Gimeno RE, Klaman LD. Adipose tissue as an active endocrine organ: recent advances.
CurrOpinPharmacol 2005;5:122–128.
NIH-PA Author Manuscript

31. Lazar MA. How obesity causes diabetes: not a tall tale. Science 2005;307:373–375. [PubMed:
15662001]
32. Wellen KE, Hotamisligil GS. Inflammation, stress, and diabetes. JClinInvest 2005;115:1111–1119.
33. de Luca C, Olefsky JM. Stressed out about obesity and insulin resistance. NatMed 2006;12:41–42.
34. Lazar MA. The humoral side of insulin resistance. NatMed 2006;12:43–44.
35. Hotamisligil GS, Peraldi P, Budavari A, Ellis R, White MF, Spiegelman BM. IRS-1-mediated
inhibition of insulin receptor tyrosine kinase activity in TNF-alpha- and obesity-induced insulin
resistance. Science 1996;271:665–668. [PubMed: 8571133]
36. Aguirre V, Werner ED, Giraud J, Lee YH, Shoelson SE, White MF. Phosphorylation of Ser307 in
insulin receptor substrate-1 blocks interactions with the insulin receptor and inhibits insulin action.
JBiolChem 2002;277:1531–1537.
37. Aguirre V, Uchida T, Yenush L, Davis R, White MF. The c-Jun NH(2)-terminal kinase promotes
insulin resistance during association with insulin receptor substrate-1 and phosphorylation of Ser
(307). JBiolChem 2000;275:9047–9054.
38. Paz K, Hemi R, LeRoith D, Karasik A, Elhanany E, Kanety H, Zick Y. A molecular basis for insulin
resistance. Elevated serine/threonine phosphorylation of IRS-1 and IRS-2 inhibits their binding to
the juxtamembrane region of the insulin receptor and impairs their ability to undergo insulin-induced
tyrosine phosphorylation. JBiolChem 1997;272:29911–29918.
NIH-PA Author Manuscript

39. Weisberg SP, McCann D, Desai M, Rosenbaum M, Leibel RL, Ferrante AW Jr. Obesity is associated
with macrophage accumulation in adipose tissue. JClinInvest 2003;112:1796–1808.
40. Xu H, Barnes GT, Yang Q, Tan G, Yang D, Chou CJ, Sole J, Nichols A, Ross JS, Tartaglia LA, Chen
H. Chronic inflammation in fat plays a crucial role in the development of obesity-related insulin
resistance. JClinInvest 2003;112:1821–1830.
41. Ozawa K, Miyazaki M, Matsuhisa M, Takano K, Nakatani Y, Hatazaki M, Tamatani T, Yamagata
K, Miyagawa J, Kitao Y, Hori O, Yamasaki Y, Ogawa S. The endoplasmic reticulum chaperone
improves insulin resistance in type 2 diabetes. Diabetes 2005;54:657–663. [PubMed: 15734840]
42. Ozcan U, Cao Q, Yilmaz E, Lee AH, Iwakoshi NN, Ozdelen E, Tuncman G, Gorgun C, Glimcher
LH, Hotamisligil GS. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes.
Science 2004;306:457–461. [PubMed: 15486293]
43. Nakatani Y, Kaneto H, Kawamori D, Yoshiuchi K, Hatazaki M, Matsuoka TA, Ozawa K, Ogawa S,
Hori M, Yamasaki Y, Matsuhisa M. Involvement of endoplasmic reticulum stress in insulin resistance
and diabetes. JBiol Chem 2005;280:847–851. [PubMed: 15509553]

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 18

44. Harding HP, Ron D. Endoplasmic reticulum stress and the development of diabetes: a review.
Diabetes 2002;51(Suppl 3):S455–61. [PubMed: 12475790]S455–S461
45. Arkan MC, Hevener AL, Greten FR, Maeda S, Li ZW, Long JM, Wynshaw-Boris A, Poli G, Olefsky
NIH-PA Author Manuscript

J, Karin M. IKK-beta links inflammation to obesity-induced insulin resistance. NatMed


2005;11:191–198.
46. Cai D, Yuan M, Frantz DF, Melendez PA, Hansen L, Lee J, Shoelson SE. Local and systemic insulin
resistance resulting from hepatic activation of IKK-beta and NF-kappaB. NatMed 2005;11:183–190.
47. Lin Y, Berg AH, Iyengar P, Lam TK, Giacca A, Combs TP, Rajala MW, Du X, Rollman B, Li W,
Hawkins M, Barzilai N, Rhodes CJ, Fantus IG, Brownlee M, Scherer PE. The hyperglycemia-induced
inflammatory response in adipocytes: the role of reactive oxygen species. JBiolChem
2005;280:4617–4626.
48. Furukawa S, Fujita T, Shimabukuro M, Iwaki M, Yamada Y, Nakajima Y, Nakayama O, Makishima
M, Matsuda M, Shimomura I. Increased oxidative stress in obesity and its impact on metabolic
syndrome. JClinInvest 2004;114:1752–1761.
49. Urakawa H, Katsuki A, Sumida Y, Gabazza EC, Murashima S, Morioka K, Maruyama N, Kitagawa
N, Tanaka T, Hori Y, Nakatani K, Yano Y, Adachi Y. Oxidative stress is associated with adiposity
and insulin resistance in men. J ClinEndocrinolMetab 2003;88:4673–4676.
50. Houstis N, Rosen ED, Lander ES. Reactive oxygen species have a causal role in multiple forms of
insulin resistance. Nature 2006;440:944–948. [PubMed: 16612386]
51. Patti ME, Butte AJ, Crunkhorn S, Cusi K, Berria R, Kashyap S, Miyazaki Y, Kohane I, Costello M,
Saccone R, Landaker EJ, Goldfine AB, Mun E, DeFronzo R, Finlayson J, Kahn CR, Mandarino LJ.
NIH-PA Author Manuscript

Coordinated reduction of genes of oxidative metabolism in humans with insulin resistance and
diabetes: Potential role of PGC1 and NRF1. Proc NatlAcadSciUSA 2003;100:8466–8471.
52. Mootha VK, Lindgren CM, Eriksson KF, Subramanian A, Sihag S, Lehar J, Puigserver P, Carlsson
E, Ridderstrale M, Laurila E, Houstis N, Daly MJ, Patterson N, Mesirov JP, Golub TR, Tamayo P,
Spiegelman B, Lander ES, Hirschhorn JN, Altshuler D, Groop LC. PGC-1alpha-responsive genes
involved in oxidative phosphorylation are coordinately downregulated in human diabetes. NatGenet
2003;34:267–273.
53. Lowell BB, Shulman GI. Mitochondrial dysfunction and type 2 diabetes. Science 2005;307:384–387.
[PubMed: 15662004]
54. Petersen KF, Dufour S, Befroy D, Garcia R, Shulman GI. Impaired mitochondrial activity in the
insulin-resistant offspring of patients with type 2 diabetes. NEnglJMed 2004;350:664–671.
55. Maechler P, Wollheim CB. Mitochondrial function in normal and diabetic beta-cells. Nature
2001;414:807–812. [PubMed: 11742413]
56. Morino K, Petersen KF, Dufour S, Befroy D, Frattini J, Shatzkes N, Neschen S, White MF, Bilz S,
Sono S, Pypaert M, Shulman GI. Reduced mitochondrial density and increased IRS-1 serine
phosphorylation in muscle of insulin-resistant offspring of type 2 diabetic parents. JClinInvest
2005;115:3587–3593.
57. Krauss S, Zhang CY, Scorrano L, Dalgaard LT, St Pierre J, Grey ST, Lowell BB. Superoxide-mediated
activation of uncoupling protein 2 causes pancreatic beta cell dysfunction. JClinInvest
NIH-PA Author Manuscript

2003;112:1831–1842.
58. Chan CB, De Leo D, Joseph JW, McQuaid TS, Ha XF, Xu F, Tsushima RG, Pennefather PS, Salapatek
AM, Wheeler MB. Increased uncoupling protein-2 levels in beta-cells are associated with impaired
glucose-stimulated insulin secretion: mechanism of action. Diabetes 2001;50:1302–1310. [PubMed:
11375330]
59. Zhang CY, Baffy G, Perret P, Krauss S, Peroni O, Grujic D, Hagen T, Vidal-Puig AJ, Boss O, Kim
YB, Zheng XX, Wheeler MB, Shulman GI, Chan CB, Lowell BB. Uncoupling protein-2 negatively
regulates insulin secretion and is a major link between obesity, beta cell dysfunction, and type 2
diabetes. Cell 2001;105:745–755. [PubMed: 11440717]
60. Porte D Jr, Baskin DG, Schwartz MW. Insulin signaling in the central nervous system: a critical role
in metabolic homeostasis and disease from C. elegans to humans. Diabetes 2005;54:1264–1276.
[PubMed: 15855309]
61. Schwartz MW, Porte D Jr. Diabetes, obesity, and the brain. Science 2005;307:375–379. [PubMed:
15662002]

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 19

62. Seeley RJ, Tschop M. How diabetes went to our heads. NatMed 2006;12:47–49.
63. Obici S, Feng Z, Arduini A, Conti R, Rossetti L. Inhibition of hypothalamic carnitine
palmitoyltransferase-1 decreases food intake and glucose production. NatMed 2003;9:756–761.
NIH-PA Author Manuscript

64. Pocai A, Lam TK, Gutierrez-Juarez R, Obici S, Schwartz GJ, Bryan J, Aguilar-Bryan L, Rossetti L.
Hypothalamic K(ATP) channels control hepatic glucose production. Nature 2005;434:1026–1031.
[PubMed: 15846348]
65. Wolfrum C, Asilmaz E, Luca E, Friedman JM, Stoffel M. Foxa2 regulates lipid metabolism and
ketogenesis in the liver during fasting and in diabetes. Nature 2004;432:1027–1032. [PubMed:
15616563]
66. Matsumoto M, Accili D. The tangled path to glucose production. NatMed 2006;12:33–34.
67. Shaw RJ, Lamia KA, Vasquez D, Koo SH, Bardeesy N, Depinho RA, Montminy M, Cantley LC.
The kinase LKB1 mediates glucose homeostasis in liver and therapeutic effects of metformin. Science
2005;310:1642–1646. [PubMed: 16308421]
68. Koo SH, Flechner L, Qi L, Zhang X, Screaton RA, Jeffries S, Hedrick S, Xu W, Boussouar F, Brindle
P, Takemori H, Montminy M. The CREB coactivator TORC2 is a key regulator of fasting glucose
metabolism. Nature 2005;437:1109–1111. [PubMed: 16148943]% 20
69. Zhou XY, Shibusawa N, Naik K, Porras D, Temple K, Ou H, Kaihara K, Roe MW, Brady MJ,
Wondisford FE. Insulin regulation of hepatic gluconeogenesis through phosphorylation of CREB-
binding protein. NatMed 2004;10:633–637.
70. Gunton JE, Kulkarni RN, Yim S, Okada T, Hawthorne WJ, Tseng YH, Roberson RS, Ricordi C,
O’Connell PJ, Gonzalez FJ, Kahn CR. Loss of ARNT/HIF1beta mediates altered gene expression
NIH-PA Author Manuscript

and pancreatic-islet dysfunction in human type 2 diabetes. Cell 2005;122:337–349. [PubMed:


16096055]
71. Czech MP. ARNT misbehavin’ in diabetic beta cells. NatMed 2006;12:39–40.
72. Rhodes CJ. Type 2 diabetes-a matter of beta-cell life and death? Science 2005;307:380–384.
[PubMed: 15662003]
73. Donath MY, Halban PA. Decreased beta-cell mass in diabetes: significance, mechanisms and
therapeutic implications. Diabetologia 2004;47:581–589. [PubMed: 14767595]
74. Lingohr MK, Buettner R, Rhodes CJ. Pancreatic beta-cell growth and survival--a role in obesity-
linked type 2 diabetes? Trends MolMed 2002;8:375–384.
75. Unger RH, Zhou YT. Lipotoxicity of beta-cells in obesity and in other causes of fatty acid spillover.
Diabetes 2001;50 (Suppl 1):S118–S121. [PubMed: 11272168]
76. Poitout V, Robertson RP. Minireview: Secondary beta-cell failure in type 2 diabetes--a convergence
of glucotoxicity and lipotoxicity. Endocrinology 2002;143:339–342. [PubMed: 11796484]
77. Kaneto H, Nakatani Y, Kawamori D, Miyatsuka T, Matsuoka TA, Matsuhisa M, Yamasaki Y. Role
of oxidative stress, endoplasmic reticulum stress, and c-Jun N-terminal kinase in pancreatic beta-cell
dysfunction and insulin resistance. IntJBiochemCell Biol 2006;38:782–793.
78. Donath MY, Storling J, Maedler K, Mandrup-Poulsen T. Inflammatory mediators and islet beta-cell
failure: a link between type 1 and type 2 diabetes. JMolMed 2003;81:455–470.
NIH-PA Author Manuscript

79. Hennige AM, Burks DJ, Ozcan U, Kulkarni RN, Ye J, Park S, Schubert M, Fisher TL, Dow MA,
Leshan R, Zakaria M, Mossa-Basha M, White MF. Upregulation of insulin receptor substrate-2 in
pancreatic beta cells prevents diabetes. JClinInvest 2003;112:1521–1532.
80. Werner ED, Lee J, Hansen L, Yuan M, Shoelson SE. Insulin resistance due to phosphorylation of
insulin receptor substrate-1 at serine 302. J Biol Chem 2004;279:35298–35305. [PubMed:
15199052]%20
81. White MF. IRS proteins and the common path to diabetes. AmJPhysiol EndocrinolMetab
2002;283:E413–E422.
82. Herbert A, Gerry NP, McQueen MB, Heid IM, Pfeufer A, Illig T, Wichmann HE, Meitinger T, Hunter
D, Hu FB, Colditz G, Hinney A, Hebebrand J, Koberwitz K, Zhu X, Cooper R, Ardlie K, Lyon H,
Hirschhorn JN, Laird NM, Lenburg ME, Lange C, Christman MF. A common genetic variant is
associated with adult and childhood obesity. Science 2006;312:279–283. [PubMed: 16614226]
83. Hanis CL, Boerwinkle E, Chakraborty R, Ellsworth DL, Concannon P, Stirling B, Morrison VA,
Wapelhorst B, Spielman RS, Gogolin-Ewens KJ, Shephard JM, Williams SR, Risch N, Hinds D,
Iwasaki N, Ogata M, Ommoi Y, Petzold C, Rietzsch H, Schroeder H-E, Schulze J, Cox NJ, Menzel

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 20

S, Boriraj VV, Chen X, Lim LR, Linder T, Mereu LE, Wang Y-Q, Xiang K, Yamagata K, Yang Y,
Bell GI. A genome-wide search for human non-insulin-dependent (type 2) diabetes genes reveals a
major susceptibility locus on chromosome 2. NatGenet 1996;13:161–166.
NIH-PA Author Manuscript

84. Cox NJ, Frigge M, Nicolae DL, Concannon P, Hanis CL, Bell GI, Kong A. Loci on chromosomes 2
(NIDDM1) and 15 interact to increase susceptibility to diabetes in Mexican Americans. NatGenet
1999;21:213–215.
85. Elbein SC, Hoffman MD, Teng K, Leppert MF, Hasstedt SJ. A genome-wide search for type 2 diabetes
susceptibility genes in Utah Caucasians. Diabetes 1999;48:1175–1182. [PubMed: 10331426]
86. Luo TH, Zhao Y, Li G, Yuan WT, Zhao JJ, Chen JL, Huang W, Luo M. A genome-wide search for
type II diabetes susceptibility genes in Chinese Hans. Diabetologia 2001;44:501–506. [PubMed:
11357482]
87. Mori Y, Otabe S, Dina C, Yasuda K, Populaire C, Lecoeur C, Vatin V, Durand E, Hara K, Okada T,
Tobe K, Boutin P, Kadowaki T, Froguel P. Genome-wide search for type 2 diabetes in Japanese
affected sib-pairs confirms susceptibility genes on 3q, 15q, and 20q and identifies two new candidate
Loci on 7p and 11p. Diabetes 2002;51:1247–1255. [PubMed: 11916952]
88. Busfield F, Duffy DL, Kesting JB, Walker SM, Lovelock P, Good D, Tate H, Watego D, Marczak
M, Hayman N, Shaw JT. A Genomewide Search for Type 2 Diabetes-Susceptibility Genes in
Indigenous Australians. AmJHum Genet 2002;70:349–357.
89. Bowden DW, Sale M, Howard TD, Qadri A, Spray BJ, Rothschild CB, Akots G, Rich SS, Freedman
BI. Linkage of genetic markers on human chromosomes 20 and 12 to NIDDM in Caucasian sib pairs
with a history of diabetic nephropathy. Diabetes 1997;46:882–886. [PubMed: 9133559]
NIH-PA Author Manuscript

90. Hani, e; Zouali, H.; Philippi, A.; Beaudoin, JC.; Vionnet, N.; Passa, P.; Demenais, F.; Froguel, P.
Indication for genetic linkage of the phosphoenolpyruvate carboxykinase (PCK1) gene region on
chromosome 20q to non-insulin-dependent diabetes mellitus. Diabetes Metab 1996;22:451–454.
[PubMed: 8985654]
91. Zouali H, Hani EH, Philippi A, Vionnet N, Beckmann JS, Demenais F, Froguel P. A susceptibility
locus for early-onset non-insulin dependent (type 2) diabetes mellitus maps to chromosome 20q,
proximal to the phosphoenolpyruvate carboxykinase gene. Hum MolGenet 1997;6:1401–1408.
92. Ghosh S, Watanabe RM, Hauser ER, Valle T, Magnuson VL, Erdos MR, Langefeld CD, Balow J Jr,
Ally DS, Kohtamaki K, Chines P, Birznieks G, Kaleta HS, Musick A, Te C, Tannenbaum J, Eldridge
W, Shapiro S, Martin C, Witt A, So A, Chang J, Shurtleff B, Porter R, Boehnke M. Type 2 diabetes:
evidence for linkage on chromosome 20 in 716 Finnish affected sib pairs. ProcNatlAcadSciUSA
1999;96:2198–2203.
93. Permutt MA, Wasson JC, Suarez BK, Lin J, Thomas J, Meyer J, Lewitzky S, Rennich JS, Parker A,
Duprat L, Maruti S, Chayen S, Glaser B. A genome scan for type 2 diabetes susceptibility loci in a
genetically isolated population. Diabetes 2001;50:681–685. [PubMed: 11246891]
94. Rotimi CN, Chen G, Adeyemo AA, Furbert-Harris P, Guass D, Zhou J, Berg K, Adegoke O, Amoah
A, Owusu S, Acheampong J, Agyenim-Boateng K, Eghan BA Jr, Oli J, Okafor G, Ofoegbu E,
Osotimehin B, Abbiyesuku F, Johnson T, Rufus T, Fasanmade O, Kittles R, Daniel H, Chen Y,
Dunston G, Collins FS. A genome-wide search for type 2 diabetes susceptibility genes in West
NIH-PA Author Manuscript

Africans: the Africa America Diabetes Mellitus (AADM) Study. Diabetes 2004;53:838–841.
[PubMed: 14988271]
95. Ghosh S, Watanabe RM, Valle TT, Hauser ER, Magnuson VL, Langefeld CD, Ally DS, Mohlke KL,
Silander K, Kohtamaki K, Chines P, Balow JJ, Birznieks G, Chang J, Eldridge W, Erdos MR,
Karanjawala ZE, Knapp JI, Kudelko K, Martin C, Morales-Mena A, Musick A, Musick T, Pfahl C,
Porter R, Rayman JB, Rha D, Segal L, Shapiro S, Sharaf R, Shurtleff B, So A, Tannenbaum J, Te C,
Tovar J, Unni A, Welch C, Whiten R, Witt A, Blaschak-Harvan J, Douglas JA, Duren WL, Epstein
MP, Fingerlin TE, Kaleta HS, Lange EM, Li C, McEachin RC, Stringham HM, Trager E, White PP,
Eriksson J, Toivanen L, Vidgren G, Nylund SJ, Tuomilehto-Wolf E, Ross EH, Demirchyan E,
Hagopian WA, Buchanan TA, Tuomilehto J, Bergman RN, Collins FS, Boehnke M. The Finland-
United States Investigation of Non-Insulin-Dependent Diabetes Mellitus Genetics (FUSION) Study.
I An Autosomal Genome Scan for Genes That Predispose to Type 2 Diabetes. AmJHumGenet
2000;67:1174–1185.
96. Mahtani MM, Widen E, Lehto M, Thomas J, McCarthy M, Brayer J, Bryant B, Chan G, Daly M,
Forsblom C, Kanninen T, Kirby A, Kruglyak L, Munnelly K, Parkkonen M, Reeve-Daly MP, Weaver

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 21

A, Brettin T, Duyk G, Lander ES, Groop LC. Mapping of a gene for type 2 diabetes associated with
an insulin secretion defect by a genome scan in Finnish families. NatGenet 1996;14:90–94.
97. Lindgren CM, Mahtani MM, Widen E, McCarthy MI, Daly MJ, Kirby A, Reeve MP, Kruglyak L,
NIH-PA Author Manuscript

Parker A, Meyer J, Almgren P, Lehto M, Kanninen T, Tuomi T, Groop LC, Lander ES. Genomewide
search for type 2 diabetes mellitus susceptibility Loci in finnish families: the botnia study.
AmJHumGenet 2002;70:509–516.
98. Ehm MG, Karnoub MC, Sakul H, Gottschalk K, Holt DC, Weber JL, Vaske D, Briley D, Briley L,
Kopf J, McMillen P, Nguyen Q, Reisman M, Lai EH, Joslyn G, Shepherd NS, Bell C, Wagner MJ,
Burns DK. Genomewide Search for Type 2 Diabetes Susceptibility Genes in Four American
Populations. AmJHumGenet 2000;66:1871–1881.
99. Bektas A, Suprenant ME, Wogan LT, Plengvidhya N, Rich SS, Warram JH, Krolewski AS, Doria A.
Evidence of a novel type 2 diabetes locus 50 cM centromeric to NIDDM2 on chromosome 12q.
Diabetes 1999;48:2246–2251. [PubMed: 10535461]
100. Bektas A, Hughes JN, Warram JH, Krolewski AS, Doria A. Type 2 diabetes locus on 12q15. Further
mapping and mutation screening of two candidate genes. Diabetes 2001;50:204–208. [PubMed:
11147789]
101. Shaw JT, Lovelock PK, Kesting JB, Cardinal J, Duffy D, Wainwright B, Cameron DP. Novel
susceptibility gene for late-onset NIDDM is localized to human chromosome 12q. Diabetes
1998;47:1793–1796. [PubMed: 9792550]
102. Wiltshire S, Frayling TM, Groves CJ, Levy JC, Hitman GA, Sampson M, Walker M, Menzel S,
Hattersley AT, Cardon LR, McCarthy MI. Evidence from a large U.K. family collection that genes
influencing age of onset of type 2 diabetes map to chromosome 12p and to the MODY3/NIDDM2
NIH-PA Author Manuscript

locus on 12q24. Diabetes 2004;53:855–860. [PubMed: 14988275]


103. Hanson RL, Ehm MG, Pettitt DJ, Prochazka M, Thompson DB, Timberlake D, Foroud T, Kobes S,
Baier L, Burns DK, Almasy L, Blangero J, Garvey WT, Bennett PH, Knowler WC. An autosomal
genomic scan for loci linked to type II diabetes mellitus and body-mass index in Pima Indians.
AmJHumGenet 1998;63:1130–1138.
104. Hsueh WC, St Jean PL, Mitchell BD, Pollin TI, Knowler WC, Ehm MG, Bell CJ, Sakul H, Wagner
MJ, Burns DK, Shuldiner AR. Genome-Wide and Fine-Mapping Linkage Studies of Type 2
Diabetes and Glucose Traits in the Old Order Amish: Evidence for a New Diabetes Locus on
Chromosome 14q11 and Confirmation of a Locus on Chromosome 1q21–q24. Diabetes
2003;52:550–557. [PubMed: 12540634]
105. Vionnet N, Hani EH, Dupont S, Gallina S, Francke S, Dotte S, De Matos F, Durand E, Lepretre F,
Lecoeur C, Gallina P, Zekiri L, Dina C, Froguel P. Genomewide Search for Type 2 Diabetes-
Susceptibility Genes in French Whites: Evidence for a Novel Susceptibility Locus for Early-Onset
Diabetes on Chromosome 3q27-qter and Independent Replication of a Type 2-Diabetes Locus on
Chromosome 1q21–q24. AmJHumGenet 2000;67:1470–1480.
106. Wiltshire S, Hattersley AT, Hitman GA, Walker M, Levy JC, Sampson M, O’Rahilly S, Frayling
TM, Bell JI, Lathrop GM, Bennett A, Dhillon R, Fletcher C, Groves CJ, Jones E, Prestwich P,
Simecek N, Rao PV, Wishart M, Foxon R, Howell S, Smedley D, Cardon LR, Menzel S, McCarthy
NIH-PA Author Manuscript

MI. A Genomewide Scan for Loci Predisposing to Type 2 Diabetes in a U.K. Population (The
Diabetes UK Warren 2 Repository): Analysis Of 573 Pedigrees Provides Independent Replication
of a Susceptibility Locus on Chromosome 1q. Am J Hum Genet 2001;69:553–569. [PubMed:
11484155]
107. Meigs JB, Panhuysen CI, Myers RH, Wilson PW, Cupples LA. A genome-wide scan for loci linked
to plasma levels of glucose and HbA(1c) in a community-based sample of Caucasian pedigrees:
The Framingham Offspring Study. Diabetes 2002;51:833–840. [PubMed: 11872688]
108. Ng MC, So WY, Cox NJ, Lam VK, Cockram CS, Critchley JA, Bell GI, Chan JC. Genome-wide
scan for type 2 diabetes loci in Hong Kong Chinese and confirmation of a susceptibility locus on
chromosome 1q21–q25. Diabetes 2004;53:1609–1613. [PubMed: 15161769]
109. Du W, Sun H, Wang H, Qiang B, Shen Y, Yao Z, Gu J, Xiong M, Huang W, Chen Z, Zuo J, Hua
X, Gao W, Sun Q, Fang F. Confirmation of susceptibility gene loci on chromosome 1 in northern
China Han families with type 2 diabetes. Chin MedJ(Engl) 2001;114:876–878.
110. McCarthy MI. Growing evidence for diabetes susceptibility genes from genome scan data.
CurrDiabRep 2003;3:159–167.

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 22

111. Das SK, Hasstedt SJ, Zhang Z, Elbein SC. Linkage and Association Mapping of a Chromosome
1q21–q24 Type 2 Diabetes Susceptibility Locus in Northern European Caucasians. Diabetes
2004;53:492–499. [PubMed: 14747303]
NIH-PA Author Manuscript

112. Wang H, Chu W, Das SK, Ren Q, Hasstedt SJ, Elbein SC. Liver pyruvate kinase polymorphisms
are associated with type 2 diabetes in northern European Caucasians. Diabetes 2002;51:2861–2865.
[PubMed: 12196482]
113. Das SK, Chu W, Zhang Z, Hasstedt SJ, Elbein SC. Calsquestrin 1 (CASQ1) gene polymorphisms
under chromosome 1q21 linkage peak are associated with type 2 diabetes in Northern European
Caucasians. Diabetes 2004;53:3300–3306. [PubMed: 15561963]
114. Fu M, Damcott CM, Sabra M, Pollin TI, Ott SH, Wang J, Garant MJ, O’Connell JR, Mitchell BD,
Shuldiner AR. Polymorphism in the Calsequestrin 1 (CASQ1) Gene on Chromosome 1q21 Is
Associated With Type 2 Diabetes in the Old Order Amish. Diabetes 2004;53:3292–3299. [PubMed:
15561962]
115. Parker A, Meyer J, Lewitzky S, Rennich JS, Chan G, Thomas JD, Orho-Melander M, Lehtovirta M,
Forsblom C, Hyrkko A, Carlsson M, Lindgren C, Groop LC. A gene conferring susceptibility to
type 2 diabetes in conjunction with obesity is located on chromosome 18p11. Diabetes
2001;50:675–680. [PubMed: 11246890]
116. Aulchenko YS, Vaessen N, Heutink P, Pullen J, Snijders PJ, Hofman A, Sandkuijl LA, Houwing-
Duistermaat JJ, Edwards M, Bennett S, Oostra BA, van Duijn CM. A genome-wide search for genes
involved in type 2 diabetes in a recently genetically isolated population from the Netherlands.
Diabetes 2003;52:3001–3004. [PubMed: 14633863]
117. van Tilburg JH, Sandkuijl LA, Strengman E, van Someren H, Rigters-Aris CA, Pearson PL, van
NIH-PA Author Manuscript

Haeften TW, Wijmenga C. A genome-wide scan in type 2 diabetes mellitus provides independent
replication of a susceptibility locus on 18p11 and suggests the existence of novel Loci on 2q12 and
19q13. JClinEndocrinolMetab 2003;88:2223–2230.
118. Francke S, Manraj M, Lacquemant C, Lecoeur C, Lepretre F, Passa P, Hebe A, Corset L, Yan SL,
Lahmidi S, Jankee S, Gunness TK, Ramjuttun US, Balgobin V, Dina C, Froguel P. A genome-wide
scan for coronary heart disease suggests in Indo- Mauritians a susceptibility locus on chromosome
16p13 and replicates linkage with the metabolic syndrome on 3q27. HumMolGenet 2001;10:2751–
2765.
119. Duggirala R, Almasy L, Blangero J, Jenkinson CP, Arya R, DeFronzo RA, Stern MP, O’Connell P.
Further evidence for a type 2 diabetes susceptibility locus on chromosome 11q. GenetEpidemiol
2003;24:240–242.
120. van Tilburg JH, Sandkuijl LA, Franke L, Strengman E, Pearson PL, van Haeften TW, Wijmenga C.
Genome-wide screen in obese pedigrees with type 2 diabetes mellitus from a defined Dutch
population. EurJClinInvest 2003;33:1070–1074.
121. Horikawa Y, Oda N, Cox NJ, Li X, Orho-Melander M, Hara M, Hinokio Y, Lindner TH, Mashima
H, Schwarz PE, Bosque-Plata L, Horikawa Y, Oda Y, Yoshiuchi I, Colilla S, Polonsky KS, Wei S,
Concannon P, Iwasaki N, Schulze J, Baier LJ, Bogardus C, Groop L, Boerwinkle E, Hanis CL, Bell
GI. Genetic variation in the gene encoding calpain-10 is associated with type 2 diabetes mellitus.
NIH-PA Author Manuscript

NatGenet 2000;26:163–175.
122. Baier LJ, Permana PA, Yang X, Pratley RE, Hanson RL, Shen GQ, Mott D, Knowler WC, Cox NJ,
Horikawa Y, Oda N, Bell GI, Bogardus C. A calpain-10 gene polymorphism is associated with
reduced muscle mRNA levels and insulin resistance. JClinInvest 2000;106:R69–R73.
123. Cox NJ, Hayes MG, Roe CA, Tsuchiya T, Bell GI. Linkage of calpain 10 to type 2 diabetes: the
biological rationale. Diabetes 2004;53(Suppl 1):S19–25. [PubMed: 14749261]S19–S25
124. Weedon MN, Schwarz PE, Horikawa Y, Iwasaki N, Illig T, Holle R, Rathmann W, Selisko T, Schulze
J, Owen KR, Evans J, Bosque-Plata L, Hitman G, Walker M, Levy JC, Sampson M, Bell GI,
McCarthy MI, Hattersley AT, Frayling TM. Meta-analysis and a large association study confirm a
role for calpain-10 variation in type 2 diabetes susceptibility. AmJHumGenet 2003;73:1208–1212.
125. Song Y, Niu T, Manson JE, Kwiatkowski DJ, Liu S. Are variants in the CAPN10 gene related to
risk of type 2 diabetes? A quantitative assessment of population and family-based association
studies. AmJHumGenet 2004;74:208–222.
126. Elbein SC, Chu W, Ren Q, Hemphill C, Schay J, Cox NJ, Hanis CL, Hasstedt SJ. Role of calpain-10
gene variants in familial type 2 diabetes in Caucasians. JClinEndocrinolMetab 2002;87:650–654.

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 23

127. Orho-Melander M, Klannemark M, Svensson MK, Ridderstrale M, Lindgren CM, Groop L. Variants
in the calpain-10 gene predispose to insulin resistance and elevated free fatty acid levels. Diabetes
2002;51:2658–2664. [PubMed: 12145185]
NIH-PA Author Manuscript

128. Lyssenko V, Almgren P, Anevski D, Orho-Melander M, Sjogren M, Saloranta C, Tuomi T, Groop


L. Botnia Study Group. Genetic Prediction of Future Type 2 Diabetes. PLOS Medicine
2005;2:1299–1308.
129. Silander K, Mohlke KL, Scott LJ, Peck EC, Hollstein P, Skol AD, Jackson AU, Deloukas P, Hunt
S, Stavrides G, Chines PS, Erdos MR, Narisu N, Conneely KN, Li C, Fingerlin TE, Dhanjal SK,
Valle TT, Bergman RN, Tuomilehto J, Watanabe RM, Boehnke M, Collins FS. Genetic Variation
Near the Hepatocyte Nuclear Factor-4alpha Gene Predicts Susceptibility to Type 2 Diabetes.
Diabetes 2004;53:1141–1149. [PubMed: 15047633]
130. Love-Gregory LD, Wasson J, Ma J, Jin CH, Glaser B, Suarez BK, Permutt MA. A Common
Polymorphism in the Upstream Promoter Region of the Hepatocyte Nuclear Factor-4alpha 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 2004;53:1134–1140. [PubMed:
15047632]
131. Vaxillaire M, Dina C, Lobbens S, Dechaume A, Vasseur-Delannoy V, Helbecque N, Charpentier
G, Froguel P. Effect of common polymorphisms in the HNF4alpha promoter on susceptibility to
type 2 diabetes in the French Caucasian population. Diabetologia 2005;48:440–444. [PubMed:
15735892]
132. Hansen SK, Rose CS, Glumer C, Drivsholm T, Borch-Johnsen K, Jorgensen T, Pedersen O, Hansen
T. Variation near the hepatocyte nuclear factor (HNF)-4alpha gene associates with type 2 diabetes
NIH-PA Author Manuscript

in the Danish population. Diabetologia 2005;48:452–458. [PubMed: 15735891]


133. Weedon MN, Owen KR, Shields B, Hitman G, Walker M, McCarthy MI, Love-Gregory LD, Permutt
MA, Hattersley AT, Frayling TM. Common variants of the hepatocyte nuclear factor-4alpha P2
promoter are associated with type 2 diabetes in the U.K. population. Diabetes 2004;53:3002–3006.
[PubMed: 15504983]
134. Damcott CM, Hoppman N, Ott SH, Reinhart LJ, Wang J, Pollin TI, O’Connell JR, Mitchell BD,
Shuldiner AR. Polymorphisms in both promoters of hepatocyte nuclear factor 4-alpha are associated
with type 2 diabetes in the Amish. Diabetes 2004;53:3337–3341. [PubMed: 15561969]
135. Muller YL, Infante AM, Hanson RL, Love-Gregory L, Knowler W, Bogardus C, Baier LJ. Variants
in hepatocyte nuclear factor 4alpha are modestly associated with type 2 diabetes in Pima Indians.
Diabetes 2005;54:3035–3039. [PubMed: 16186411]
136. Winckler W, Graham RR, de Bakker PI, Sun M, Almgren P, Tuomi T, Gaudet D, Hudson TJ, Ardlie
KG, Daly MJ, Hirschhorn JN, Groop L, Altshuler D. Association testing of variants in the hepatocyte
nuclear factor 4alpha gene with risk of type 2 diabetes in 7,883 people. Diabetes 2005;54:886–892.
[PubMed: 15734869]
137. Bento JL, Palmer ND, Mychaleckyj JC, Lange LA, Langefeld CD, Rich SS, Freedman BI, Bowden
DW. Association of protein tyrosine phosphatase 1B gene polymorphisms with type 2 diabetes.
Diabetes 2004;53:3007–3012. [PubMed: 15504984]
NIH-PA Author Manuscript

138. Palmer ND, Bento JL, Mychaleckyj JC, Langefeld CD, Campbell JK, Norris JM, Haffner SM,
Bergman RN, Bowden DW. Association of protein tyrosine phosphatase 1B gene polymorphisms
with measures of glucose homeostasis in Hispanic Americans: the insulin resistance atherosclerosis
study (IRAS) family study. Diabetes 2004;53:3013–3019. [PubMed: 15504985]
139. Florez JC, Agapakis CM, Burtt NP, Sun M, Almgren P, Rastam L, Tuomi T, Gaudet D, Hudson TJ,
Daly MJ, Ardlie KG, Hirschhorn JN, Groop L, Altshuler D. Association testing of the protein
tyrosine phosphatase 1B gene (PTPN1) with type 2 diabetes in 7,883 people. Diabetes
2005;54:1884–1891. [PubMed: 15919813]
140. Fu M, Sabra M, Damcott C, Pollini TI, Ott S, Tanner K, Wang J, Shi X, Connell JM, Mitchell BD,
Shuldiner AR: Strong association of intragenic SNPs with type 2 diabetes in the 152 megabase
region of chromosome 1q21 in the Old Order Amish (Abstract). American Society of Human
Genetics 54th Meeting Abstracts, 2004
141. Zeggini E, Rayner W, Groves CJ, Hanson RL, Mitchell BD, Osapos J, Connell MV, Jia W, Ng
MCY, Knowler WC, Baier LJ, Froguel P, Xiang K, Chan JCN, Deloukas P, Cardon L, Bogardus
C, Elbein SC, Shuldiner A, McCarthy M. Meta-Analysis of 3000 Single Nucleotide Polymorphisms

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 24

from Chromosome 1q in Samples from Seven Linked Populations Reveals Shared Type 2 Diabetes
Susceptibility Variants (Abstract). Diabetes 2005;54(Suppl 2)
142. Gibson F, Froguel P. Genetics of the APM1 locus and its contribution to type 2 diabetes susceptibility
NIH-PA Author Manuscript

in French Caucasians. Diabetes 2004;53:2977–2983. [PubMed: 15504979]


143. Zacharova J, Chiasson JL, Laakso M. The common polymorphisms (single nucleotide
polymorphism [SNP] +45 and SNP +276) of the adiponectin gene predict the conversion from
impaired glucose tolerance to type 2 diabetes: the STOP-NIDDM trial. Diabetes 2005;54:893–899.
[PubMed: 15734870]
144. Vozarova, de Court; Hanson, RL.; Funahashi, T.; Lindsay, RS.; Matsuzawa, Y.; Tanaka, S.;
Thameem, F.; Gruber, JD.; Froguel, P.; Wolford, JK. Common Polymorphisms in the Adiponectin
Gene ACDC Are Not Associated With Diabetes in Pima Indians. Diabetes 2005;54:284–289.
[PubMed: 15616040]
145. Grant SF, Thorleifsson G, Reynisdottir I, Benediktsson R, Manolescu A, Sainz J, Helgason A,
Stefansson H, Emilsson V, Helgadottir A, Styrkarsdottir U, Magnusson KP, Walters GB, Palsdottir
E, Jonsdottir T, Gudmundsdottir T, Gylfason A, Saemundsdottir J, Wilensky RL, Reilly MP, Rader
DJ, Bagger Y, Christiansen C, Gudnason V, Sigurdsson G, Thorsteinsdottir U, Gulcher JR, Kong
A, Stefansson K. Variant of transcription factor 7-like 2 (TCF7L2) gene confers risk of type 2
diabetes. NatGenet 2006;38:320–323.
146. Altshuler D, Hirschhorn JN, Klannemark M, Lindgren CM, Vohl MC, Nemesh J, Lane CR, Schaffner
SF, Bolk S, Brewer C, Tuomi T, Gaudet D, Hudson TJ, Daly M, Groop L, Lander ES. The common
PPARgamma Pro12Ala polymorphism is associated with decreased risk of type 2 diabetes.
NatGenet 2000;26:76–80.
NIH-PA Author Manuscript

147. Memisoglu A, Hu FB, Hankinson SE, Liu S, Meigs JB, Altshuler DM, Hunter DJ, Manson JE.
Prospective study of the association between the proline to alanine codon 12 polymorphism in the
PPARgamma gene and type 2 diabetes. Diabetes Care 2003;26:2915–2917. [PubMed: 14514601]
148. Inoue H, Ferrer J, Welling CM, Elbein SC, Hoffman M, Mayorga R, Warren-Perry M, Zhang Y,
Millns H, Turner R, Province M, Bryan J, Permutt MA, Aguilar-Bryan L. Sequence variants in the
sulfonylurea receptor (SUR) gene are associated with NIDDM in Caucasians. Diabetes
1996;45:825–831. [PubMed: 8635661]
149. Hart LM, de Knijff P, Dekker JM, Stolk RP, Nijpels G, van der Does FE, Ruige JB, Grobbee DE,
Heine RJ, Maassen JA. Variants in the sulphonylurea receptor gene: association of the exon 16-3t
variant with Type II diabetes mellitus in Dutch Caucasians. Diabetologia 1999;42:617–620.
[PubMed: 10333056]
150. Meirhaeghe A, Helbecque N, Cottel D, Arveiler D, Ruidavets JB, Haas B, Ferrieres J, Tauber JP,
Bingham A, Amouyel P. Impact of sulfonylurea receptor 1 genetic variability on non-insulin-
dependent diabetes mellitus prevalence and treatment: a population study. AmJMedGenet
2001;101:4–8.
151. Barroso I, Luan J, Middelberg RP, Harding AH, Franks PW, Jakes RW, Clayton D, Schafer AJ,
O’Rahilly S, Wareham NJ. Candidate Gene Association Study in Type 2 Diabetes Indicates a Role
for Genes Involved in beta-Cell Function as Well as Insulin Action. PLoSBiol 2003;1:E20.
NIH-PA Author Manuscript

152. Hani EH, Hager J, Philippi A, Demenais F, Froguel P, Vionnet N. Mapping NIDDM susceptibility
loci in French families: studies with markers in the region of NIDDM1 on chromosome 2q. Diabetes
1997;46:1225–1226. [PubMed: 9200659]
153. Gloyn AL, Weedon MN, Owen KR, Turner MJ, Knight BA, Hitman G, Walker M, Levy JC,
Sampson M, Halford S, McCarthy MI, Hattersley AT, Frayling TM. 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
2003;52:568–572. [PubMed: 12540637]
154. Florez JC, Burtt N, de Bakker PI, Almgren P, Tuomi T, Holmkvist J, Gaudet D, Hudson TJ, Schaffner
SF, Daly MJ, Hirschhorn JN, Groop L, Altshuler D. Haplotype structure and genotype-phenotype
correlations of the sulfonylurea receptor and the islet ATP-sensitive potassium channel gene region.
Diabetes 2004;53:1360–1368. [PubMed: 15111507]
155. Elbein SC, Sun J, Scroggin E, Teng K, Hasstedt SJ. Role of common sequence variants in insulin
secretion in familial type 2 diabetic kindreds: the sulfonylurea receptor, glucokinase, and hepatocyte
nuclear factor 1alpha genes. Diabetes Care 2001;24:472–478. [PubMed: 11289470]

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 25

156. Hansen T, Echwald SM, Hansen L, Moller AM, Almind K, Clausen JO, Urhammer SA, Inoue H,
Ferrer J, Bryan J, Aguilar-Bryan L, Permutt MA, Pedersen O. Decreased tolbutamide-stimulated
insulin secretion in healthy subjects with sequence variants in the high-affinity sulfonylurea receptor
NIH-PA Author Manuscript

gene. Diabetes 1998;47:598–605. [PubMed: 9568693]


157. Inoue H, Ferrer J, Warren-Perry M, Zhang Y, Millns H, Turner RC, Elbein SC, Hampe CL, Suarez
BK, Inagaki N, Seino S, Permutt MA. Sequence variants in the pancreatic islet beta-cell inwardly
rectifying K+ channel Kir6.2 (Bir) gene: identification and lack of role in Caucasian patients with
NIDDM. Diabetes 1997;46:502–507. [PubMed: 9032109]
158. Sakura H, Wat N, Horton V, Millns H, Turner RC, Ashcroft FM. Sequence variations in the human
Kir6.2 gene, a subunit of the beta- cell ATP-sensitive K-channel: no association with NIDDM in
while Caucasian subjects or evidence of abnormal function when expressed in vitro. Diabetologia
1996;39:1233–1236. [PubMed: 8897013]
159. 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
2002;51:875–879. [PubMed: 11872696]
160. Nielsen EM, Hansen L, Carstensen B, Echwald SM, Drivsholm T, Glumer C, Thorsteinsson B,
Borch-Johnsen K, Hansen T, Pedersen O. The E23K variant of Kir6.2 associates with impaired
post-OGTT serum insulin response and increased risk of type 2 diabetes. Diabetes 2003;52:573–
577. [PubMed: 12540638]
161. ‘t Hart LM, van Haeften TW, Dekker JM, Bot M, Heine RJ, Maassen JA. 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 2002;51:3135–3138. [PubMed: 12351459]
NIH-PA Author Manuscript

162. Elbein S, Rotwein P, Permutt MA, Bell GI, Sanz N, Karam JH. Lack of association of the
polymorphic locus in the 5′-flanking region of the human insulin gene and diabetes in American
blacks. Diabetes 1985;34:433–439. [PubMed: 3886460]
163. Le Stunff C, Fallin D, Schork NJ, Bougneres P. The insulin gene VNTR is associated with fasting
insulin levels and development of juvenile obesity. NatGenet 2000;26:444–446.
164. Dunger DB, Ong KK, Huxtable SJ, Sherriff A, Woods KA, Ahmed ML, Golding J, Pembrey ME,
Ring S, Bennett ST, Todd JA. Association of the INS VNTR with size at birth. ALSPAC Study
Team Avon Longitudinal Study of Pregnancy and Childhood. NatGenet 1998;19:98–100.
165. Lindsay RS, Hanson RL, Wiedrich C, Knowler WC, Bennett PH, Baier LJ. The insulin gene variable
number tandem repeat class I/III polymorphism is in linkage disequilibrium with birth weight but
not Type 2 diabetes in the Pima population. Diabetes 2003;52:187–193. [PubMed: 12502511]
166. Mitchell SM, Hattersley AT, Knight B, Turner T, Metcalf BS, Voss LD, Davies D, McCarthy A,
Wilkin TJ, Smith GD, Ben Shlomo Y, Frayling TM. Lack of support for a role of the insulin gene
variable number of tandem repeats minisatellite (INS-VNTR) locus in fetal growth or type 2
diabetes-related intermediate traits in United Kingdom populations. JClinEndocrinolMetab
2004;89:310–317.
167. Huxtable SJ, Saker PJ, Haddad L, Walker M, Frayling TM, Levy JC, Hitman GA, O’Rahilly S,
Hattersley AT, McCarthy MI. Analysis of parent-offspring trios provides evidence for linkage and
NIH-PA Author Manuscript

association between the insulin gene and type 2 diabetes mediated exclusively through paternally
transmitted class III variable number tandem repeat alleles. Diabetes 2000;49:126–130. [PubMed:
10615960]
168. Almind K, Bjorbaek C, Vestergaard H, Hansen T, Echwald S, Pedersen O. Aminoacid
polymorphisms of insulin receptor substrate-1 in non- insulin- dependent diabetes mellitus. Lancet
1993;342:828–832. [PubMed: 8104271]
169. Almind K, Inoue G, Pedersen O, Kahn CR. A common amino acid polymorphism in insulin receptor
substrate-1 causes impaired insulin signaling. Evidence from transfection studies. JClinInvest
1996;97:2569–2575.
170. Zeggini E, Parkinson J, Halford S, Owen KR, Frayling TM, Walker M, Hitman GA, Levy JC,
Sampson MJ, Feskens EJ, Hattersley AT, McCarthy MI. Association studies of insulin receptor
substrate 1 gene (IRS1) variants in type 2 diabetes samples enriched for family history and early
age of onset. Diabetes 2004;53:3319–3322. [PubMed: 15561966]
171. Florez JC, Sjogren M, Burtt N, Orho-Melander M, Schayer S, Sun M, Almgren P, Lindblad U,
Tuomi T, Gaudet D, Hudson TJ, Daly MJ, Ardlie KG, Hirschhorn JN, Altshuler D, Groop L.

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 26

Association testing in 9,000 people fails to confirm the association of the insulin receptor substrate-1
G972R polymorphism with type 2 diabetes. Diabetes 2004;53:3313–3318. [PubMed: 15561965]
172. Hara K, Tobe K, Okada T, Kadowaki H, Akanuma Y, Ito C, Kimura S, Kadowaki T. A genetic
NIH-PA Author Manuscript

variation in the PGC-1 gene could confer insulin resistance and susceptibility to Type II diabetes.
Diabetologia 2002;45:740–743. [PubMed: 12107756]
173. Ek J, Andersen G, Urhammer SA, Gaede PH, Drivsholm T, Borch-Johnsen K, Hansen T, Pedersen
O. Mutation analysis of peroxisome proliferator-activated receptor-gamma coactivator-1 (PGC-1)
and relationships of identified amino acid polymorphisms to Type II diabetes mellitus. Diabetologia
2001;44:2220–2226. [PubMed: 11793024]
174. Vimaleswaran KS, Radha V, Ghosh S, Majumder PP, Deepa R, Babu HN, Rao MR, Mohan V.
Peroxisome proliferator-activated receptor-gamma co-activator-1alpha (PGC-1alpha) gene
polymorphisms and their relationship to Type 2 diabetes in Asian Indians. DiabetMed
2005;22:1516–1521.
175. Kim JH, Shin HD, Park BL, Cho YM, Kim SY, Lee HK, Park KS. Peroxisome proliferator-activated
receptor gamma coactivator 1 alpha promoter polymorphisms are associated with early-onset type
2 diabetes mellitus in the Korean population. Diabetologia 2005;48:1323–1330. [PubMed:
15937669]
176. Fanelli M, Filippi E, Sentinelli F, Romeo S, Fallarino M, Buzzetti R, Leonetti F, Baroni MG. The
Gly482Ser missense mutation of the Peroxisome Proliferator-activated receptor gamma
coactivator-1alpha (PGC-1alpha) gene associates with reduced insulin sensitivity in normal and
glucose-intolerant obese subjects. DisMarkers 2005;21:175–180.
177. Pihlajamaki J, Kinnunen M, Ruotsalainen E, Salmenniemi U, Vauhkonen I, Kuulasmaa T,
NIH-PA Author Manuscript

Kainulainen S, Laakso M. Haplotypes of PPARGC1A are associated with glucose tolerance, body
mass index and insulin sensitivity in offspring of patients with type 2 diabetes. Diabetologia
2005;48:1331–1334. [PubMed: 15912394]
178. Muller YL, Bogardus C, Pedersen O, Baier L. A Gly482Ser missense mutation in the peroxisome
proliferator-activated receptor gamma coactivator-1 is associated with altered lipid oxidation and
early insulin secretion in Pima Indians. Diabetes 2003;52:895–898. [PubMed: 12606537]
179. Ambye L, Rasmussen S, Fenger M, Jorgensen T, Borch-Johnsen K, Madsbad S, Urhammer SA.
Studies of the Gly482Ser polymorphism of the peroxisome proliferator-activated receptor gamma
coactivator 1alpha (PGC-1alpha) gene in Danish subjects with the metabolic syndrome. Diabetes
ResClinPract 2005;67:175–179.
180. Meyre D, Bouatia-Naji N, Tounian A, Samson C, Lecoeur C, Vatin V, Ghoussaini M, Wachter C,
Hercberg S, Charpentier G, Patsch W, Pattou F, Charles MA, Tounian P, Clement K, Jouret B,
Weill J, Maddux BA, Goldfine ID, Walley A, Boutin P, Dina C, Froguel P. Variants of ENPP1 are
associated with childhood and adult obesity and increase the risk of glucose intolerance and type 2
diabetes. NatGenet 2005;37:863–867.
181. Matsuoka N, Patki A, Tiwari HK, Allison DB, Johnson SB, Gregersen PK, Leibel RL, Chung WK:
Association of K121Q polymorphism in ENPP1 (PC-1) with BMI in Caucasian and African-
American adults. Int.J.Obes.(Lond). .: 2005
NIH-PA Author Manuscript

182. Bacci S, Ludovico O, Prudente S, Zhang YY, Di Paola R, Mangiacotti D, Rauseo A, Nolan D, Duffy
J, Fini G, Salvemini L, Amico C, Vigna C, Pellegrini F, Menzaghi C, Doria A, Trischitta V. The
K121Q polymorphism of the ENPP1/PC-1 gene is associated with insulin resistance/atherogenic
phenotypes, including earlier onset of type 2 diabetes and myocardial infarction. Diabetes
2005;54:3021–3025. [PubMed: 16186408]
183. Abate N, Chandalia M, Satija P, Adams-Huet B, Grundy SM, Sandeep S, Radha V, Deepa R, Mohan
V. ENPP1/PC-1 K121Q polymorphism and genetic susceptibility to type 2 diabetes. Diabetes
2005;54:1207–1213. [PubMed: 15793263]
184. Ioannidis JP, Ntzani EE, Trikalinos TA. ‘Racial’ differences in genetic effects for complex diseases.
NatGenet 2004;36:1312–1318.
185. Beck-Nielsen H, Vaag A, Poulsen P, Gaster M. Metabolic and genetic influence on glucose
metabolism in type 2 diabetic subjects--experiences from relatives and twin studies.
BestPractResClinEndocrinolMetab 2003;17:445–467.

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 27

186. Kapranov P, Drenkow J, Cheng J, Long J, Helt G, Dike S, Gingeras TR. Examples of the complex
architecture of the human transcriptome revealed by RACE and high-density tiling arrays. Genome
Res 2005;15:987–997. [PubMed: 15998911]
NIH-PA Author Manuscript

187. Carninci P, Kasukawa T, Katayama S, Gough J, Frith MC, Maeda N, Oyama R, Ravasi T, Lenhard
B, Wells C, Kodzius R, Shimokawa K, Bajic VB, Brenner SE, Batalov S, Forrest AR, Zavolan M,
Davis MJ, Wilming LG, Aidinis V, Allen JE, Ambesi-Impiombato A, Apweiler R, Aturaliya RN,
Bailey TL, Bansal M, Baxter L, Beisel KW, Bersano T, Bono H, Chalk AM, Chiu KP, Choudhary
V, Christoffels A, Clutterbuck DR, Crowe ML, Dalla E, Dalrymple BP, de Bono B, Della GG, di
Bernardo D, Down T, Engstrom P, Fagiolini M, Faulkner G, Fletcher CF, Fukushima T, Furuno M,
Futaki S, Gariboldi M, Georgii-Hemming P, Gingeras TR, Gojobori T, Green RE, Gustincich S,
Harbers M, Hayashi Y, Hensch TK, Hirokawa N, Hill D, Huminiecki L, Iacono M, Ikeo K, Iwama
A, Ishikawa T, Jakt M, Kanapin A, Katoh M, Kawasawa Y, Kelso J, Kitamura H, Kitano H, Kollias
G, Krishnan SP, Kruger A, Kummerfeld SK, Kurochkin IV, Lareau LF, Lazarevic D, Lipovich L,
Liu J, Liuni S, McWilliam S, Madan BM, Madera M, Marchionni L, Matsuda H, Matsuzawa S,
Miki H, Mignone F, Miyake S, Morris K, Mottagui-Tabar S, Mulder N, Nakano N, Nakauchi H,
Ng P, Nilsson R, Nishiguchi S, Nishikawa S, Nori F, Ohara O, Okazaki Y, Orlando V, Pang KC,
Pavan WJ, Pavesi G, Pesole G, Petrovsky N, Piazza S, Reed J, Reid JF, Ring BZ, Ringwald M,
Rost B, Ruan Y, Salzberg SL, Sandelin A, Schneider C, Schonbach C, Sekiguchi K, Semple CA,
Seno S, Sessa L, Sheng Y, Shibata Y, Shimada H, Shimada K, Silva D, Sinclair B, Sperling S,
Stupka E, Sugiura K, Sultana R, Takenaka Y, Taki K, Tammoja K, Tan SL, Tang S, Taylor MS,
Tegner J, Teichmann SA, Ueda HR, van Nimwegen E, Verardo R, Wei CL, Yagi K, Yamanishi H,
Zabarovsky E, Zhu S, Zimmer A, Hide W, Bult C, Grimmond SM, Teasdale RD, Liu ET, Brusic
V, Quackenbush J, Wahlestedt C, Mattick JS, Hume DA, Kai C, Sasaki D, Tomaru Y, Fukuda S,
NIH-PA Author Manuscript

Kanamori-Katayama M, Suzuki M, Aoki J, Arakawa T, Iida J, Imamura K, Itoh M, Kato T, Kawaji


H, Kawagashira N, Kawashima T, Kojima M, Kondo S, Konno H, Nakano K, Ninomiya N, Nishio
T, Okada M, Plessy C, Shibata K, Shiraki T, Suzu ki S, Tagami M, Waki K, Watahiki A, Okamura-
Oho Y, Suzuki H, Kawai J, Hayashizaki Y. The transcriptional landscape of the mammalian
genome. Science 2005;309:1559–1563. [PubMed: 16141072]
188. Cheng J, Kapranov P, Drenkow J, Dike S, Brubaker S, Patel S, Long J, Stern D, Tammana H, Helt
G, Sementchenko V, Piccolboni A, Bekiranov S, Bailey DK, Ganesh M, Ghosh S, Bell I, Gerhard
DS, Gingeras TR. Transcriptional maps of 10 human chromosomes at 5-nucleotide resolution.
Science 2005;308:1149–1154. [PubMed: 15790807]
189. Janssens AC, Gwinn M, Khoury MJ, Subramonia-Iyer S. Does genetic testing really improve the
prediction of future type 2 diabetes? PlosMed 2006;3:e114.
NIH-PA Author Manuscript

Cellscience. Author manuscript; available in PMC 2006 August 3.


Das and Elbein Page 28

Table 1
Summary of published Genome-wide Scans for T2DM
The table summarizes published genome wide scans for T2DM or related traits that are at least 10cM density.
Replicated regions are marked by red color. Regions are selected for reported LOD score of at least 2.0 or
NIH-PA Author Manuscript

equivalent, some regions represent subset analysis or below this threshold but replicate other findings. Locations
are presented according to current Marshfield Linkage Map, based on most significant marker or region.

Study Population Size and type Suggestive linkage (LOD>2 or


equivalent)

Hanis (1996) Mexican American 330 affected sib pairs 2q(260cM);15(45cM)


Mahtani (1996) Botanian Finn 26 extended families 12q(140cM)
Hanson (1998) Pima Indian 264 nuclear families 11q(139cM);1q(175–200cM);6
(130cM)
Duggirala (1999) Mexican American 32 extended families 10q(150cM);3p(70cM);4q(170cM);
9p(14cM)
Elbein (1999) US Caucasian 42 extended families 1q(165.62–170.84cM); 2q
(245.4cM); 18p(0cM)
Ghosh (1999, 2000) Finnish Caucasian (FUSION I) 477 families (716 affected sib 20p(20cM); 20q(60–85cM); 11q
pairs) (85cM)
Ehm (2000) US Caucasian 77 Caucasian families 5(78cM); 12(94cM); X(68.74cM) :
for Caucasian
[GENNID] African American 65 African-American families 10p(28cM) : for African American
Mexican American 53 Mexican American families 3p(58cM) : for Mexican American
Vionnet (2000) French Caucasian 143 families 1q21–q24(169.68cM); 3q
(207.73cM)
Permutt (2001) Ashkenazi Caucasian 267 families 4q(178cM);20q(50–63cM)
Parker (2001) Finnish Swedish Caucasian 353 families 18p11(18–30cM)
Wiltshire (2001) UK Caucasian 573 families (743 sib pairs) 1q21(181.49–191.52cM); 8p(33–
NIH-PA Author Manuscript

44cM); 10(120cM)
Francke (2001) Indo-Mauritian 99 families 1q44(318cM);3q22(165cM)
Luo (2001) Chinese Han 102 families 20q (75cM);9p21.3(42.73cM);
9q21.13(70.33cM); 2q36.1
(221.13cM)
Meigs (2002) Framingham Caucasian 330 families 1q21(192.05cM) HbA1c level;
1q42.2(247.23) Mean glucose
Mori (2002) Japanese 224 affected sib pairs 2q(194cM); 7p(7cM); 11p13(46cM);
15q13(42cM); 20q12(62cM)
Lindgren (2002) Botnian Finn 58 families 9q21(66.32cM)
Bushfield (2002) Indigenous Australian 1 extended families(232 2q24(171.04cM); 8p22
members) (31.73cM); 3q27(224.88cM)
Hsueh (2003) Amish Caucasian lextended family (691 member) 1q31(116cM); 1q21(166cM); 14q11
(7cM)
Van Tilburg (2003) Dutch Caucasian 178 families 12q12–q14(48–68cM);18p11(18–
28cM)
Aulchenko (2003) Dutch Caucasian 1 extended pedigree (79 nuclear 18p11(8cM)
families)
Reynisdottir (2003) Iceland Caucasian 227 families 5q34–q35.2(177.06cM)
Iwasaki (2003) Japanese 256 affected sib pairs 15q(45.8cM); 21q(48cM); 9q
(140cM) : in ordered subset analysis
Xiang (2004) Chinese Han 257 families 1q21(192cM);6q21(129cM)
Ng (2004) Chinese (Hong Kong) 64 families (126 sibpairs) 1q21–q25(173cM); 4q(135cM)
Silander (2004) Finnish Caucasian (FUSION II) 242 affected sib pairs 6q16–q22(113.03cM);14q23
(66.81cM)
Sale (2004) African American 247 families 6q25.3(164.78cM)
Rotimi (2004) West Africa 343 sib pairs 20q13.3(62–95cM); 19
NIH-PA Author Manuscript

(43.4cM); 12q(131cM)
Nawata (2004) Japanese 102 affected sib pairs 11p13–p12(99.01cM);6q15–q16
(51.95cM)
Hunt (2005) Mexican American (SAFDGS) 39 extended families 3p(102cM)
Zhao (2005) Chinese Han 34 Multiplex nuclear families 1p34.2(73.21cM);1q24.3
(188.85cM)

Cellscience. Author manuscript; available in PMC 2006 August 3.

You might also like