Entry - *138079 - GLUCOKINASE; GCK - OMIM
* 138079

GLUCOKINASE; GCK


Alternative titles; symbols

GK; GLK
HEXOKINASE 4; HK4
LIVER GLUCOKINASE; LGLK


HGNC Approved Gene Symbol: GCK

Cytogenetic location: 7p13     Genomic coordinates (GRCh38): 7:44,143,213-44,189,439 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
7p13 Diabetes mellitus, noninsulin-dependent, late onset 125853 AD 3
Diabetes mellitus, permanent neonatal 1 606176 AR 3
Hyperinsulinemic hypoglycemia, familial, 3 602485 AD 3
MODY, type II 125851 AD 3

TEXT

Description

The phosphorylation of glucose at the sixth carbon position is the first step in glycolysis. The reaction is catalyzed by a family of enzymes called hexokinases, types I (142600) through IV (glucokinase). Glucokinase (GCK; EC 2.7.1.1) is a structurally and functionally unique member of this family. Glucokinase is expressed only in mammalian liver and pancreatic islet beta cells. Because of its unique functional characteristics, the enzyme plays an important regulatory role in glucose metabolism. The rate of glucose metabolism in liver and pancreas is a function of the activity of the enzyme (summary by Matsutani et al., 1992).


Cloning and Expression

Using rat islet glucokinase cDNA to screen a human islet cDNA library, followed by screening a human liver cDNA library and RT-PCR of liver RNA, Tanizawa et al. (1991) cloned 2 GCK splice variants, which they called LGLK1 and LGLK2. LGLK1 encodes a deduced 464-amino acid protein with a calculated molecular mass of 52 kD. LGLK2 encodes a deduced 466-amino acid protein with 16 different N-terminal amino acids compared with LGLK1. LGLK2 shares 98% identity with rat islet glucokinase. Northern blot analysis detected a 2.8-kb transcript in liver total RNA. In vitro translation of LGLK1 cDNA resulted in proteins with apparent molecular masses of 52 and 48 kD by SDS-PAGE. Proteins of similar masses were translated from LGLK2.


Gene Structure

Stoffel et al. (1992) determined that the GCK gene contains 12 exons. Alternative promoters and alternative splicing encode different-sized proteins.


Mapping

Matsutani et al. (1992) identified a region of compound dinucleotide repeats located approximately 10 kb 3-prime to the GCK gene and used oligonucleotide primers and PCR amplification to demonstrate a number of alleles created by this repeat region. The variable numbers of CT and CA repeats represented altogether 6 alleles that range in length from +10 to -15 nucleotides compared to the most common (Z) allele. Two alleles, Z+10 and Z-15, appeared to be unique to American blacks, while a Z+6 allele was observed only in the Caucasian population studied. Using the PCR assay, they localized the human glucokinase gene to chromosome 7 in a panel of rodent/human somatic cell lines. Genetic analysis in CEPH pedigrees placed the dinucleotide repeat element, and thereby GCK, on 7p between TCRG (see 186970) and the RFLP locus D7S57.

Mishra et al. (1992) placed GCK between D7S57 proximally and D7S65 distally. It was estimated to lie about 4.7 cM from D7S65; D7S65 was found to be about 3.4 cM from TCRG, which is located at 7p15-p14. (Mishra et al. (1992) gave the location of TCRG as 7p15 in their Figure 2.)


Gene Function

Byrne et al. (1994) studied pancreatic beta-cell function in 4 males and 2 females from kindreds with known GCK mutations and in 6 controls pair-matched for age, weight, and sex. All of the GCK subjects with GCK mutations were found to have elevated fasting and postprandial glucose levels in comparison to the 6 controls. Insulin secretion rates (ISRs) were estimated. The results supported a key role for GCK in determining the in vivo glucose/ISR dose-response relationships and defined the alterations in beta-cell responsiveness that occur in subjects with GCK mutations.

Heimberg et al. (1996) reported the expression of glucokinase in rat glucagon-producing islet alpha cells, which are negatively regulated by glucose. Purified rat alpha cells expressed glucokinase mRNA and protein with the same transcript length, nucleotide sequence, and immunoreactivity as the beta-cell isoform. Glucokinase activity accounted for more than 50% of glucose phosphorylation in extracts of alpha cells and for more than 90% of glucose utilization in intact cells. A glucagon-producing tumor also contained glucokinase mRNA, protein, and enzymatic activity. The data indicated that glucokinase may serve as a metabolic glucose sensor in pancreatic alpha cells and, hence, mediate a mechanism for direct regulation of glucagon release by extracellular glucose. Since the alpha cells do not express GLUT2 (138160), Heimberg et al. (1996) suggested that glucose sensing does not necessarily require the coexpression of GLUT2 and glucokinase.

On the basis of studies in 7 glucokinase-deficient subjects with normal glycosylated hemoglobin and 12 control subjects using (13)C nuclear magnetic spectroscopy during a day in which 3 isocaloric mixed meals were ingested, Velho et al. (1996) observed results suggesting that in addition to the altered beta-cell function, abnormalities in liver glycogen metabolism play an important role in the pathogenesis of hypoglycemia in patients with maturity-onset diabetes of the young type 2 (MODY2; 125851). Average fasting hepatic glycogen content was similar in both groups and increased in both after the meals with a continuous pattern throughout the day. However, the net increment in hepatic glycogen content after each meal was 30 to 60% lower in glucokinase-deficient than in control subjects. Glucokinase-deficient subjects had decreased net accumulation of hepatic glycogen and relatively augmented hepatic gluconeogenesis after meals.

Danial et al. (2003) undertook a proteomic analysis to assess whether BAD (603167) might participate in mitochondrial physiology. In liver mitochondria, BAD resides in a functional holoenzyme complex together with protein kinase A (see 176911) and protein phosphatase-1 (PP1; see 176875) catalytic units, WAVE1 (605035) as an A kinase-anchoring protein, and glucokinase. Using mitochondria from hepatocytes of Bad-deficient mice, Danial et al. (2003) demonstrated that BAD is required to assemble the complex, the lack of which results in diminished mitochondria-based glucokinase activity and blunted mitochondrial respiration in response to glucose. Glucose deprivation results in dephosphorylation of BAD, and BAD-dependent cell death. Moreover, Danial et al. (2003) demonstrated that the phosphorylation status of BAD helps regulate glucokinase activity. Mice deficient in BAD or bearing a nonphosphorylatable BAD (3SA) mutant (Datta et al., 2002) display abnormal glucose homeostasis, including profound defects in glucose tolerance. Danial et al. (2003) concluded that this combination of proteomics, genetics, and physiology indicates an unanticipated role for BAD in integrating pathways of glucose metabolism and apoptosis.

Using a yeast 2-hybrid assay and other protein interaction assays, Hofmeister-Brix et al. (2013) found that Midn (606700), via its the ubiquitin-like domain, bound glucokinase in rat and mouse pancreatic beta cells and cell lines. Binding was strongest at low glucose concentration, when glucokinase exists mainly in its inactive super-open-to-open conformation. Overexpression of fluorescence-tagged Midn inhibited glucokinase activity, reduced the affinity of glucokinase for glucose, and reduced glucose-induced insulin secretion in rat pancreatic MIN6 cells.

Using mouse models and the rat INS-1 pancreatic beta cell line, Kim et al. (2014) found that chronic ethanol consumption induced activating transcription factor-3 (ATF3; 603148), which then inhibited Gck transcriptional activity. Atf3 directly bound to a putative ATF/CREB site in the Gck promoter. Atf3 also counteracted the positive effect of Pdx1 (600733) on Gck transcriptional activity.


Molecular Genetics

Maturity-Onset Diabetes of the Young, Type 2

In 16 French families with maturity-onset diabetes of the young (MODY2; 125851), Froguel et al. (1992) found linkage of the disease with GCK. There was statistically significant evidence of genetic heterogeneity, with an estimated 45 to 95% of the 16 families showing linkage to glucokinase. Because glucokinase is a key enzyme of blood glucose homeostasis, the results suggested a pathogenetic connection. The relationship was clinched by the demonstration by Vionnet et al. (1992) of a nonsense mutation in the GCK gene (138079.0001).

Velho et al. (1992) studied pancreatic beta-cell secretory function in 9 patients from 4 GCK-linked MODY kindreds. They found that beta-cell secretory response to continuous glucose stimulus during a hyperglycemic glucose clamp was significantly reduced. This finding was different from that for noninsulin-dependent diabetes mellitus with late age of onset or MODY not linked to GCK. Fasting plasma insulin and C-peptide levels in patients were inappropriately low in relation to concomitant plasma glucose level. Furthermore, during a hyperinsulinemic euglycemic clamp, endogenous insulin secretion at euglycemia was suppressed in patients but not in controls. Velho et al. (1992) interpreted these results as suggesting that mutant GCK leads to chronic hyperglycemia by raising the threshold of circulating glucose levels which induces insulin secretion. This was the first demonstration of a primary pancreatic secretory defect associated with a form of NIDDM.

Froguel et al. (1993) detected 16 mutations of the GCK gene in 18 of 32 families with MODY. An explanation for the way a mutation in glucokinase might cause diabetes follows from the work of Matschinsky (1990), who developed the concept that glucokinase acts as the glucose sensor of beta cells. According to this concept, the rates of glucose metabolism and insulin secretion are closely linked, with both being determined by the plasma glucose concentration. In beta cells and hepatocytes, the rate of glucose metabolism is determined by the rate of glucose phosphorylation, which is catalyzed by glucokinase. Beta cells and hepatocytes also contain glucose transporter-2 (GLUT2; 138160), an insulin-independent cellular protein that mediates the transport of glucose into cells. The capacity of GLUT2 to transport glucose is very high, facilitating rapid equilibrium between extracellular and intracellular glucose. Thus, in effect, the extracellular glucose concentrations are sensed by glucokinase.

Randle (1993) provided a review of glucokinase and candidate genes for type II MODY. Matschinsky et al. (1993) also reviewed glucokinase as the pancreatic beta-cell glucose sensor and 'diabetes gene.' They tabulated 16 separate mutations in the GCK gene found in families with MODY.

Johansen et al. (2005) examined the prevalence and nature of mutations in the 3 common MODY genes HNF4A (600281), GCK, and TCF1 (142410) in Danish patients with a clinical diagnosis of MODY and determined metabolic differences in probands with and without mutations in HNF4A, GCK, and TCF1. They identified 29 different mutations in 38 MODY families. Fifteen of the mutations were novel. The variants segregated with diabetes within the families, and none of the variants were found in 100 normal Danish chromosomes. Their findings suggested a relative prevalence of 3% of MODY1 (125850) (2 different mutations in 2 families), 10% of MODY2 (7 in 8), and 36% of MODY3 (600496) (21 in 28) among Danish kindred clinically diagnosed as MODY. No significant differences in biochemical and anthropometric measurements were observed at baseline examinations. Forty-nine percent of the families carried mutations in the 3 examined MODY genes.

Vits et al. (2006) identified 19 different GCK mutations, including 11 novel mutations (see, e.g., 138079.0014), in 33 (26.6%) of 124 Belgian probands with MODY.

Pinterova et al. (2007) screened the GCK gene in 92 Czech probands fulfilling classic MODY criteria and identified 15 different missense mutations in 27 (29%) patients; the mutations were not found in 50 unrelated healthy Czech individuals. Pinterova et al. (2007) concluded that mutations in GCK are a common cause of MODY in the Czech population.

In a mainland Chinese family with a clinical profile similar to that of previously reported MODY2 families, Shen et al. (2011) analyzed the GCK gene and identified a heterozygous missense mutation (E339K; 138079.0016) that segregated with the disease and was not found in 200 controls. Functional analysis indicated that the mutation inactivated enzyme kinetics and severely impaired GCK protein stability.

The variable phenotype observed with mutations in the GCK gene includes gestational diabetes mellitus. This fact prompted Stoffel et al. (1993) to screen a group of women with gestational diabetes who also had a first-degree relative with diabetes mellitus for the presence of GCK mutations. Among 40 subjects, they identified 2 with heterozygous mutations (138079.0007-138070.0008), suggesting a prevalence of approximately 5%. Extrapolating from this result, the prevalence of glucokinase-deficient MODY among Americans may be approximately 1 in 2,500.

Fetal insulin secretion in response to maternal glycemia plays a key role in fetal growth, and adult insulin secretion is a primary determinant of glucose tolerance. Hattersley et al. (1998) hypothesized that a defect in the sensing of glucose by the pancreas, caused by a heterozygous mutation in the glucokinase gene, could reduce fetal growth and birth weight in addition to causing hyperglycemia after birth. In 58 offspring, where one parent had a glucokinase mutation, the inheritance of a glucokinase mutation by the fetus resulted in a mean reduction of birth weight of 533 g (P = 0.002). In 19 of 21 sib pairs discordant for the presence of a glucokinase mutation, the child with the mutation had a lower birth weight, with a mean difference of 521 g (P = 0.0002). Maternal hyperglycemia due to a glucokinase mutation resulted in a mean increase in birth weight of 601 g (P = 0.001). The effects of maternal and fetal glucokinase mutations on birth weight were additive. Hattersley et al. (1998) proposed that these changes in birth weight reflect changes in fetal insulin secretion which are influenced directly by the fetal genotype and indirectly, through maternal hyperglycemia, by the maternal genotype.

Dunger et al. (1998) demonstrated an association between the insulin VNTR (variable number tandem repeat) 'locus,' and specifically the VNTR III genotype, and larger size at birth. McCarthy (1998) discussed the significance of these observations in relation to the 'thrifty genotype' hypothesis originally proposed by Neel (1962) and updated by Neel et al. (1998).

Permanent Neonatal Diabetes Mellitus 1

Njolstad et al. (2001) described 2 patients in whom complete deficiency of glucokinase caused permanent neonatal-onset diabetes (PNDM1; 606176). One patient was homozygous for an M210K mutation (138079.0010); the other was homozygous for a T228M mutation (138079.0003). Both patients showed total absence of basal insulin release.

Hyperinsulinemic Hypoglycemia

In 'family 3' studied by Thornton et al. (1998), in which affected members had hyperinsulinemic hypoglycemia (see HHF3, 602485), Glaser et al. (1998) identified heterozygosity for an activating mutation (138079.0009) in the GCK gene.

In a 14-year-old obese boy with a history of neonatal hypoglycemia treated with diazoxide, who was experiencing hypoglycemic episodes associated with seizures and unconsciousness, Christesen et al. (2002) identified heterozygosity for an activating mutation in the GCK gene (138079.0012). The boy's normal-weight mother, who had asymptomatic fasting hypoglycemia, carried the same mutation. Christesen et al. (2002) noted the striking contrast in clinical presentation between the mother and son with the same mutation, and recalled a similar situation in the family reported by Glaser et al. (1998), in which the proband was obese and severely hyperinsulinemic, whereas his sister who carried the same mutation was of normal weight, had relatively low insulin levels, and milder clinical symptoms.

In a Finnish woman with severe hyperinsulinemic hypoglycemia from birth, who had severe mental retardation and died at age 29 still having hypoglycemic seizures, Cuesta-Munoz et al. (2004) identified heterozygosity for a de novo activating mutation in the GCK gene (138079.0013).

Kassem et al. (2010) reported a young girl with severe neonatal hypoglycemia due to a missense mutation in the GCK gene (138079.0015) in whom mean islet cell areas in both the head and the tail of the pancreas were significantly larger than those of 5 age-matched controls and those of 2 age-matched patients with diffuse hypoglycemia due to ABCC8 mutations (600509; see HHF1, 256450). Noting a previously reported HHF3 patient in whom quantitative histologic analysis of pancreatic specimens showed a similar increase in the mean islet profile (Cuesta-Munoz et al. (2004)), Kassem et al. (2010) suggested that histologic findings in infants with hyperinsulinemic hypoglycemia might differ according to the genetic cause of the condition.

Penetrance of GCK Mutations in Diabetes

Mirshahi et al. (2022) comprehensively assessed the penetrance and prevalence of pathogenic variants in HNF1A (142410), HNF4A (600281), and GCK that account for more than 80% of monogenic diabetes. Mirshahi et al. (2022) analyzed clinical and genetic data from 1,742 clinically referred probands, 2,194 family members, clinically unselected individuals from a US health system-based cohort of 132,194 individuals, and a UK population-based cohort of 198,748 individuals, and found that 1 in 1,500 individuals harbor a pathogenic variant in one of these genes. The penetrance of pathogenic GCK variants was similar (89 to 97%) across all cohorts. The penetrance of diabetes for HNF1A and HNF4A pathogenic variants was substantially lower in the clinically unselected individuals compared to clinically referred probands and was dependent on the setting (32% in the population, 49% in the health system cohort, 86% in a family member, and 98% in probands for HNF1A). The relative risk of diabetes was similar across the clinically unselected cohorts, highlighting the role of environment/ other genetic factors. The authors suggested that for HNF1A and HNF4A, genetic interpretation and counseling should be tailored to the setting in which a pathogenic monogenic variant was identified. GCK is an exception with near-complete penetrance in all settings.

Other Associations

Rowe et al. (1995) evaluated 35 microsatellite marker loci on human chromosome 7 for linkage to insulin-dependent diabetes mellitus (IDDM) in 339 affected sib-pair families. Increased sharing of parental haplotypes in affected sib pairs was detected for 2 microsatellite markers flanking GCK. Preferential transmission of alleles to affected offspring was observed at one of these marker loci, GCK3, indicating linkage disequilibrium between the marker and the disease susceptibility locus. This combination of linkage and disease association suggested to Rowe et al. (1995) that glucokinase, or a gene in the vicinity, plays an important role in IDDM susceptibility. (Rowe et al. (1995) used the designations GCK1, GCK2, and GCK3 for 3 modestly polymorphic microsatellite markers near the GCK gene (Concannon, 1995). These are not symbols for separate glucokinase genes; there is only a single GCK gene on 7p, although as a result of alternative, tissue-specific promoters, there are islet- and liver-specific forms that differ in their amino-terminal sequences.)

Stone et al. (1996) used SSCP analysis to determine whether a G-to-A variant at position -30 of the beta-cell promoter of the GCK gene is associated with impaired beta-cell function. The variant was observed more frequently in Japanese-American men with impaired glucose tolerance than in Japanese-American men with normal glucose tolerance. Beta-cell function was assessed using the ratio of the incremental response in immunoreactive insulin to that of glucose during the first 30 minutes of the oral glucose tolerance test. They concluded that the -30 promoter variant is associated with reduced beta-cell function in middle-aged Japanese-American men and may contribute to the high risk of abnormal glucose tolerance in this population. They noted that the polymorphism has been observed in other populations.

Marz et al. (2004) analyzed the GCK -30G-A variant in 2,567 patients with and in 731 individuals without coronary artery disease (CAD; see CHDS1, 607339) by angiography and found that the A allele of the pancreatic GCK promoter increased the risk of CAD in individuals both with and without type II diabetes mellitus (ORs, 1.92 and 1.27, respectively). The A allele was also associated with an increased prevalence of type II diabetes mellitus, particularly among CAD patients.

For discussion of an association between variation in the GCK gene and fasting plasma glucose levels, see FGQTL2 (613219).


Genotype/Phenotype Correlations

Gloyn (2003) stated that given the central role of glucokinase in the regulation of insulin release, it is understandable that mutations in the GCK gene can cause both hyper- and hypoglycemia. Heterozygous inactivating mutations in GCK cause MODY, characterized by mild hyperglycemia, which is present at birth, but is often only detected later in life. Homozygous inactivating GCK mutations result in a more severe phenotype, presenting at birth as permanent neonatal diabetes mellitus. Several heterozygous activating GCK mutations causing hypoglycemia had been reported as instances of familial hyperinsulinism (see HHF3, 602485). Gloyn (2003) tabulated 195 mutations in the GCK gene, found in 285 families. There were no common mutations and the mutations were distributed throughout the gene. Mutations causing hypoglycemia are located in various exons in a discrete region of the protein termed the heterotropic allosteric activator site. Patients with MODY type 2 (due to GCK mutations) rarely need pharmacologic treatment and most are managed on diet alone. During pregnancy, some women with GCK mutations are treated with insulin, which, due to its role in fetal growth, can result in babies that are large for gestational age if they do not also have the GCK mutation (Hattersley et al., 1998). Microvascular complications are rare in MODY2. These patients do not need to be followed in a diabetes clinic, but should have an HbA1c checked annually. Some patients with hyperinsulinism due to GCK mutations may not require pharmacologic intervention and may be able to control their symptoms by eating regularly. Patients with hyperinsulinism due to mutation in the ABCC8 gene (600509) or the KCNJ11 gene (600937) respond poorly or are totally refractory to the potassium channel opener diazoxide, which is used to inhibit insulin secretion. Conversely, patients with hyperinsulinism resulting from GCK mutations respond well or tolerated diazoxide (Gloyn et al., 2003).


Animal Model

Using homologous recombination in mouse embryonic stem cells to disrupt the glucokinase gene, Bali et al. (1995) showed that heterozygous mice showed a phenotype similar to MODY. Reduced islet glucokinase activity caused mildly elevated fasting blood glucose levels. Hyperglycemic clamp studies revealed decreased glucose tolerance and abnormal liver glucose metabolism. These findings demonstrated a key role for glucokinase in glucose homeostasis and implicated both islets and liver in the MODY disease. Aizawa et al. (1996) made a detailed study of pancreatic beta-cell function in the isolated pancreatic islets of heterozygous GCK knockout mice. The beta cells of these mice displayed impaired glucose sensitivity, poor discrimination of alpha- and beta-glucose anomers, and normal activity of the ATP-sensitive K+ channel. The mice displayed impaired insulin release after insulin treatment and age-related worsening of glucose tolerance.

Understanding how the beta cells of the pancreatic islet sense an alteration in glucose levels is critical to understanding how glucose homeostasis is maintained. Grupe et al. (1995) noted that glucose sensing is mediated through an increase in the rate of intracellular catabolism of glucose rather than a ligand-receptor interaction. However, it is likely that there is still a need for a rate-limiting step in glucose catabolism to serve the role of glucose sensor. A compelling case can be made for glucokinase fulfilling this role, since phosphorylation by GLK appears to be the rate-limiting step for glucose catabolism in beta cells. To test this possibility and to resolve the relative roles of liver and beta-cell GLK in maintaining glucose levels, Grupe et al. (1995) generated mice completely deficient in GLK and transgenic mice in which GLK is expressed only in beta cells. In mice with only 1 GLK allele, blood glucose levels were elevated and insulin secretion was reduced. GLK-deficient mice died perinatally with severe hyperglycemia. Expression of GLK in beta cells in the absence of expression in the liver is sufficient for survival.

Grimsby et al. (2003) identified a class of antidiabetic agents that act as nonessential, mixed-type glucokinase activators that increase the glucose affinity and maximum velocity of glucokinase. Glucokinase activators augmented both hepatic glucose metabolism and glucose-induced insulin secretion from isolated rodent pancreatic islets, consistent with the expression and function of glucokinase in both cell types. In several rodent models of NIDDM, glucokinase activators lowered blood glucose levels, improved the results of glucose tolerance tests, and increased hepatic glucose uptake.

Using N-ethyl-N-nitrosourea (ENU) mutagenesis, Inoue et al. (2004) generated diabetic mice. The authors screened 9,375 animals and identified 11 mutations in the glucokinase (Gk) gene that were associated with hyperglycemia. Four had previously been found in human MODY2 (125851) patients, and 1 was found previously in a patient with permanent neonatal diabetes mellitus (PNDM; 606176). Some of the Gk mutant lines displayed impaired glucose-responsive insulin secretion, and the mutations had different effects on Gk mRNA levels and/or the stability of the GK protein.

Terauchi et al. (2007) generated mice with haploinsufficiency of beta cell-specific Gck and observed that on a high-fat diet, Gck +/- mice had decreased beta cell replication and insufficient beta cell hyperplasia compared to wildtype mice despite a similar degree of insulin resistance. On a high-fat diet, Gck +/- mouse islets showed decreased insulin receptor substrate-2 (Irs2; 600797) expression compared with wildtype islets. Overexpression of Irs2 in beta cells of Gck +/- mice partially prevented diabetes by increasing beta cell mass. Terauchi et al. (2007) suggested that both GCK and IRS2 are critical for beta cell hyperplasia to occur in response to high-fat diet-induced insulin resistance.


ALLELIC VARIANTS ( 16 Selected Examples):

.0001 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLU279TER
  
RCV000017512

In the process of single-strand conformation polymorphism (SSCP) analysis of exon 7 in a French family with GCK-linked MODY (MODY2; 125851), previously studied by Froguel et al. (1992), Vionnet et al. (1992) demonstrated a G-to-T substitution in codon 279 which changed GAG (glutamic acid) to TAG, an amber termination codon. An individual in this family who had noninsulin-dependent diabetes mellitus (NIDDM) did not show linkage to GCK; moreover, she did not have the nonsense mutation in codon 279. Thus, there were 2 forms of NIDDM in this kindred. Hattersley et al. (1992) likewise found tight linkage of MODY to a macrosatellite polymorphism associated with the GCK locus; peak lod = 4.60 at theta = 0.0. In a second MODY pedigree, they excluded linkage; lod = -7.36 at theta = 0.0.


.0002 TYPE 2 DIABETES MELLITUS

GCK, ARG186TER
  
RCV000017513...

In a Japanese family, Katagiri et al. (1992) found a correlation between the presence of late-onset noninsulin-dependent diabetes mellitus (125853) or impaired glucose tolerance and a nonsense mutation in exon 5 of the glucokinase gene: a C-to-T transition changing codon 186 from CGA (arginine) to TGA (an amber termination codon). The mutation was predicted to delete 60% of the amino acid residues of the glucokinase derived from the affected allele. The family was detected through a 59-year-old woman who was the twenty-third subject screened. Froguel and Velho (1993) raised the question as to whether the Japanese family may in fact have had MODY, because in their experience in French families carrying mutations in GCK, hyperglycemia often went undiagnosed for a long time. Chiu et al. (1993) likewise questioned whether the disorder in the Japanese families should be termed NIDDM and pointed out that in young patients with glucokinase mutations the degree of hyperglycemia is so mild that values often do not exceed the renal threshold. Therefore, absence of glycosuria cannot be used as a criterion for distinguishing MODY from NIDDM. Permutt et al. (1992) pointed out that structural mutations in the GCK gene are very rare (less than 2%) in American black and Caucasian NIDDM patients.


.0003 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

DIABETES MELLITUS, PERMANENT NEONATAL, 1, INCLUDED
GCK, THR228MET
  
RCV000020167...

Maturity-Onset Diabetes of the Young, Type 2

Stoffel et al. (1992) stated that DNA polymorphisms in the GCK gene has been shown to be tightly linked to MODY (MODY2; 125851) in approximately 80% of French families. They identified 2 further missense mutations in exon 7 in families with MODY: thr228 to met (T228M) and gly261 to arg (G261R; 138079.0004). A TGC-to-TAC change at codon 228 and a CCC-to-TCC change at codon 261 were responsible. Using computer-assisted modeling and the crystal structure of the related yeast hexokinase B as a simple model for human beta-cell glucokinase, Stoffel et al. (1992) obtained data which suggested to them that mutation of thr228 affects affinity for ATP and mutation of gly261 alters glucose binding. The identification of mutations in glucokinase, a protein that plays an important role in hepatic and beta-cell glucose metabolism, indicates that early-onset noninsulin-dependent diabetes mellitus may be primarily a disorder of carbohydrate metabolism.

Velho et al. (1997) identified the T228M mutation in affected members of a family with glucokinase-related MODY.

Diabetes Mellitus, Permanent Neonatal 1

In an 8-year-old girl of Italian ancestry, Njolstad et al. (2001) identified homozygosity for the T228M mutation in the GCK gene as the cause of neonatal-onset diabetes mellitus (PNDM1; 606176). The child had shown hyperglycemia and marked growth retardation at birth. Her father had impaired glucose tolerance and her mother had impaired fasting glycemia.


.0004 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLY261ARG
  
RCV000017515...

Stoffel et al. (1992) identified missense mutations in exon 7 of the GCK gene in families with maturity diabetes of the young (MODY2; 125851): thr228 to met (T228M; 138079.0003) and gly261 to arg (G261R). A TGC-to-TAC change at codon 228 and a CCC-to-TCC change at codon 261 were responsible. Using computer-assisted modeling and the crystal structure of the related yeast hexokinase B as a simple model for human beta-cell glucokinase, Stoffel et al. (1992) obtained data which suggested to them that mutation of thr228 affects affinity for ATP and mutation of gly261 alters glucose binding. The identification of mutations in glucokinase, a protein that plays an important role in hepatic and beta-cell glucose metabolism, indicates that early-onset noninsulin-dependent diabetes mellitus may be primarily a disorder of carbohydrate metabolism.


.0005 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLY299ARG
  
RCV000017516...

In a large British pedigree with many cases of MODY (MODY2; 125851) in 5 generations, Stoffel et al. (1992) demonstrated a G-to-C transversion in the GCK gene, which converted codon 266 from glycine to arginine. The mutation also created a HhaI site which allowed them to construct a rapid PCR test for the mutation. Applying this to cases of classic late-onset type 2 diabetes mellitus, they found the same mutation in 1 of 50 patients. All 9 relatives of this patient who had inherited the mutation had type 2 diabetes, although 6 others without the mutation also had diabetes. MODY had not previously been considered in this family.


.0006 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, IVS4DS, 15-BP DEL
  
RCV000992052...

Sun et al. (1993) analyzed the nucleotide sequence of exon 4 and its flanking intronic regions of the GCK gene in 4 hyperglycemic members of a MODY (MODY2; 125851) family and found a deletion of 15 bp, which removed the T of the GT in the donor splice site of intron 4 and the following 14 basepairs. This deletion resulted in 2 aberrant transcripts: one with exon 5 missing and the other with activation of a cryptic splice site leading to the removal of the last 8 codons of exon 4. This intronic deletion seemed to cause a more severe form of glucose intolerance than did the missense and nonsense mutations previously reported.


.0007 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, SER131PRO
  
RCV000017518...

Stoffel et al. (1993) found heterozygosity for a ser131-to-pro mutation in the GCK gene in an obese 31-year-old Puerto Rican woman with gestational diabetes (MODY2; 125851) in her first pregnancy. She reportedly had borderline elevated blood glucose levels at age 26. Stoffel et al. (1993) showed defective enzymatic properties of the enzyme carrying the ser131-to-pro mutation.


.0008 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLU265TER
  
RCV000017519...

Stoffel et al. (1993) found a glu265-to-ter mutation in a thin 32-year-old Caucasian woman with gestational diabetes mellitus (MODY2; 125851) who was in her second pregnancy. She reportedly had elevated fasting blood glucose levels since age 16. The subject's mother and 2 sisters had diabetes mellitus treated with diet or oral hypoglycemic agents.


.0009 HYPERINSULINEMIC HYPOGLYCEMIA, FAMILIAL, 3

GCK, VAL455MET
  
RCV000017520...

In a 31-year-old man and his 36-year-old sister from the 3-generation 'family 3' with autosomal dominant hyperinsulinemic hypoglycemia (602485) previously studied by Thornton et al. (1998), Glaser et al. (1998) identified a val455-to-met (V455M) mutation in the glucokinase gene. When expressed in vitro, the V455M mutation increased the affinity of glucokinase for glucose, resulting in higher rates of glycolysis at low glucose concentrations and therefore a higher rate of insulin secretion at any plasma glucose concentration. The finding confirmed the importance of glucokinase as a primary regulator of glucose-controlled insulin secretion in beta cells. This mutation was not found in 37 unrelated white families with hyperinsulinism, including 6 with an apparently autosomal dominant form.


.0010 DIABETES MELLITUS, PERMANENT NEONATAL, 1

MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2, INCLUDED
GCK, MET210LYS
  
RCV000017521...

Njolstad et al. (2001) described an infant girl of Norwegian ancestry with neonatal diabetes mellitus (PNDM1; 606176) that persisted thereafter. The parents were first cousins, and both had glucose intolerance. At 5 years of age, the patient developed epilepsy, probably as a sequela of a neonatal brain abscess. A sister had typical type I diabetes developing at the age of 7 years. The mother had a diagnosis of gestational diabetes at the age of 25 years. The father had impaired fasting glycemia that was treated with diet. After excluding other candidate genes, Njolstad et al. (2001) found that the child was homozygous for a T-to-A transversion (ATG-AAG) at nucleotide 629 in exon 6 of the GCK gene, resulting in a met210-to-lys mutation (M210K). Her parents and sister were heterozygous for the mutation, which cosegregated with diabetes or hyperglycemia in other members of the family. Thus, in this family, heterozygosity caused GCK-related MODY (MODY2; 125851) and homozygosity caused neonatal-onset diabetes.


.0011 MOVED TO 138079.0003


.0012 HYPERINSULINEMIC HYPOGLYCEMIA, FAMILIAL, 3

GCK, ALA456VAL
  
RCV000017525...

In a 14-year-old obese boy with a history of neonatal hypoglycemia treated with diazoxide, who was experiencing hypoglycemic episodes associated with seizures and unconsciousness (see HHF3, 602485), Christesen et al. (2002) identified heterozygosity for an ala456-to-val (A456V) substitution in exon 10 of the GCK gene. Kinetic analysis showed this to be an activating mutation. The boy's normal-weight mother, who had asymptomatic fasting hypoglycemia, carried the same mutation; the mutation was not found in his normoglycemic brother nor in 80 controls.


.0013 HYPERINSULINEMIC HYPOGLYCEMIA, FAMILIAL, 3

GCK, TYR214CYS
  
RCV000017526

In a Finnish woman with severe hyperinsulinemic hypoglycemia from birth (602485), who had severe mental retardation and was still having hypoglycemic seizures when she died at age 29, Cuesta-Munoz et al. (2004) identified heterozygosity for a de novo tyr214-to-cys (Y214C) substitution in exon 6 of the GCK gene. Although paternity was confirmed, the mutation was not found in her parents or her 2 healthy sisters. Kinetic analysis revealed that this mutation had the highest activity index (130-fold over wildtype) of all naturally occurring activating GCK mutations described. Cuesta-Munoz et al. (2004) noted that this phenotype was considerably more severe than that of previously reported patients (see 138079.0009 and 138079.0012).


.0014 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, ALA378THR
  
RCV000017527...

In affected members from 5 Belgian families with MODY2 (125851), Vits et al. (2006) identified a 1132G-A transition in exon 9 of the GCK gene, resulting in an ala378-to-thr (A378T) substitution. All the probands originated from the Belgian province of West Flanders, suggesting a founder mutation; this was confirmed in 3 families by haplotype analysis.


.0015 HYPERINSULINEMIC HYPOGLYCEMIA, FAMILIAL, 3

GCK, VAL91LEU
   RCV000017528

Kassem et al. (2010) studied a young girl with severe neonatal hypoglycemia-3 (602485) due to a val91-to-leu (V91L) missense mutation in the GCK gene. Her father had a similar clinical course but neither his DNA nor pancreatic tissue was available for study. Quantitative histologic examination after subtotal pancreatectomy due to refractory disease revealed abnormally large islets, with some beta cells containing a large nucleus, and mean islet cell areas in both the head and the tail of the pancreas were significantly larger than those of 5 age-matched controls and those of 2 age-matched HHF1 (256450) patients.


.0016 MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLU339LYS
  
RCV000032978...

In 5 affected members over 3 generations of a mainland Chinese family with type 2 maturity-onset diabetes of the young (MODY2; 125851), Shen et al. (2011) identified heterozygosity for a G-A transition in exon 8 of the GCK gene, resulting in a glu339-to-lys (E339K) substitution. The mutation was not found in unaffected family members or in 200 controls. SDS-PAGE analysis of a bacterial expression system demonstrated that the protein yield of mutant GCK was significantly lower than wildtype; kinetic analysis showed that the E339K mutant had 1.4-fold and 9.9-fold lower affinity for glucose and ATP, respectively, compared to wildtype. The mutant GCK also exhibited thermal instability, with a dramatic decrease in activity at 45 degrees centigrade compared to 55 degrees centigrade for wildtype; in addition, wildtype GCK maintained 50% activity at 55 degrees centigrade for 30 minutes, whereas wildtype lost 97% of activity within 30 minutes.


REFERENCES

  1. Aizawa, T., Asanuma, N., Terauchi, Y., Suzuki, N., Komatsu, M., Itoh, N., Nakabayashi, T., Hidaka, H., Ohnota, H., Yamauchi, K., Yasuda, K., Yazaki, Y., Kadowaki, T., Hashizume, K. Analysis of the pancreatic beta cell in the mouse with targeted disruption of the pancreatic beta cell-specific glucokinase gene. Biochem. Biophys. Res. Commun. 229: 460-465, 1996. [PubMed: 8954920, related citations] [Full Text]

  2. Bali, D., Svetlanov, A., Lee, H.-W., Fusco-DeMane, D., Leiser, M., Li, B., Barzilai, N., Surana, M., Hou, H., Fleischer, N., DePinho, R., Rossetti, L., Efrat, S. Animal model for maturity-onset diabetes of the young generated by disruption of the mouse glucokinase gene. J. Biol. Chem. 270: 21464-21467, 1995. [PubMed: 7665557, related citations] [Full Text]

  3. Byrne, M. M., Sturis, J., Clement, K., Vionnet, N., Pueyo, M. E., Stoffel, M., Takeda, J., Passa, P., Cohen, D., Bell, G. I., Velho, G., Froguel, P., Polonsky, K. S. Insulin secretory abnormalities in subjects with hyperglycemia due to glucokinase mutations. J. Clin. Invest. 93: 1120-1130, 1994. [PubMed: 8132752, related citations] [Full Text]

  4. Chiu, K. C., Tanizawa, Y., Permutt, M. A. Non-sense mutation of glucokinase gene. (Letter) Lancet 341: 385-386, 1993. [PubMed: 8094164, related citations]

  5. Christesen, H. B. T., Jacobsen, B. B., Odili, S., Buettger, C., Cuesta-Munoz, A., Hansen, T., Brusgaard, K., Massa, O., Magnuson, M. A., Shiota, C., Matschinsky, F., Barbetti, F. The second activating glucokinase mutation (A456V): implications for glucose homeostasis and diabetes therapy. Diabetes 51: 1240-1246, 2002. [PubMed: 11916951, related citations] [Full Text]

  6. Concannon, P. Personal Communication. Seattle, Wash. 7/26/1995.

  7. Cuesta-Munoz, A. L., Huopio, H., Otonkoski, T., Gomez-Zumaquero, J. M., Nanto-Salonen, K., Rahier, J., Lopez-Enriquez, S., Garcia-Gimeno, M. A., Sanz, P., Soriguer, F. C., Laakso, M. Severe persistent hyperinsulinemic hypoglycemia due to a de novo glucokinase mutation. Diabetes 53: 2164-2168, 2004. [PubMed: 15277402, related citations] [Full Text]

  8. Danial, N. N., Gramm, C. F., Scorrano, L., Zhang, C.-Y., Krauss, S., Ranger, A. M., Datta, S. R., Greenberg, M. E., Licklider, L. J., Lowell, B. B., Gygi, S. P., Korsmeyer, S. J. BAD and glucokinase reside in a mitochondrial complex that integrates glycolysis and apoptosis. Nature 424: 952-956, 2003. [PubMed: 12931191, related citations] [Full Text]

  9. Datta, S. R., Ranger, A. M., Lin, M. Z., Sturgill, J. F., Ma, Y.-C., Cowan, C. W., Dikkes, P., Korsmeyer, S. J., Greenberg, M. E. Survival factor-mediated BAD phosphorylation raises the mitochondrial threshold for apoptosis. Dev. Cell 3: 631-643, 2002. [PubMed: 12431371, related citations] [Full Text]

  10. Dunger, D. B., Ong, K. K. L., Huxtable, S. J., Sherriff, A., Woods, K. A., Ahmed, M. L., Golding, J., Pembrey, M. E., Ring, S., ALSPAC Study Team, Bennett, S. T., Todd, J. A. Association of the INS VNTR with size at birth. Nature Genet. 19: 98-100, 1998. [PubMed: 9590300, related citations] [Full Text]

  11. Froguel, P., Vaxillaire, M., Sun, F., Velho, G., Zouali, H., Butel, M. O., Lesage, S., Vionnet, N., Clement, K., Fougerousse, F., Tanizawa, Y., Weissenbach, J., Beckmann, J. S., Lathrop, G. M., Passa, P., Permutt, M. A., Cohen, D. Close linkage of glucokinase locus on chromosome 7p to early-onset non-insulin-dependent diabetes mellitus. Nature 356: 162-164, 1992. Note: Erratum: Nature 357: 607 only, 1992. [PubMed: 1545870, related citations] [Full Text]

  12. Froguel, P., Velho, G. Non-sense mutation of glucokinase gene. (Letter) Lancet 341: 385 only, 1993. [PubMed: 8094163, related citations]

  13. Froguel, P., Zouali, H., Vionnet, N., Velho, G., Vaxillaire, M., Sun, F., Lesage, S., Stoffel, M., Takeda, J., Passa, P., Permutt, M. A., Beckmann, J. S., Bell, G. I., Cohen, D. Familial hyperglycemia due to mutations in glucokinase: definition of a subtype of diabetes mellitus. New Eng. J. Med. 328: 697-702, 1993. [PubMed: 8433729, related citations] [Full Text]

  14. Glaser, B., Kesavan, P., Heyman, M., Davis, E., Cuesta, A., Buchs, A., Stanley, C. A., Thornton, P. S., Permutt, M. A., Matschinsky, F. M., Herold, K. C. Familial hyperinsulinism caused by an activating glucokinase mutation. New Eng. J. Med. 338: 226-230, 1998. [PubMed: 9435328, related citations] [Full Text]

  15. Gloyn, A. L., Noordam, K., Willemsen, M. A. A. P., Ellard, S., Lam, W. W. K., Campbell, I. W., Midgley, P., Shiota, C., Buettger, C., Magnuson, M. A., Matschinsky, F. M., Hattersley, A. T. Insights into the biochemical and genetic basis of glucokinase activation from naturally occurring hypoglycemia mutations. Diabetes 52: 2433-2440, 2003. [PubMed: 12941786, related citations] [Full Text]

  16. Gloyn, A. L. Glucokinase (GCK) mutations in hyper- and hypoglycemia: maturity-onset diabetes of the young, permanent neonatal diabetes, and hyperinsulinemia of infancy. Hum. Mutat. 22: 353-362, 2003. [PubMed: 14517946, related citations] [Full Text]

  17. Grimsby, J., Sarabu, R., Corbett, W. L., Haynes, N.-E., Bizzarro, F. T., Coffey, J. W., Guertin, K. R., Hilliard, D. W., Kester, R. F., Mahaney, P. E., Marcus, L., Qi, L., Spence, C. L., Tengi, J., Magnuson, M. A., Chu, C. A., Dvorozniak, M. T., Matschinsky, F. M., Grippo, J. F. Allosteric activators of glucokinase: potential role in diabetes therapy. Science 301: 370-373, 2003. [PubMed: 12869762, related citations] [Full Text]

  18. Grupe, A., Hultgren, B., Ryan, A., Ma, Y. H., Bauer, M., Stewart, T. A. Transgenic knockouts reveal a critical requirement for pancreatic beta cell glucokinase in maintaining glucose homeostasis. Cell 83: 69-78, 1995. [PubMed: 7553875, related citations] [Full Text]

  19. Hattersley, A. T., Beards, F., Ballantyne, E., Appleton, M., Harvey, R., Ellard, S. Mutations in the glucokinase gene of the fetus result in reduced birth weight. Nature Genet. 19: 268-270, 1998. [PubMed: 9662401, related citations] [Full Text]

  20. Hattersley, A. T., Turner, R. C., Permutt, M. A., Patel, P., Tanizawa, Y., Chiu, K. C., O'Rahilly, S., Watkins, P. J., Wainscoat, J. S. Linkage of type 2 diabetes to the glucokinase gene. Lancet 339: 1307-1310, 1992. [PubMed: 1349989, related citations] [Full Text]

  21. Heimberg, H., De Vos, A., Moens, K., Quartier, E., Bouwens, L., Pipeleers, D., Van Schaftingen, E., Madsen, O., Schuit, F. The glucose sensor protein glucokinase is expressed in glucagon-producing alpha-cells. Proc. Nat. Acad. Sci. 93: 7036-7041, 1996. [PubMed: 8692940, related citations] [Full Text]

  22. Hofmeister-Brix, A., Kollmann, K., Langer, S., Schultz, J., Lenzen, S., Baltrusch, S. Identification of the ubiquitin-like domain of midnolin as a new glucokinase interaction partner. J. Biol. Chem. 288: 35824-35839, 2013. [PubMed: 24187134, images, related citations] [Full Text]

  23. Inoue, M., Sakuraba, Y., Motegi, H., Kubota, N., Toki, H., Matsui, J., Toyoda, Y., Miwa, I., Terauchi, Y., Kadowaki, T., Shigeyama, Y., Kasuga, M. A series of maturity onset diabetes of the young, type 2 (MODY2) mouse models generated by a large-scale ENU mutagenesis program. Hum. Molec. Genet. 13: 1147-1157, 2004. [PubMed: 15102714, related citations] [Full Text]

  24. Johansen, A., Ek, J., Mortensen, H. B., Pedersen, O., Hansen, T. Half of clinically defined maturity-onset diabetes of the young patients in Denmark do not have mutations in HNF4A, GCK, and TCF1. J. Clin. Endocr. Metab. 90: 4607-4614, 2005. [PubMed: 15928245, related citations] [Full Text]

  25. Kassem, S., Bhandari, S., Rodriguez-Bada, P., Motaghedi, R., Heyman, M., Garcia-Gimeno, M. A., Cobo-Vuilleumier, N., Sanz, P., Maclaren, N. K., Rahier, J., Glaser, B., Cuesta-Munoz, A. L. Large islets, beta-cell proliferation, and a glucokinase mutation. New Eng. J. Med. 362: 1348-1350, 2010. Note: Erratum: New Eng. J. Med. 363: 2178 only, 2010. [PubMed: 20375417, related citations] [Full Text]

  26. Katagiri, H., Asano, T., Ishihara, H., Inukai, K., Anai, M., Miyazaki, J., Tsukuda, K., Kikuchi, M., Yazaki, Y., Oka, Y. Nonsense mutation of glucokinase gene in late-onset non-insulin-dependent diabetes mellitus. Lancet 340: 1316-1317, 1992. [PubMed: 1360036, related citations] [Full Text]

  27. Kim, J. Y., Hwang, J.-Y., Lee, D. Y., Song, E. H., Park, K. J., Kim, G. H., Jeong, E. A., Lee, Y. J., Go, M. J., Kim, D. J., Lee, S. S., Kim, B.-J., Song, J., Roh, G. S., Gao, B., Kim, W.-H. Chronic ethanol consumption inhibits glucokinase transcriptional activity by Atf3 and triggers metabolic syndrome in vivo. J. Biol. Chem. 289: 27065-27079, 2014. [PubMed: 25074928, images, related citations] [Full Text]

  28. Marz, W., Nauck, M., Hoffmann, M. M., Nagel, D., Boehm, B. O., Koenig, W., Rothenbacher, D., Winkelmann, B. R. G(-30)A polymorphism in the pancreatic promoter of the glucokinase gene associated with angiographic coronary artery disease and type 2 diabetes mellitus. Circulation 109: 2844-2849, 2004. [PubMed: 15173029, related citations] [Full Text]

  29. Matschinsky, F., Liang, Y., Kesavan, P., Wang, L., Froguel, P., Velho, G., Cohen, D., Permutt, M. A., Tanizawa, Y., Jetton, T. L., Niswender, K., Magnuson, M. A. Glucokinase as pancreatic beta-cell glucose sensor and diabetes gene. J. Clin. Invest. 92: 2092-2098, 1993. [PubMed: 8227324, related citations] [Full Text]

  30. Matschinsky, F. M. Glucokinase as glucose sensor and metabolic signal generator in pancreatic beta-cells and hepatocytes. Diabetes 39: 647-652, 1990. [PubMed: 2189759, related citations] [Full Text]

  31. Matsutani, A., Janssen, R., Donis-Keller, H., Permutt, M. A. A polymorphic (CA)n repeat element maps the human glucokinase gene (GCK) to chromosome 7p. Genomics 12: 319-325, 1992. [PubMed: 1740341, related citations] [Full Text]

  32. McCarthy, M. Weighing in on diabetes risk. Nature Genet. 19: 209-210, 1998. [PubMed: 9662384, related citations] [Full Text]

  33. Mirshahi, U. L., Colclough, K., Wright, C. F., Wood, A. R., Beaumont, R. N., Tyrrell, J., Laver, T. W., Stahl, R., Golden, A., Goehringer, J. M, Geisinger-Regeneron DiscovEHR Collaboration, Frayling, T. F., Hattersley, A. T., Carey, D. J., Weedon, M. N., Patel, K. A. Reduced penetrance of MODY-associated HNF1A/HNF4A variants but not GCK variants in clinically unselected cohorts. Am. J. Hum. Genet. 109: 2018-2028, 2022. [PubMed: 36257325, images, related citations] [Full Text]

  34. Mishra, S. K., Helms, C., Dorsey, D., Permutt, M. A., Donis-Keller, H. A 2-cM genetic linkage map of human chromosome 7p that includes 47 loci. Genomics 12: 326-334, 1992. [PubMed: 1740342, related citations] [Full Text]

  35. Neel, J. V., Weder, A. B., Julius, S. Type II diabetes, essential hypertension, and obesity as 'syndromes of impaired genetic homeostasis': the 'thrifty genotype' hypothesis enters the 21st century. Perspect. Biol. Med. 42: 44-74, 1998. [PubMed: 9894356, related citations] [Full Text]

  36. Neel, J. V. Diabetes mellitus: a 'thrifty' genotype rendered detrimental by 'progress'? Am. J. Hum. Genet. 14: 353-362, 1962. [PubMed: 13937884, related citations]

  37. Njolstad, P. R., Sovik, O., Cuesta-Munoz, A., Bjorkhaug, L., Massa, O., Barbetti, F., Undlien, D. E., Shiota, C., Magnuson, M. A., Molven, A., Matschinsky, F. M., Bell, G. I. Neonatal diabetes mellitus due to complete glucokinase deficiency. New Eng. J. Med. 344: 1588-1592, 2001. [PubMed: 11372010, related citations] [Full Text]

  38. Permutt, M. A., Chiu, K. C., Tanizawa, Y. Glucokinase and NIDDM: a candidate gene that paid off. Diabetes 41: 1367-1372, 1992. [PubMed: 1397713, related citations] [Full Text]

  39. Pinterova, D., Ek, J., Kolostova, K., Pruhova, S., Novota, P., Romzova, M., Feigerlova, E., Cerna, M., Lebl, J., Pedersen, O., Hansen, T. Six novel mutations in the GCK gene in MODY patients. (Letter) Clin. Genet. 71: 95-96, 2007. [PubMed: 17204055, related citations] [Full Text]

  40. Randle, P. J. Glucokinase and candidate genes for type 2 (non-insulin-dependent) diabetes mellitus. Diabetologia 36: 269-275, 1993. [PubMed: 8386682, related citations] [Full Text]

  41. Rowe, R. E., Wapelhorst, B., Bell, G. I., Risch, N., Spielman, R. S., Concannon, P. Linkage and association between insulin-dependent diabetes mellitus (IDDM) susceptibility and markers near the glucokinase gene on chromosome 7. Nature Genet. 10: 240-245, 1995. [PubMed: 7663523, related citations] [Full Text]

  42. Shen, Y., Cai, M., Liang, H., Wang, H., Weng, J. Insight into the biochemical characteristics of a novel glucokinase gene mutation. Hum. Genet. 129: 231-238, 2011. [PubMed: 21104275, related citations] [Full Text]

  43. Stoffel, M., Bell, K. L., Blackburn, C. L., Powell, K. L., Seo, T. S., Takeda, J., Vionnet, N., Xiang, K.-S., Gidh-Jain, M., Pilkis, S. J., Ober, C., Bell, G. I. Identification of glucokinase mutations in subjects with gestational diabetes mellitus. Diabetes 42: 937-940, 1993. [PubMed: 8495817, related citations] [Full Text]

  44. Stoffel, M., Froguel, P., Takeda, J., Zouali, H., Vionnet, N., Nishi, S., Weber, I. T., Harrison, R. W., Pilkis, S. J., Lesage, S., Vaxillaire, M., Velho, G., Sun, F., Iris, F., Passa, P., Cohen, D., Bell, G. I. Human glucokinase gene: isolation, characterization, and identification of two missense mutations linked to early-onset non-insulin-dependent (type 2) diabetes mellitus. Proc. Nat. Acad. Sci. 89: 7698-7702, 1992. Note: Erratum: Proc. Nat. Acad Sci. 89: 10562 only, 1992. [PubMed: 1502186, related citations] [Full Text]

  45. Stoffel, M., Patel, P., Lo, Y.-M. D., Hattersley, A. T., Lucassen, A. M., Page, R., Bell, J. I., Bell, G. I., Turner, R. C., Wainscoat, J. S. Missense glucokinase mutation in maturity-onset diabetes of the young and mutation screening in late-onset diabetes. Nature Genet. 2: 153-156, 1992. [PubMed: 1303265, related citations] [Full Text]

  46. Stone, L. M., Kahn, S. E., Fujimoto, W. Y., Deeb, S. S., Porte, D. Jr. A variation at position -30 of the beta-cell glucokinase gene promoter is associated with reduced beta-cell function in middle-aged Japanese-American men. Diabetes 45: 422-428, 1996. [PubMed: 8603762, related citations] [Full Text]

  47. Sun, F., Knebelmann, B., Pueyo, M. E., Zouali, H., Lesage, S., Vaxillaire, M., Passa, P., Cohen, D., Velho, G., Antignac, C., Froguel, P. Deletion of the donor splice site of intron 4 in the glucokinase gene causes maturity-onset diabetes of the young. J. Clin. Invest. 92: 1174-1180, 1993. [PubMed: 8376578, related citations] [Full Text]

  48. Tanizawa, Y., Koranyi, L. I., Welling, C. M., Permutt, M. A. Human liver glucokinase gene: cloning and sequence determination of two alternatively spliced cDNAs. Proc. Nat. Acad. Sci. 88: 7294-7297, 1991. [PubMed: 1871135, related citations] [Full Text]

  49. Terauchi, Y., Takamoto, I., Kubota, N., Matsui, J., Suzuki, R., Komeda, K., Hara, A., Toyoda, Y., Miwa, I., Aizawa, S., Tsutsumi, S., Tsubamoto, Y., Hashimoto, S., Eto, K., Nakamura, A., Noda, M., Tobe, K., Aburatani, H., Nagai, R., Kadowaki, T. Glucokinase and IRS-2 are required for compensatory beta cell hyperplasia in response to high-fat diet-induced insulin resistance. J. Clin. Invest. 117: 246-257, 2007. [PubMed: 17200721, images, related citations] [Full Text]

  50. Thornton, P. S., Satin-Smith, M. S., Herold, K., Glaser, B., Chiu, K. C., Nestorowicz, A., Permutt, M. A., Baker, L., Stanley, C. A. Familial hyperinsulinism with apparent autosomal dominant inheritance: clinical and genetic differences from the autosomal recessive variant. J. Pediat. 132: 9-14, 1998. [PubMed: 9469993, related citations] [Full Text]

  51. Velho, G., Blanche, H., Vaxillaire, M., Bellane-Chantelot, C., Pardini, V. C., Timsit, J., Passa, P., Deschamps, I., Robert, J.-J., Weber, I. T., Marotta, D., Pilkis, S. J., Lipkind, G. M., Bell, G. I., Froguel, P. Identification of 14 new glucokinase mutations and description of the clinical profile of 42 MODY-2 families. Diabetologia 40: 217-224, 1997. [PubMed: 9049484, related citations] [Full Text]

  52. Velho, G., Froguel, P., Clement, K., Pueyo, M. E., Rakotoambinina, B., Zouali, H., Passa, P., Cohen, D., Robert, J.-J. Primary pancreatic beta-cell secretory defect caused by mutations in glucokinase gene in kindreds of maturity onset diabetes of the young. Lancet 340: 444-448, 1992. [PubMed: 1354782, related citations] [Full Text]

  53. Velho, G., Petersen, K. F., Perseghin, G., Hwang, J.-H., Rothman, D. L., Pueyo, M. E., Cline, G. W., Froguel, P., Shulman, G. I. Impaired hepatic glycogen synthesis in glucokinase-deficient (MODY-2) subjects. J. Clin. Invest. 98: 1755-1761, 1996. [PubMed: 8878425, related citations] [Full Text]

  54. Vionnet, N., Stoffel, M., Takeda, J., Yasuda, K., Bell, G. I., Zouali, H., Lesage, S., Velho, G., Iris, F., Passa, P., Froguel, P., Cohen, D. Nonsense mutation in the glucokinase gene causes early-onset non-insulin-dependent diabetes mellitus. Nature 356: 721-722, 1992. [PubMed: 1570017, related citations] [Full Text]

  55. Vits, L., Beckers, D., Craen, M., de Beaufort, C., Vanfleteren, E., Dahan, K., Nollet, A., Vanhaverbeke, G., Imschoot, S. V., Bourguignon, J.-P., Beauloye, V., Storm, K., Massa, G., Giri, M., Nobels, F., De Schepper, J., Rooman, R., Van den Bruel, A., Mathieu, C., Wuyts, W. Identification of novel and recurrent glucokinase mutations in Belgian and Luxembourg maturity onset diabetes of the young patients. (Letter) Clin. Genet. 70: 355-359, 2006. [PubMed: 16965331, related citations] [Full Text]


Ada Hamosh - updated : 01/17/2023
Patricia A. Hartz - updated : 02/14/2018
Patricia A. Hartz - updated : 6/24/2014
Marla J. F. O'Neill - updated : 11/2/2012
Patricia A. Hartz - updated : 4/15/2010
Marla J. F. O'Neill - reorganized : 4/14/2010
Marla J. F. O'Neill - updated : 4/14/2010
Ada Hamosh - updated : 1/15/2010
Victor A. McKusick - updated : 5/23/2007
Marla J. F. O'Neill - updated : 4/17/2007
Marla J. F. O'Neill - updated : 3/9/2007
Victor A. McKusick - updated : 11/28/2006
Cassandra L. Kniffin - updated : 11/2/2006
John A. Phillips, III - updated : 10/19/2006
George E. Tiller - updated : 8/31/2006
Marla J. F. O'Neill - updated : 3/17/2006
Marla J. F. O'Neill - updated : 1/25/2006
Victor A. McKusick - updated : 11/19/2003
Ada Hamosh - updated : 9/16/2003
Ada Hamosh - updated : 8/5/2003
Ada Hamosh - updated : 10/18/2001
Ada Hamosh - updated : 8/23/2001
Victor A. McKusick - updated : 6/25/2001
Victor A. McKusick - updated : 2/2/1999
Victor A. McKusick - updated : 6/25/1998
Victor A. McKusick - updated : 4/15/1998
Jennifer P. Macke - updated : 9/2/1997
Creation Date:
Victor A. McKusick : 2/1/1992
alopez : 01/17/2023
alopez : 12/01/2020
carol : 09/02/2020
carol : 07/14/2020
alopez : 04/30/2020
mgross : 02/14/2018
alopez : 08/04/2016
carol : 08/12/2015
carol : 8/12/2015
mgross : 6/24/2014
mgross : 10/7/2013
mgross : 10/7/2013
terry : 4/4/2013
carol : 11/6/2012
terry : 11/2/2012
terry : 7/5/2012
carol : 12/15/2010
alopez : 4/22/2010
mgross : 4/16/2010
terry : 4/15/2010
carol : 4/14/2010
carol : 4/14/2010
wwang : 2/3/2010
alopez : 1/27/2010
alopez : 1/26/2010
terry : 1/15/2010
terry : 9/11/2009
joanna : 9/4/2009
alopez : 5/29/2007
terry : 5/23/2007
wwang : 4/17/2007
wwang : 3/13/2007
terry : 3/9/2007
alopez : 12/8/2006
terry : 11/28/2006
carol : 11/3/2006
ckniffin : 11/2/2006
alopez : 10/19/2006
alopez : 8/31/2006
carol : 3/17/2006
carol : 3/16/2006
carol : 3/16/2006
carol : 3/16/2006
carol : 3/16/2006
wwang : 2/2/2006
terry : 1/25/2006
carol : 12/3/2004
carol : 12/3/2004
tkritzer : 12/30/2003
tkritzer : 12/18/2003
tkritzer : 11/24/2003
terry : 11/19/2003
alopez : 9/16/2003
alopez : 8/7/2003
terry : 8/5/2003
terry : 6/27/2002
carol : 10/18/2001
carol : 8/23/2001
mcapotos : 7/6/2001
mcapotos : 6/29/2001
terry : 6/25/2001
carol : 12/26/2000
mgross : 3/16/1999
carol : 2/15/1999
carol : 2/15/1999
terry : 2/2/1999
dkim : 12/10/1998
alopez : 6/29/1998
terry : 6/25/1998
carol : 4/20/1998
terry : 4/15/1998
mark : 10/20/1997
mark : 10/17/1997
alopez : 10/7/1997
alopez : 10/7/1997
alopez : 10/6/1997
terry : 11/25/1996
terry : 11/12/1996
terry : 11/1/1996
terry : 9/17/1996
marlene : 8/15/1996
terry : 11/2/1995
mark : 8/18/1995
mimadm : 9/24/1994
jason : 7/14/1994
carol : 5/10/1994
carol : 12/7/1993

* 138079

GLUCOKINASE; GCK


Alternative titles; symbols

GK; GLK
HEXOKINASE 4; HK4
LIVER GLUCOKINASE; LGLK


HGNC Approved Gene Symbol: GCK

SNOMEDCT: 237604008, 44054006;   ICD10CM: E11;  


Cytogenetic location: 7p13     Genomic coordinates (GRCh38): 7:44,143,213-44,189,439 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
7p13 Diabetes mellitus, noninsulin-dependent, late onset 125853 Autosomal dominant 3
Diabetes mellitus, permanent neonatal 1 606176 Autosomal recessive 3
Hyperinsulinemic hypoglycemia, familial, 3 602485 Autosomal dominant 3
MODY, type II 125851 Autosomal dominant 3

TEXT

Description

The phosphorylation of glucose at the sixth carbon position is the first step in glycolysis. The reaction is catalyzed by a family of enzymes called hexokinases, types I (142600) through IV (glucokinase). Glucokinase (GCK; EC 2.7.1.1) is a structurally and functionally unique member of this family. Glucokinase is expressed only in mammalian liver and pancreatic islet beta cells. Because of its unique functional characteristics, the enzyme plays an important regulatory role in glucose metabolism. The rate of glucose metabolism in liver and pancreas is a function of the activity of the enzyme (summary by Matsutani et al., 1992).


Cloning and Expression

Using rat islet glucokinase cDNA to screen a human islet cDNA library, followed by screening a human liver cDNA library and RT-PCR of liver RNA, Tanizawa et al. (1991) cloned 2 GCK splice variants, which they called LGLK1 and LGLK2. LGLK1 encodes a deduced 464-amino acid protein with a calculated molecular mass of 52 kD. LGLK2 encodes a deduced 466-amino acid protein with 16 different N-terminal amino acids compared with LGLK1. LGLK2 shares 98% identity with rat islet glucokinase. Northern blot analysis detected a 2.8-kb transcript in liver total RNA. In vitro translation of LGLK1 cDNA resulted in proteins with apparent molecular masses of 52 and 48 kD by SDS-PAGE. Proteins of similar masses were translated from LGLK2.


Gene Structure

Stoffel et al. (1992) determined that the GCK gene contains 12 exons. Alternative promoters and alternative splicing encode different-sized proteins.


Mapping

Matsutani et al. (1992) identified a region of compound dinucleotide repeats located approximately 10 kb 3-prime to the GCK gene and used oligonucleotide primers and PCR amplification to demonstrate a number of alleles created by this repeat region. The variable numbers of CT and CA repeats represented altogether 6 alleles that range in length from +10 to -15 nucleotides compared to the most common (Z) allele. Two alleles, Z+10 and Z-15, appeared to be unique to American blacks, while a Z+6 allele was observed only in the Caucasian population studied. Using the PCR assay, they localized the human glucokinase gene to chromosome 7 in a panel of rodent/human somatic cell lines. Genetic analysis in CEPH pedigrees placed the dinucleotide repeat element, and thereby GCK, on 7p between TCRG (see 186970) and the RFLP locus D7S57.

Mishra et al. (1992) placed GCK between D7S57 proximally and D7S65 distally. It was estimated to lie about 4.7 cM from D7S65; D7S65 was found to be about 3.4 cM from TCRG, which is located at 7p15-p14. (Mishra et al. (1992) gave the location of TCRG as 7p15 in their Figure 2.)


Gene Function

Byrne et al. (1994) studied pancreatic beta-cell function in 4 males and 2 females from kindreds with known GCK mutations and in 6 controls pair-matched for age, weight, and sex. All of the GCK subjects with GCK mutations were found to have elevated fasting and postprandial glucose levels in comparison to the 6 controls. Insulin secretion rates (ISRs) were estimated. The results supported a key role for GCK in determining the in vivo glucose/ISR dose-response relationships and defined the alterations in beta-cell responsiveness that occur in subjects with GCK mutations.

Heimberg et al. (1996) reported the expression of glucokinase in rat glucagon-producing islet alpha cells, which are negatively regulated by glucose. Purified rat alpha cells expressed glucokinase mRNA and protein with the same transcript length, nucleotide sequence, and immunoreactivity as the beta-cell isoform. Glucokinase activity accounted for more than 50% of glucose phosphorylation in extracts of alpha cells and for more than 90% of glucose utilization in intact cells. A glucagon-producing tumor also contained glucokinase mRNA, protein, and enzymatic activity. The data indicated that glucokinase may serve as a metabolic glucose sensor in pancreatic alpha cells and, hence, mediate a mechanism for direct regulation of glucagon release by extracellular glucose. Since the alpha cells do not express GLUT2 (138160), Heimberg et al. (1996) suggested that glucose sensing does not necessarily require the coexpression of GLUT2 and glucokinase.

On the basis of studies in 7 glucokinase-deficient subjects with normal glycosylated hemoglobin and 12 control subjects using (13)C nuclear magnetic spectroscopy during a day in which 3 isocaloric mixed meals were ingested, Velho et al. (1996) observed results suggesting that in addition to the altered beta-cell function, abnormalities in liver glycogen metabolism play an important role in the pathogenesis of hypoglycemia in patients with maturity-onset diabetes of the young type 2 (MODY2; 125851). Average fasting hepatic glycogen content was similar in both groups and increased in both after the meals with a continuous pattern throughout the day. However, the net increment in hepatic glycogen content after each meal was 30 to 60% lower in glucokinase-deficient than in control subjects. Glucokinase-deficient subjects had decreased net accumulation of hepatic glycogen and relatively augmented hepatic gluconeogenesis after meals.

Danial et al. (2003) undertook a proteomic analysis to assess whether BAD (603167) might participate in mitochondrial physiology. In liver mitochondria, BAD resides in a functional holoenzyme complex together with protein kinase A (see 176911) and protein phosphatase-1 (PP1; see 176875) catalytic units, WAVE1 (605035) as an A kinase-anchoring protein, and glucokinase. Using mitochondria from hepatocytes of Bad-deficient mice, Danial et al. (2003) demonstrated that BAD is required to assemble the complex, the lack of which results in diminished mitochondria-based glucokinase activity and blunted mitochondrial respiration in response to glucose. Glucose deprivation results in dephosphorylation of BAD, and BAD-dependent cell death. Moreover, Danial et al. (2003) demonstrated that the phosphorylation status of BAD helps regulate glucokinase activity. Mice deficient in BAD or bearing a nonphosphorylatable BAD (3SA) mutant (Datta et al., 2002) display abnormal glucose homeostasis, including profound defects in glucose tolerance. Danial et al. (2003) concluded that this combination of proteomics, genetics, and physiology indicates an unanticipated role for BAD in integrating pathways of glucose metabolism and apoptosis.

Using a yeast 2-hybrid assay and other protein interaction assays, Hofmeister-Brix et al. (2013) found that Midn (606700), via its the ubiquitin-like domain, bound glucokinase in rat and mouse pancreatic beta cells and cell lines. Binding was strongest at low glucose concentration, when glucokinase exists mainly in its inactive super-open-to-open conformation. Overexpression of fluorescence-tagged Midn inhibited glucokinase activity, reduced the affinity of glucokinase for glucose, and reduced glucose-induced insulin secretion in rat pancreatic MIN6 cells.

Using mouse models and the rat INS-1 pancreatic beta cell line, Kim et al. (2014) found that chronic ethanol consumption induced activating transcription factor-3 (ATF3; 603148), which then inhibited Gck transcriptional activity. Atf3 directly bound to a putative ATF/CREB site in the Gck promoter. Atf3 also counteracted the positive effect of Pdx1 (600733) on Gck transcriptional activity.


Molecular Genetics

Maturity-Onset Diabetes of the Young, Type 2

In 16 French families with maturity-onset diabetes of the young (MODY2; 125851), Froguel et al. (1992) found linkage of the disease with GCK. There was statistically significant evidence of genetic heterogeneity, with an estimated 45 to 95% of the 16 families showing linkage to glucokinase. Because glucokinase is a key enzyme of blood glucose homeostasis, the results suggested a pathogenetic connection. The relationship was clinched by the demonstration by Vionnet et al. (1992) of a nonsense mutation in the GCK gene (138079.0001).

Velho et al. (1992) studied pancreatic beta-cell secretory function in 9 patients from 4 GCK-linked MODY kindreds. They found that beta-cell secretory response to continuous glucose stimulus during a hyperglycemic glucose clamp was significantly reduced. This finding was different from that for noninsulin-dependent diabetes mellitus with late age of onset or MODY not linked to GCK. Fasting plasma insulin and C-peptide levels in patients were inappropriately low in relation to concomitant plasma glucose level. Furthermore, during a hyperinsulinemic euglycemic clamp, endogenous insulin secretion at euglycemia was suppressed in patients but not in controls. Velho et al. (1992) interpreted these results as suggesting that mutant GCK leads to chronic hyperglycemia by raising the threshold of circulating glucose levels which induces insulin secretion. This was the first demonstration of a primary pancreatic secretory defect associated with a form of NIDDM.

Froguel et al. (1993) detected 16 mutations of the GCK gene in 18 of 32 families with MODY. An explanation for the way a mutation in glucokinase might cause diabetes follows from the work of Matschinsky (1990), who developed the concept that glucokinase acts as the glucose sensor of beta cells. According to this concept, the rates of glucose metabolism and insulin secretion are closely linked, with both being determined by the plasma glucose concentration. In beta cells and hepatocytes, the rate of glucose metabolism is determined by the rate of glucose phosphorylation, which is catalyzed by glucokinase. Beta cells and hepatocytes also contain glucose transporter-2 (GLUT2; 138160), an insulin-independent cellular protein that mediates the transport of glucose into cells. The capacity of GLUT2 to transport glucose is very high, facilitating rapid equilibrium between extracellular and intracellular glucose. Thus, in effect, the extracellular glucose concentrations are sensed by glucokinase.

Randle (1993) provided a review of glucokinase and candidate genes for type II MODY. Matschinsky et al. (1993) also reviewed glucokinase as the pancreatic beta-cell glucose sensor and 'diabetes gene.' They tabulated 16 separate mutations in the GCK gene found in families with MODY.

Johansen et al. (2005) examined the prevalence and nature of mutations in the 3 common MODY genes HNF4A (600281), GCK, and TCF1 (142410) in Danish patients with a clinical diagnosis of MODY and determined metabolic differences in probands with and without mutations in HNF4A, GCK, and TCF1. They identified 29 different mutations in 38 MODY families. Fifteen of the mutations were novel. The variants segregated with diabetes within the families, and none of the variants were found in 100 normal Danish chromosomes. Their findings suggested a relative prevalence of 3% of MODY1 (125850) (2 different mutations in 2 families), 10% of MODY2 (7 in 8), and 36% of MODY3 (600496) (21 in 28) among Danish kindred clinically diagnosed as MODY. No significant differences in biochemical and anthropometric measurements were observed at baseline examinations. Forty-nine percent of the families carried mutations in the 3 examined MODY genes.

Vits et al. (2006) identified 19 different GCK mutations, including 11 novel mutations (see, e.g., 138079.0014), in 33 (26.6%) of 124 Belgian probands with MODY.

Pinterova et al. (2007) screened the GCK gene in 92 Czech probands fulfilling classic MODY criteria and identified 15 different missense mutations in 27 (29%) patients; the mutations were not found in 50 unrelated healthy Czech individuals. Pinterova et al. (2007) concluded that mutations in GCK are a common cause of MODY in the Czech population.

In a mainland Chinese family with a clinical profile similar to that of previously reported MODY2 families, Shen et al. (2011) analyzed the GCK gene and identified a heterozygous missense mutation (E339K; 138079.0016) that segregated with the disease and was not found in 200 controls. Functional analysis indicated that the mutation inactivated enzyme kinetics and severely impaired GCK protein stability.

The variable phenotype observed with mutations in the GCK gene includes gestational diabetes mellitus. This fact prompted Stoffel et al. (1993) to screen a group of women with gestational diabetes who also had a first-degree relative with diabetes mellitus for the presence of GCK mutations. Among 40 subjects, they identified 2 with heterozygous mutations (138079.0007-138070.0008), suggesting a prevalence of approximately 5%. Extrapolating from this result, the prevalence of glucokinase-deficient MODY among Americans may be approximately 1 in 2,500.

Fetal insulin secretion in response to maternal glycemia plays a key role in fetal growth, and adult insulin secretion is a primary determinant of glucose tolerance. Hattersley et al. (1998) hypothesized that a defect in the sensing of glucose by the pancreas, caused by a heterozygous mutation in the glucokinase gene, could reduce fetal growth and birth weight in addition to causing hyperglycemia after birth. In 58 offspring, where one parent had a glucokinase mutation, the inheritance of a glucokinase mutation by the fetus resulted in a mean reduction of birth weight of 533 g (P = 0.002). In 19 of 21 sib pairs discordant for the presence of a glucokinase mutation, the child with the mutation had a lower birth weight, with a mean difference of 521 g (P = 0.0002). Maternal hyperglycemia due to a glucokinase mutation resulted in a mean increase in birth weight of 601 g (P = 0.001). The effects of maternal and fetal glucokinase mutations on birth weight were additive. Hattersley et al. (1998) proposed that these changes in birth weight reflect changes in fetal insulin secretion which are influenced directly by the fetal genotype and indirectly, through maternal hyperglycemia, by the maternal genotype.

Dunger et al. (1998) demonstrated an association between the insulin VNTR (variable number tandem repeat) 'locus,' and specifically the VNTR III genotype, and larger size at birth. McCarthy (1998) discussed the significance of these observations in relation to the 'thrifty genotype' hypothesis originally proposed by Neel (1962) and updated by Neel et al. (1998).

Permanent Neonatal Diabetes Mellitus 1

Njolstad et al. (2001) described 2 patients in whom complete deficiency of glucokinase caused permanent neonatal-onset diabetes (PNDM1; 606176). One patient was homozygous for an M210K mutation (138079.0010); the other was homozygous for a T228M mutation (138079.0003). Both patients showed total absence of basal insulin release.

Hyperinsulinemic Hypoglycemia

In 'family 3' studied by Thornton et al. (1998), in which affected members had hyperinsulinemic hypoglycemia (see HHF3, 602485), Glaser et al. (1998) identified heterozygosity for an activating mutation (138079.0009) in the GCK gene.

In a 14-year-old obese boy with a history of neonatal hypoglycemia treated with diazoxide, who was experiencing hypoglycemic episodes associated with seizures and unconsciousness, Christesen et al. (2002) identified heterozygosity for an activating mutation in the GCK gene (138079.0012). The boy's normal-weight mother, who had asymptomatic fasting hypoglycemia, carried the same mutation. Christesen et al. (2002) noted the striking contrast in clinical presentation between the mother and son with the same mutation, and recalled a similar situation in the family reported by Glaser et al. (1998), in which the proband was obese and severely hyperinsulinemic, whereas his sister who carried the same mutation was of normal weight, had relatively low insulin levels, and milder clinical symptoms.

In a Finnish woman with severe hyperinsulinemic hypoglycemia from birth, who had severe mental retardation and died at age 29 still having hypoglycemic seizures, Cuesta-Munoz et al. (2004) identified heterozygosity for a de novo activating mutation in the GCK gene (138079.0013).

Kassem et al. (2010) reported a young girl with severe neonatal hypoglycemia due to a missense mutation in the GCK gene (138079.0015) in whom mean islet cell areas in both the head and the tail of the pancreas were significantly larger than those of 5 age-matched controls and those of 2 age-matched patients with diffuse hypoglycemia due to ABCC8 mutations (600509; see HHF1, 256450). Noting a previously reported HHF3 patient in whom quantitative histologic analysis of pancreatic specimens showed a similar increase in the mean islet profile (Cuesta-Munoz et al. (2004)), Kassem et al. (2010) suggested that histologic findings in infants with hyperinsulinemic hypoglycemia might differ according to the genetic cause of the condition.

Penetrance of GCK Mutations in Diabetes

Mirshahi et al. (2022) comprehensively assessed the penetrance and prevalence of pathogenic variants in HNF1A (142410), HNF4A (600281), and GCK that account for more than 80% of monogenic diabetes. Mirshahi et al. (2022) analyzed clinical and genetic data from 1,742 clinically referred probands, 2,194 family members, clinically unselected individuals from a US health system-based cohort of 132,194 individuals, and a UK population-based cohort of 198,748 individuals, and found that 1 in 1,500 individuals harbor a pathogenic variant in one of these genes. The penetrance of pathogenic GCK variants was similar (89 to 97%) across all cohorts. The penetrance of diabetes for HNF1A and HNF4A pathogenic variants was substantially lower in the clinically unselected individuals compared to clinically referred probands and was dependent on the setting (32% in the population, 49% in the health system cohort, 86% in a family member, and 98% in probands for HNF1A). The relative risk of diabetes was similar across the clinically unselected cohorts, highlighting the role of environment/ other genetic factors. The authors suggested that for HNF1A and HNF4A, genetic interpretation and counseling should be tailored to the setting in which a pathogenic monogenic variant was identified. GCK is an exception with near-complete penetrance in all settings.

Other Associations

Rowe et al. (1995) evaluated 35 microsatellite marker loci on human chromosome 7 for linkage to insulin-dependent diabetes mellitus (IDDM) in 339 affected sib-pair families. Increased sharing of parental haplotypes in affected sib pairs was detected for 2 microsatellite markers flanking GCK. Preferential transmission of alleles to affected offspring was observed at one of these marker loci, GCK3, indicating linkage disequilibrium between the marker and the disease susceptibility locus. This combination of linkage and disease association suggested to Rowe et al. (1995) that glucokinase, or a gene in the vicinity, plays an important role in IDDM susceptibility. (Rowe et al. (1995) used the designations GCK1, GCK2, and GCK3 for 3 modestly polymorphic microsatellite markers near the GCK gene (Concannon, 1995). These are not symbols for separate glucokinase genes; there is only a single GCK gene on 7p, although as a result of alternative, tissue-specific promoters, there are islet- and liver-specific forms that differ in their amino-terminal sequences.)

Stone et al. (1996) used SSCP analysis to determine whether a G-to-A variant at position -30 of the beta-cell promoter of the GCK gene is associated with impaired beta-cell function. The variant was observed more frequently in Japanese-American men with impaired glucose tolerance than in Japanese-American men with normal glucose tolerance. Beta-cell function was assessed using the ratio of the incremental response in immunoreactive insulin to that of glucose during the first 30 minutes of the oral glucose tolerance test. They concluded that the -30 promoter variant is associated with reduced beta-cell function in middle-aged Japanese-American men and may contribute to the high risk of abnormal glucose tolerance in this population. They noted that the polymorphism has been observed in other populations.

Marz et al. (2004) analyzed the GCK -30G-A variant in 2,567 patients with and in 731 individuals without coronary artery disease (CAD; see CHDS1, 607339) by angiography and found that the A allele of the pancreatic GCK promoter increased the risk of CAD in individuals both with and without type II diabetes mellitus (ORs, 1.92 and 1.27, respectively). The A allele was also associated with an increased prevalence of type II diabetes mellitus, particularly among CAD patients.

For discussion of an association between variation in the GCK gene and fasting plasma glucose levels, see FGQTL2 (613219).


Genotype/Phenotype Correlations

Gloyn (2003) stated that given the central role of glucokinase in the regulation of insulin release, it is understandable that mutations in the GCK gene can cause both hyper- and hypoglycemia. Heterozygous inactivating mutations in GCK cause MODY, characterized by mild hyperglycemia, which is present at birth, but is often only detected later in life. Homozygous inactivating GCK mutations result in a more severe phenotype, presenting at birth as permanent neonatal diabetes mellitus. Several heterozygous activating GCK mutations causing hypoglycemia had been reported as instances of familial hyperinsulinism (see HHF3, 602485). Gloyn (2003) tabulated 195 mutations in the GCK gene, found in 285 families. There were no common mutations and the mutations were distributed throughout the gene. Mutations causing hypoglycemia are located in various exons in a discrete region of the protein termed the heterotropic allosteric activator site. Patients with MODY type 2 (due to GCK mutations) rarely need pharmacologic treatment and most are managed on diet alone. During pregnancy, some women with GCK mutations are treated with insulin, which, due to its role in fetal growth, can result in babies that are large for gestational age if they do not also have the GCK mutation (Hattersley et al., 1998). Microvascular complications are rare in MODY2. These patients do not need to be followed in a diabetes clinic, but should have an HbA1c checked annually. Some patients with hyperinsulinism due to GCK mutations may not require pharmacologic intervention and may be able to control their symptoms by eating regularly. Patients with hyperinsulinism due to mutation in the ABCC8 gene (600509) or the KCNJ11 gene (600937) respond poorly or are totally refractory to the potassium channel opener diazoxide, which is used to inhibit insulin secretion. Conversely, patients with hyperinsulinism resulting from GCK mutations respond well or tolerated diazoxide (Gloyn et al., 2003).


Animal Model

Using homologous recombination in mouse embryonic stem cells to disrupt the glucokinase gene, Bali et al. (1995) showed that heterozygous mice showed a phenotype similar to MODY. Reduced islet glucokinase activity caused mildly elevated fasting blood glucose levels. Hyperglycemic clamp studies revealed decreased glucose tolerance and abnormal liver glucose metabolism. These findings demonstrated a key role for glucokinase in glucose homeostasis and implicated both islets and liver in the MODY disease. Aizawa et al. (1996) made a detailed study of pancreatic beta-cell function in the isolated pancreatic islets of heterozygous GCK knockout mice. The beta cells of these mice displayed impaired glucose sensitivity, poor discrimination of alpha- and beta-glucose anomers, and normal activity of the ATP-sensitive K+ channel. The mice displayed impaired insulin release after insulin treatment and age-related worsening of glucose tolerance.

Understanding how the beta cells of the pancreatic islet sense an alteration in glucose levels is critical to understanding how glucose homeostasis is maintained. Grupe et al. (1995) noted that glucose sensing is mediated through an increase in the rate of intracellular catabolism of glucose rather than a ligand-receptor interaction. However, it is likely that there is still a need for a rate-limiting step in glucose catabolism to serve the role of glucose sensor. A compelling case can be made for glucokinase fulfilling this role, since phosphorylation by GLK appears to be the rate-limiting step for glucose catabolism in beta cells. To test this possibility and to resolve the relative roles of liver and beta-cell GLK in maintaining glucose levels, Grupe et al. (1995) generated mice completely deficient in GLK and transgenic mice in which GLK is expressed only in beta cells. In mice with only 1 GLK allele, blood glucose levels were elevated and insulin secretion was reduced. GLK-deficient mice died perinatally with severe hyperglycemia. Expression of GLK in beta cells in the absence of expression in the liver is sufficient for survival.

Grimsby et al. (2003) identified a class of antidiabetic agents that act as nonessential, mixed-type glucokinase activators that increase the glucose affinity and maximum velocity of glucokinase. Glucokinase activators augmented both hepatic glucose metabolism and glucose-induced insulin secretion from isolated rodent pancreatic islets, consistent with the expression and function of glucokinase in both cell types. In several rodent models of NIDDM, glucokinase activators lowered blood glucose levels, improved the results of glucose tolerance tests, and increased hepatic glucose uptake.

Using N-ethyl-N-nitrosourea (ENU) mutagenesis, Inoue et al. (2004) generated diabetic mice. The authors screened 9,375 animals and identified 11 mutations in the glucokinase (Gk) gene that were associated with hyperglycemia. Four had previously been found in human MODY2 (125851) patients, and 1 was found previously in a patient with permanent neonatal diabetes mellitus (PNDM; 606176). Some of the Gk mutant lines displayed impaired glucose-responsive insulin secretion, and the mutations had different effects on Gk mRNA levels and/or the stability of the GK protein.

Terauchi et al. (2007) generated mice with haploinsufficiency of beta cell-specific Gck and observed that on a high-fat diet, Gck +/- mice had decreased beta cell replication and insufficient beta cell hyperplasia compared to wildtype mice despite a similar degree of insulin resistance. On a high-fat diet, Gck +/- mouse islets showed decreased insulin receptor substrate-2 (Irs2; 600797) expression compared with wildtype islets. Overexpression of Irs2 in beta cells of Gck +/- mice partially prevented diabetes by increasing beta cell mass. Terauchi et al. (2007) suggested that both GCK and IRS2 are critical for beta cell hyperplasia to occur in response to high-fat diet-induced insulin resistance.


ALLELIC VARIANTS 16 Selected Examples):

.0001   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLU279TER
SNP: rs104894005, gnomAD: rs104894005, ClinVar: RCV000017512

In the process of single-strand conformation polymorphism (SSCP) analysis of exon 7 in a French family with GCK-linked MODY (MODY2; 125851), previously studied by Froguel et al. (1992), Vionnet et al. (1992) demonstrated a G-to-T substitution in codon 279 which changed GAG (glutamic acid) to TAG, an amber termination codon. An individual in this family who had noninsulin-dependent diabetes mellitus (NIDDM) did not show linkage to GCK; moreover, she did not have the nonsense mutation in codon 279. Thus, there were 2 forms of NIDDM in this kindred. Hattersley et al. (1992) likewise found tight linkage of MODY to a macrosatellite polymorphism associated with the GCK locus; peak lod = 4.60 at theta = 0.0. In a second MODY pedigree, they excluded linkage; lod = -7.36 at theta = 0.0.


.0002   TYPE 2 DIABETES MELLITUS

GCK, ARG186TER
SNP: rs104894006, gnomAD: rs104894006, ClinVar: RCV000017513, RCV000516235, RCV002225072, RCV002345246

In a Japanese family, Katagiri et al. (1992) found a correlation between the presence of late-onset noninsulin-dependent diabetes mellitus (125853) or impaired glucose tolerance and a nonsense mutation in exon 5 of the glucokinase gene: a C-to-T transition changing codon 186 from CGA (arginine) to TGA (an amber termination codon). The mutation was predicted to delete 60% of the amino acid residues of the glucokinase derived from the affected allele. The family was detected through a 59-year-old woman who was the twenty-third subject screened. Froguel and Velho (1993) raised the question as to whether the Japanese family may in fact have had MODY, because in their experience in French families carrying mutations in GCK, hyperglycemia often went undiagnosed for a long time. Chiu et al. (1993) likewise questioned whether the disorder in the Japanese families should be termed NIDDM and pointed out that in young patients with glucokinase mutations the degree of hyperglycemia is so mild that values often do not exceed the renal threshold. Therefore, absence of glycosuria cannot be used as a criterion for distinguishing MODY from NIDDM. Permutt et al. (1992) pointed out that structural mutations in the GCK gene are very rare (less than 2%) in American black and Caucasian NIDDM patients.


.0003   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

DIABETES MELLITUS, PERMANENT NEONATAL, 1, INCLUDED
GCK, THR228MET
SNP: rs80356655, gnomAD: rs80356655, ClinVar: RCV000020167, RCV000498792, RCV001269032, RCV002362586, RCV003147295, RCV003147296, RCV003883118

Maturity-Onset Diabetes of the Young, Type 2

Stoffel et al. (1992) stated that DNA polymorphisms in the GCK gene has been shown to be tightly linked to MODY (MODY2; 125851) in approximately 80% of French families. They identified 2 further missense mutations in exon 7 in families with MODY: thr228 to met (T228M) and gly261 to arg (G261R; 138079.0004). A TGC-to-TAC change at codon 228 and a CCC-to-TCC change at codon 261 were responsible. Using computer-assisted modeling and the crystal structure of the related yeast hexokinase B as a simple model for human beta-cell glucokinase, Stoffel et al. (1992) obtained data which suggested to them that mutation of thr228 affects affinity for ATP and mutation of gly261 alters glucose binding. The identification of mutations in glucokinase, a protein that plays an important role in hepatic and beta-cell glucose metabolism, indicates that early-onset noninsulin-dependent diabetes mellitus may be primarily a disorder of carbohydrate metabolism.

Velho et al. (1997) identified the T228M mutation in affected members of a family with glucokinase-related MODY.

Diabetes Mellitus, Permanent Neonatal 1

In an 8-year-old girl of Italian ancestry, Njolstad et al. (2001) identified homozygosity for the T228M mutation in the GCK gene as the cause of neonatal-onset diabetes mellitus (PNDM1; 606176). The child had shown hyperglycemia and marked growth retardation at birth. Her father had impaired glucose tolerance and her mother had impaired fasting glycemia.


.0004   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLY261ARG
SNP: rs104894008, gnomAD: rs104894008, ClinVar: RCV000017515, RCV000426797, RCV001507008, RCV003325940, RCV003445078

Stoffel et al. (1992) identified missense mutations in exon 7 of the GCK gene in families with maturity diabetes of the young (MODY2; 125851): thr228 to met (T228M; 138079.0003) and gly261 to arg (G261R). A TGC-to-TAC change at codon 228 and a CCC-to-TCC change at codon 261 were responsible. Using computer-assisted modeling and the crystal structure of the related yeast hexokinase B as a simple model for human beta-cell glucokinase, Stoffel et al. (1992) obtained data which suggested to them that mutation of thr228 affects affinity for ATP and mutation of gly261 alters glucose binding. The identification of mutations in glucokinase, a protein that plays an important role in hepatic and beta-cell glucose metabolism, indicates that early-onset noninsulin-dependent diabetes mellitus may be primarily a disorder of carbohydrate metabolism.


.0005   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLY299ARG
SNP: rs104894009, ClinVar: RCV000017516, RCV001288185

In a large British pedigree with many cases of MODY (MODY2; 125851) in 5 generations, Stoffel et al. (1992) demonstrated a G-to-C transversion in the GCK gene, which converted codon 266 from glycine to arginine. The mutation also created a HhaI site which allowed them to construct a rapid PCR test for the mutation. Applying this to cases of classic late-onset type 2 diabetes mellitus, they found the same mutation in 1 of 50 patients. All 9 relatives of this patient who had inherited the mutation had type 2 diabetes, although 6 others without the mutation also had diabetes. MODY had not previously been considered in this family.


.0006   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, IVS4DS, 15-BP DEL
SNP: rs1583601110, ClinVar: RCV000992052, RCV002286425, RCV003106086

Sun et al. (1993) analyzed the nucleotide sequence of exon 4 and its flanking intronic regions of the GCK gene in 4 hyperglycemic members of a MODY (MODY2; 125851) family and found a deletion of 15 bp, which removed the T of the GT in the donor splice site of intron 4 and the following 14 basepairs. This deletion resulted in 2 aberrant transcripts: one with exon 5 missing and the other with activation of a cryptic splice site leading to the removal of the last 8 codons of exon 4. This intronic deletion seemed to cause a more severe form of glucose intolerance than did the missense and nonsense mutations previously reported.


.0007   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, SER131PRO
SNP: rs104894010, ClinVar: RCV000017518, RCV002513078, RCV003147297, RCV003147298, RCV003147299

Stoffel et al. (1993) found heterozygosity for a ser131-to-pro mutation in the GCK gene in an obese 31-year-old Puerto Rican woman with gestational diabetes (MODY2; 125851) in her first pregnancy. She reportedly had borderline elevated blood glucose levels at age 26. Stoffel et al. (1993) showed defective enzymatic properties of the enzyme carrying the ser131-to-pro mutation.


.0008   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLU265TER
SNP: rs104894011, ClinVar: RCV000017519, RCV000517435

Stoffel et al. (1993) found a glu265-to-ter mutation in a thin 32-year-old Caucasian woman with gestational diabetes mellitus (MODY2; 125851) who was in her second pregnancy. She reportedly had elevated fasting blood glucose levels since age 16. The subject's mother and 2 sisters had diabetes mellitus treated with diet or oral hypoglycemic agents.


.0009   HYPERINSULINEMIC HYPOGLYCEMIA, FAMILIAL, 3

GCK, VAL455MET
SNP: rs104894012, ClinVar: RCV000017520, RCV002513079

In a 31-year-old man and his 36-year-old sister from the 3-generation 'family 3' with autosomal dominant hyperinsulinemic hypoglycemia (602485) previously studied by Thornton et al. (1998), Glaser et al. (1998) identified a val455-to-met (V455M) mutation in the glucokinase gene. When expressed in vitro, the V455M mutation increased the affinity of glucokinase for glucose, resulting in higher rates of glycolysis at low glucose concentrations and therefore a higher rate of insulin secretion at any plasma glucose concentration. The finding confirmed the importance of glucokinase as a primary regulator of glucose-controlled insulin secretion in beta cells. This mutation was not found in 37 unrelated white families with hyperinsulinism, including 6 with an apparently autosomal dominant form.


.0010   DIABETES MELLITUS, PERMANENT NEONATAL, 1

MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2, INCLUDED
GCK, MET210LYS
SNP: rs80356654, gnomAD: rs80356654, ClinVar: RCV000017521, RCV000190348, RCV001281107, RCV003318493

Njolstad et al. (2001) described an infant girl of Norwegian ancestry with neonatal diabetes mellitus (PNDM1; 606176) that persisted thereafter. The parents were first cousins, and both had glucose intolerance. At 5 years of age, the patient developed epilepsy, probably as a sequela of a neonatal brain abscess. A sister had typical type I diabetes developing at the age of 7 years. The mother had a diagnosis of gestational diabetes at the age of 25 years. The father had impaired fasting glycemia that was treated with diet. After excluding other candidate genes, Njolstad et al. (2001) found that the child was homozygous for a T-to-A transversion (ATG-AAG) at nucleotide 629 in exon 6 of the GCK gene, resulting in a met210-to-lys mutation (M210K). Her parents and sister were heterozygous for the mutation, which cosegregated with diabetes or hyperglycemia in other members of the family. Thus, in this family, heterozygosity caused GCK-related MODY (MODY2; 125851) and homozygosity caused neonatal-onset diabetes.


.0011   MOVED TO 138079.0003


.0012   HYPERINSULINEMIC HYPOGLYCEMIA, FAMILIAL, 3

GCK, ALA456VAL
SNP: rs104894014, ClinVar: RCV000017525, RCV001818166

In a 14-year-old obese boy with a history of neonatal hypoglycemia treated with diazoxide, who was experiencing hypoglycemic episodes associated with seizures and unconsciousness (see HHF3, 602485), Christesen et al. (2002) identified heterozygosity for an ala456-to-val (A456V) substitution in exon 10 of the GCK gene. Kinetic analysis showed this to be an activating mutation. The boy's normal-weight mother, who had asymptomatic fasting hypoglycemia, carried the same mutation; the mutation was not found in his normoglycemic brother nor in 80 controls.


.0013   HYPERINSULINEMIC HYPOGLYCEMIA, FAMILIAL, 3

GCK, TYR214CYS
SNP: rs104894015, ClinVar: RCV000017526

In a Finnish woman with severe hyperinsulinemic hypoglycemia from birth (602485), who had severe mental retardation and was still having hypoglycemic seizures when she died at age 29, Cuesta-Munoz et al. (2004) identified heterozygosity for a de novo tyr214-to-cys (Y214C) substitution in exon 6 of the GCK gene. Although paternity was confirmed, the mutation was not found in her parents or her 2 healthy sisters. Kinetic analysis revealed that this mutation had the highest activity index (130-fold over wildtype) of all naturally occurring activating GCK mutations described. Cuesta-Munoz et al. (2004) noted that this phenotype was considerably more severe than that of previously reported patients (see 138079.0009 and 138079.0012).


.0014   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, ALA378THR
SNP: rs104894016, gnomAD: rs104894016, ClinVar: RCV000017527, RCV002513080, RCV003325941

In affected members from 5 Belgian families with MODY2 (125851), Vits et al. (2006) identified a 1132G-A transition in exon 9 of the GCK gene, resulting in an ala378-to-thr (A378T) substitution. All the probands originated from the Belgian province of West Flanders, suggesting a founder mutation; this was confirmed in 3 families by haplotype analysis.


.0015   HYPERINSULINEMIC HYPOGLYCEMIA, FAMILIAL, 3

GCK, VAL91LEU
ClinVar: RCV000017528

Kassem et al. (2010) studied a young girl with severe neonatal hypoglycemia-3 (602485) due to a val91-to-leu (V91L) missense mutation in the GCK gene. Her father had a similar clinical course but neither his DNA nor pancreatic tissue was available for study. Quantitative histologic examination after subtotal pancreatectomy due to refractory disease revealed abnormally large islets, with some beta cells containing a large nucleus, and mean islet cell areas in both the head and the tail of the pancreas were significantly larger than those of 5 age-matched controls and those of 2 age-matched HHF1 (256450) patients.


.0016   MATURITY-ONSET DIABETES OF THE YOUNG, TYPE 2

GCK, GLU339LYS
SNP: rs397514580, ClinVar: RCV000032978, RCV003883126

In 5 affected members over 3 generations of a mainland Chinese family with type 2 maturity-onset diabetes of the young (MODY2; 125851), Shen et al. (2011) identified heterozygosity for a G-A transition in exon 8 of the GCK gene, resulting in a glu339-to-lys (E339K) substitution. The mutation was not found in unaffected family members or in 200 controls. SDS-PAGE analysis of a bacterial expression system demonstrated that the protein yield of mutant GCK was significantly lower than wildtype; kinetic analysis showed that the E339K mutant had 1.4-fold and 9.9-fold lower affinity for glucose and ATP, respectively, compared to wildtype. The mutant GCK also exhibited thermal instability, with a dramatic decrease in activity at 45 degrees centigrade compared to 55 degrees centigrade for wildtype; in addition, wildtype GCK maintained 50% activity at 55 degrees centigrade for 30 minutes, whereas wildtype lost 97% of activity within 30 minutes.


REFERENCES

  1. Aizawa, T., Asanuma, N., Terauchi, Y., Suzuki, N., Komatsu, M., Itoh, N., Nakabayashi, T., Hidaka, H., Ohnota, H., Yamauchi, K., Yasuda, K., Yazaki, Y., Kadowaki, T., Hashizume, K. Analysis of the pancreatic beta cell in the mouse with targeted disruption of the pancreatic beta cell-specific glucokinase gene. Biochem. Biophys. Res. Commun. 229: 460-465, 1996. [PubMed: 8954920] [Full Text: https://doi.org/10.1006/bbrc.1996.1826]

  2. Bali, D., Svetlanov, A., Lee, H.-W., Fusco-DeMane, D., Leiser, M., Li, B., Barzilai, N., Surana, M., Hou, H., Fleischer, N., DePinho, R., Rossetti, L., Efrat, S. Animal model for maturity-onset diabetes of the young generated by disruption of the mouse glucokinase gene. J. Biol. Chem. 270: 21464-21467, 1995. [PubMed: 7665557] [Full Text: https://doi.org/10.1074/jbc.270.37.21464]

  3. Byrne, M. M., Sturis, J., Clement, K., Vionnet, N., Pueyo, M. E., Stoffel, M., Takeda, J., Passa, P., Cohen, D., Bell, G. I., Velho, G., Froguel, P., Polonsky, K. S. Insulin secretory abnormalities in subjects with hyperglycemia due to glucokinase mutations. J. Clin. Invest. 93: 1120-1130, 1994. [PubMed: 8132752] [Full Text: https://doi.org/10.1172/JCI117064]

  4. Chiu, K. C., Tanizawa, Y., Permutt, M. A. Non-sense mutation of glucokinase gene. (Letter) Lancet 341: 385-386, 1993. [PubMed: 8094164]

  5. Christesen, H. B. T., Jacobsen, B. B., Odili, S., Buettger, C., Cuesta-Munoz, A., Hansen, T., Brusgaard, K., Massa, O., Magnuson, M. A., Shiota, C., Matschinsky, F., Barbetti, F. The second activating glucokinase mutation (A456V): implications for glucose homeostasis and diabetes therapy. Diabetes 51: 1240-1246, 2002. [PubMed: 11916951] [Full Text: https://doi.org/10.2337/diabetes.51.4.1240]

  6. Concannon, P. Personal Communication. Seattle, Wash. 7/26/1995.

  7. Cuesta-Munoz, A. L., Huopio, H., Otonkoski, T., Gomez-Zumaquero, J. M., Nanto-Salonen, K., Rahier, J., Lopez-Enriquez, S., Garcia-Gimeno, M. A., Sanz, P., Soriguer, F. C., Laakso, M. Severe persistent hyperinsulinemic hypoglycemia due to a de novo glucokinase mutation. Diabetes 53: 2164-2168, 2004. [PubMed: 15277402] [Full Text: https://doi.org/10.2337/diabetes.53.8.2164]

  8. Danial, N. N., Gramm, C. F., Scorrano, L., Zhang, C.-Y., Krauss, S., Ranger, A. M., Datta, S. R., Greenberg, M. E., Licklider, L. J., Lowell, B. B., Gygi, S. P., Korsmeyer, S. J. BAD and glucokinase reside in a mitochondrial complex that integrates glycolysis and apoptosis. Nature 424: 952-956, 2003. [PubMed: 12931191] [Full Text: https://doi.org/10.1038/nature01825]

  9. Datta, S. R., Ranger, A. M., Lin, M. Z., Sturgill, J. F., Ma, Y.-C., Cowan, C. W., Dikkes, P., Korsmeyer, S. J., Greenberg, M. E. Survival factor-mediated BAD phosphorylation raises the mitochondrial threshold for apoptosis. Dev. Cell 3: 631-643, 2002. [PubMed: 12431371] [Full Text: https://doi.org/10.1016/s1534-5807(02)00326-x]

  10. Dunger, D. B., Ong, K. K. L., Huxtable, S. J., Sherriff, A., Woods, K. A., Ahmed, M. L., Golding, J., Pembrey, M. E., Ring, S., ALSPAC Study Team, Bennett, S. T., Todd, J. A. Association of the INS VNTR with size at birth. Nature Genet. 19: 98-100, 1998. [PubMed: 9590300] [Full Text: https://doi.org/10.1038/ng0598-98]

  11. Froguel, P., Vaxillaire, M., Sun, F., Velho, G., Zouali, H., Butel, M. O., Lesage, S., Vionnet, N., Clement, K., Fougerousse, F., Tanizawa, Y., Weissenbach, J., Beckmann, J. S., Lathrop, G. M., Passa, P., Permutt, M. A., Cohen, D. Close linkage of glucokinase locus on chromosome 7p to early-onset non-insulin-dependent diabetes mellitus. Nature 356: 162-164, 1992. Note: Erratum: Nature 357: 607 only, 1992. [PubMed: 1545870] [Full Text: https://doi.org/10.1038/356162a0]

  12. Froguel, P., Velho, G. Non-sense mutation of glucokinase gene. (Letter) Lancet 341: 385 only, 1993. [PubMed: 8094163]

  13. Froguel, P., Zouali, H., Vionnet, N., Velho, G., Vaxillaire, M., Sun, F., Lesage, S., Stoffel, M., Takeda, J., Passa, P., Permutt, M. A., Beckmann, J. S., Bell, G. I., Cohen, D. Familial hyperglycemia due to mutations in glucokinase: definition of a subtype of diabetes mellitus. New Eng. J. Med. 328: 697-702, 1993. [PubMed: 8433729] [Full Text: https://doi.org/10.1056/NEJM199303113281005]

  14. Glaser, B., Kesavan, P., Heyman, M., Davis, E., Cuesta, A., Buchs, A., Stanley, C. A., Thornton, P. S., Permutt, M. A., Matschinsky, F. M., Herold, K. C. Familial hyperinsulinism caused by an activating glucokinase mutation. New Eng. J. Med. 338: 226-230, 1998. [PubMed: 9435328] [Full Text: https://doi.org/10.1056/NEJM199801223380404]

  15. Gloyn, A. L., Noordam, K., Willemsen, M. A. A. P., Ellard, S., Lam, W. W. K., Campbell, I. W., Midgley, P., Shiota, C., Buettger, C., Magnuson, M. A., Matschinsky, F. M., Hattersley, A. T. Insights into the biochemical and genetic basis of glucokinase activation from naturally occurring hypoglycemia mutations. Diabetes 52: 2433-2440, 2003. [PubMed: 12941786] [Full Text: https://doi.org/10.2337/diabetes.52.9.2433]

  16. Gloyn, A. L. Glucokinase (GCK) mutations in hyper- and hypoglycemia: maturity-onset diabetes of the young, permanent neonatal diabetes, and hyperinsulinemia of infancy. Hum. Mutat. 22: 353-362, 2003. [PubMed: 14517946] [Full Text: https://doi.org/10.1002/humu.10277]

  17. Grimsby, J., Sarabu, R., Corbett, W. L., Haynes, N.-E., Bizzarro, F. T., Coffey, J. W., Guertin, K. R., Hilliard, D. W., Kester, R. F., Mahaney, P. E., Marcus, L., Qi, L., Spence, C. L., Tengi, J., Magnuson, M. A., Chu, C. A., Dvorozniak, M. T., Matschinsky, F. M., Grippo, J. F. Allosteric activators of glucokinase: potential role in diabetes therapy. Science 301: 370-373, 2003. [PubMed: 12869762] [Full Text: https://doi.org/10.1126/science.1084073]

  18. Grupe, A., Hultgren, B., Ryan, A., Ma, Y. H., Bauer, M., Stewart, T. A. Transgenic knockouts reveal a critical requirement for pancreatic beta cell glucokinase in maintaining glucose homeostasis. Cell 83: 69-78, 1995. [PubMed: 7553875] [Full Text: https://doi.org/10.1016/0092-8674(95)90235-x]

  19. Hattersley, A. T., Beards, F., Ballantyne, E., Appleton, M., Harvey, R., Ellard, S. Mutations in the glucokinase gene of the fetus result in reduced birth weight. Nature Genet. 19: 268-270, 1998. [PubMed: 9662401] [Full Text: https://doi.org/10.1038/953]

  20. Hattersley, A. T., Turner, R. C., Permutt, M. A., Patel, P., Tanizawa, Y., Chiu, K. C., O'Rahilly, S., Watkins, P. J., Wainscoat, J. S. Linkage of type 2 diabetes to the glucokinase gene. Lancet 339: 1307-1310, 1992. [PubMed: 1349989] [Full Text: https://doi.org/10.1016/0140-6736(92)91958-b]

  21. Heimberg, H., De Vos, A., Moens, K., Quartier, E., Bouwens, L., Pipeleers, D., Van Schaftingen, E., Madsen, O., Schuit, F. The glucose sensor protein glucokinase is expressed in glucagon-producing alpha-cells. Proc. Nat. Acad. Sci. 93: 7036-7041, 1996. [PubMed: 8692940] [Full Text: https://doi.org/10.1073/pnas.93.14.7036]

  22. Hofmeister-Brix, A., Kollmann, K., Langer, S., Schultz, J., Lenzen, S., Baltrusch, S. Identification of the ubiquitin-like domain of midnolin as a new glucokinase interaction partner. J. Biol. Chem. 288: 35824-35839, 2013. [PubMed: 24187134] [Full Text: https://doi.org/10.1074/jbc.M113.526632]

  23. Inoue, M., Sakuraba, Y., Motegi, H., Kubota, N., Toki, H., Matsui, J., Toyoda, Y., Miwa, I., Terauchi, Y., Kadowaki, T., Shigeyama, Y., Kasuga, M. A series of maturity onset diabetes of the young, type 2 (MODY2) mouse models generated by a large-scale ENU mutagenesis program. Hum. Molec. Genet. 13: 1147-1157, 2004. [PubMed: 15102714] [Full Text: https://doi.org/10.1093/hmg/ddh133]

  24. Johansen, A., Ek, J., Mortensen, H. B., Pedersen, O., Hansen, T. Half of clinically defined maturity-onset diabetes of the young patients in Denmark do not have mutations in HNF4A, GCK, and TCF1. J. Clin. Endocr. Metab. 90: 4607-4614, 2005. [PubMed: 15928245] [Full Text: https://doi.org/10.1210/jc.2005-0196]

  25. Kassem, S., Bhandari, S., Rodriguez-Bada, P., Motaghedi, R., Heyman, M., Garcia-Gimeno, M. A., Cobo-Vuilleumier, N., Sanz, P., Maclaren, N. K., Rahier, J., Glaser, B., Cuesta-Munoz, A. L. Large islets, beta-cell proliferation, and a glucokinase mutation. New Eng. J. Med. 362: 1348-1350, 2010. Note: Erratum: New Eng. J. Med. 363: 2178 only, 2010. [PubMed: 20375417] [Full Text: https://doi.org/10.1056/NEJMc0909845]

  26. Katagiri, H., Asano, T., Ishihara, H., Inukai, K., Anai, M., Miyazaki, J., Tsukuda, K., Kikuchi, M., Yazaki, Y., Oka, Y. Nonsense mutation of glucokinase gene in late-onset non-insulin-dependent diabetes mellitus. Lancet 340: 1316-1317, 1992. [PubMed: 1360036] [Full Text: https://doi.org/10.1016/0140-6736(92)92494-z]

  27. Kim, J. Y., Hwang, J.-Y., Lee, D. Y., Song, E. H., Park, K. J., Kim, G. H., Jeong, E. A., Lee, Y. J., Go, M. J., Kim, D. J., Lee, S. S., Kim, B.-J., Song, J., Roh, G. S., Gao, B., Kim, W.-H. Chronic ethanol consumption inhibits glucokinase transcriptional activity by Atf3 and triggers metabolic syndrome in vivo. J. Biol. Chem. 289: 27065-27079, 2014. [PubMed: 25074928] [Full Text: https://doi.org/10.1074/jbc.M114.585653]

  28. Marz, W., Nauck, M., Hoffmann, M. M., Nagel, D., Boehm, B. O., Koenig, W., Rothenbacher, D., Winkelmann, B. R. G(-30)A polymorphism in the pancreatic promoter of the glucokinase gene associated with angiographic coronary artery disease and type 2 diabetes mellitus. Circulation 109: 2844-2849, 2004. [PubMed: 15173029] [Full Text: https://doi.org/10.1161/01.CIR.0000129306.44085.C4]

  29. Matschinsky, F., Liang, Y., Kesavan, P., Wang, L., Froguel, P., Velho, G., Cohen, D., Permutt, M. A., Tanizawa, Y., Jetton, T. L., Niswender, K., Magnuson, M. A. Glucokinase as pancreatic beta-cell glucose sensor and diabetes gene. J. Clin. Invest. 92: 2092-2098, 1993. [PubMed: 8227324] [Full Text: https://doi.org/10.1172/JCI116809]

  30. Matschinsky, F. M. Glucokinase as glucose sensor and metabolic signal generator in pancreatic beta-cells and hepatocytes. Diabetes 39: 647-652, 1990. [PubMed: 2189759] [Full Text: https://doi.org/10.2337/diab.39.6.647]

  31. Matsutani, A., Janssen, R., Donis-Keller, H., Permutt, M. A. A polymorphic (CA)n repeat element maps the human glucokinase gene (GCK) to chromosome 7p. Genomics 12: 319-325, 1992. [PubMed: 1740341] [Full Text: https://doi.org/10.1016/0888-7543(92)90380-b]

  32. McCarthy, M. Weighing in on diabetes risk. Nature Genet. 19: 209-210, 1998. [PubMed: 9662384] [Full Text: https://doi.org/10.1038/876]

  33. Mirshahi, U. L., Colclough, K., Wright, C. F., Wood, A. R., Beaumont, R. N., Tyrrell, J., Laver, T. W., Stahl, R., Golden, A., Goehringer, J. M, Geisinger-Regeneron DiscovEHR Collaboration, Frayling, T. F., Hattersley, A. T., Carey, D. J., Weedon, M. N., Patel, K. A. Reduced penetrance of MODY-associated HNF1A/HNF4A variants but not GCK variants in clinically unselected cohorts. Am. J. Hum. Genet. 109: 2018-2028, 2022. [PubMed: 36257325] [Full Text: https://doi.org/10.1016/j.ajhg.2022.09.014]

  34. Mishra, S. K., Helms, C., Dorsey, D., Permutt, M. A., Donis-Keller, H. A 2-cM genetic linkage map of human chromosome 7p that includes 47 loci. Genomics 12: 326-334, 1992. [PubMed: 1740342] [Full Text: https://doi.org/10.1016/0888-7543(92)90381-2]

  35. Neel, J. V., Weder, A. B., Julius, S. Type II diabetes, essential hypertension, and obesity as 'syndromes of impaired genetic homeostasis': the 'thrifty genotype' hypothesis enters the 21st century. Perspect. Biol. Med. 42: 44-74, 1998. [PubMed: 9894356] [Full Text: https://doi.org/10.1353/pbm.1998.0060]

  36. Neel, J. V. Diabetes mellitus: a 'thrifty' genotype rendered detrimental by 'progress'? Am. J. Hum. Genet. 14: 353-362, 1962. [PubMed: 13937884]

  37. Njolstad, P. R., Sovik, O., Cuesta-Munoz, A., Bjorkhaug, L., Massa, O., Barbetti, F., Undlien, D. E., Shiota, C., Magnuson, M. A., Molven, A., Matschinsky, F. M., Bell, G. I. Neonatal diabetes mellitus due to complete glucokinase deficiency. New Eng. J. Med. 344: 1588-1592, 2001. [PubMed: 11372010] [Full Text: https://doi.org/10.1056/NEJM200105243442104]

  38. Permutt, M. A., Chiu, K. C., Tanizawa, Y. Glucokinase and NIDDM: a candidate gene that paid off. Diabetes 41: 1367-1372, 1992. [PubMed: 1397713] [Full Text: https://doi.org/10.2337/diab.41.11.1367]

  39. Pinterova, D., Ek, J., Kolostova, K., Pruhova, S., Novota, P., Romzova, M., Feigerlova, E., Cerna, M., Lebl, J., Pedersen, O., Hansen, T. Six novel mutations in the GCK gene in MODY patients. (Letter) Clin. Genet. 71: 95-96, 2007. [PubMed: 17204055] [Full Text: https://doi.org/10.1111/j.1399-0004.2006.00729.x]

  40. Randle, P. J. Glucokinase and candidate genes for type 2 (non-insulin-dependent) diabetes mellitus. Diabetologia 36: 269-275, 1993. [PubMed: 8386682] [Full Text: https://doi.org/10.1007/BF00400227]

  41. Rowe, R. E., Wapelhorst, B., Bell, G. I., Risch, N., Spielman, R. S., Concannon, P. Linkage and association between insulin-dependent diabetes mellitus (IDDM) susceptibility and markers near the glucokinase gene on chromosome 7. Nature Genet. 10: 240-245, 1995. [PubMed: 7663523] [Full Text: https://doi.org/10.1038/ng0695-240]

  42. Shen, Y., Cai, M., Liang, H., Wang, H., Weng, J. Insight into the biochemical characteristics of a novel glucokinase gene mutation. Hum. Genet. 129: 231-238, 2011. [PubMed: 21104275] [Full Text: https://doi.org/10.1007/s00439-010-0914-4]

  43. Stoffel, M., Bell, K. L., Blackburn, C. L., Powell, K. L., Seo, T. S., Takeda, J., Vionnet, N., Xiang, K.-S., Gidh-Jain, M., Pilkis, S. J., Ober, C., Bell, G. I. Identification of glucokinase mutations in subjects with gestational diabetes mellitus. Diabetes 42: 937-940, 1993. [PubMed: 8495817] [Full Text: https://doi.org/10.2337/diab.42.6.937]

  44. Stoffel, M., Froguel, P., Takeda, J., Zouali, H., Vionnet, N., Nishi, S., Weber, I. T., Harrison, R. W., Pilkis, S. J., Lesage, S., Vaxillaire, M., Velho, G., Sun, F., Iris, F., Passa, P., Cohen, D., Bell, G. I. Human glucokinase gene: isolation, characterization, and identification of two missense mutations linked to early-onset non-insulin-dependent (type 2) diabetes mellitus. Proc. Nat. Acad. Sci. 89: 7698-7702, 1992. Note: Erratum: Proc. Nat. Acad Sci. 89: 10562 only, 1992. [PubMed: 1502186] [Full Text: https://doi.org/10.1073/pnas.89.16.7698]

  45. Stoffel, M., Patel, P., Lo, Y.-M. D., Hattersley, A. T., Lucassen, A. M., Page, R., Bell, J. I., Bell, G. I., Turner, R. C., Wainscoat, J. S. Missense glucokinase mutation in maturity-onset diabetes of the young and mutation screening in late-onset diabetes. Nature Genet. 2: 153-156, 1992. [PubMed: 1303265] [Full Text: https://doi.org/10.1038/ng1092-153]

  46. Stone, L. M., Kahn, S. E., Fujimoto, W. Y., Deeb, S. S., Porte, D. Jr. A variation at position -30 of the beta-cell glucokinase gene promoter is associated with reduced beta-cell function in middle-aged Japanese-American men. Diabetes 45: 422-428, 1996. [PubMed: 8603762] [Full Text: https://doi.org/10.2337/diab.45.4.422]

  47. Sun, F., Knebelmann, B., Pueyo, M. E., Zouali, H., Lesage, S., Vaxillaire, M., Passa, P., Cohen, D., Velho, G., Antignac, C., Froguel, P. Deletion of the donor splice site of intron 4 in the glucokinase gene causes maturity-onset diabetes of the young. J. Clin. Invest. 92: 1174-1180, 1993. [PubMed: 8376578] [Full Text: https://doi.org/10.1172/JCI116687]

  48. Tanizawa, Y., Koranyi, L. I., Welling, C. M., Permutt, M. A. Human liver glucokinase gene: cloning and sequence determination of two alternatively spliced cDNAs. Proc. Nat. Acad. Sci. 88: 7294-7297, 1991. [PubMed: 1871135] [Full Text: https://doi.org/10.1073/pnas.88.16.7294]

  49. Terauchi, Y., Takamoto, I., Kubota, N., Matsui, J., Suzuki, R., Komeda, K., Hara, A., Toyoda, Y., Miwa, I., Aizawa, S., Tsutsumi, S., Tsubamoto, Y., Hashimoto, S., Eto, K., Nakamura, A., Noda, M., Tobe, K., Aburatani, H., Nagai, R., Kadowaki, T. Glucokinase and IRS-2 are required for compensatory beta cell hyperplasia in response to high-fat diet-induced insulin resistance. J. Clin. Invest. 117: 246-257, 2007. [PubMed: 17200721] [Full Text: https://doi.org/10.1172/JCI17645]

  50. Thornton, P. S., Satin-Smith, M. S., Herold, K., Glaser, B., Chiu, K. C., Nestorowicz, A., Permutt, M. A., Baker, L., Stanley, C. A. Familial hyperinsulinism with apparent autosomal dominant inheritance: clinical and genetic differences from the autosomal recessive variant. J. Pediat. 132: 9-14, 1998. [PubMed: 9469993] [Full Text: https://doi.org/10.1016/s0022-3476(98)70477-9]

  51. Velho, G., Blanche, H., Vaxillaire, M., Bellane-Chantelot, C., Pardini, V. C., Timsit, J., Passa, P., Deschamps, I., Robert, J.-J., Weber, I. T., Marotta, D., Pilkis, S. J., Lipkind, G. M., Bell, G. I., Froguel, P. Identification of 14 new glucokinase mutations and description of the clinical profile of 42 MODY-2 families. Diabetologia 40: 217-224, 1997. [PubMed: 9049484] [Full Text: https://doi.org/10.1007/s001250050666]

  52. Velho, G., Froguel, P., Clement, K., Pueyo, M. E., Rakotoambinina, B., Zouali, H., Passa, P., Cohen, D., Robert, J.-J. Primary pancreatic beta-cell secretory defect caused by mutations in glucokinase gene in kindreds of maturity onset diabetes of the young. Lancet 340: 444-448, 1992. [PubMed: 1354782] [Full Text: https://doi.org/10.1016/0140-6736(92)91768-4]

  53. Velho, G., Petersen, K. F., Perseghin, G., Hwang, J.-H., Rothman, D. L., Pueyo, M. E., Cline, G. W., Froguel, P., Shulman, G. I. Impaired hepatic glycogen synthesis in glucokinase-deficient (MODY-2) subjects. J. Clin. Invest. 98: 1755-1761, 1996. [PubMed: 8878425] [Full Text: https://doi.org/10.1172/JCI118974]

  54. Vionnet, N., Stoffel, M., Takeda, J., Yasuda, K., Bell, G. I., Zouali, H., Lesage, S., Velho, G., Iris, F., Passa, P., Froguel, P., Cohen, D. Nonsense mutation in the glucokinase gene causes early-onset non-insulin-dependent diabetes mellitus. Nature 356: 721-722, 1992. [PubMed: 1570017] [Full Text: https://doi.org/10.1038/356721a0]

  55. Vits, L., Beckers, D., Craen, M., de Beaufort, C., Vanfleteren, E., Dahan, K., Nollet, A., Vanhaverbeke, G., Imschoot, S. V., Bourguignon, J.-P., Beauloye, V., Storm, K., Massa, G., Giri, M., Nobels, F., De Schepper, J., Rooman, R., Van den Bruel, A., Mathieu, C., Wuyts, W. Identification of novel and recurrent glucokinase mutations in Belgian and Luxembourg maturity onset diabetes of the young patients. (Letter) Clin. Genet. 70: 355-359, 2006. [PubMed: 16965331] [Full Text: https://doi.org/10.1111/j.1399-0004.2006.00686.x]


Contributors:
Ada Hamosh - updated : 01/17/2023
Patricia A. Hartz - updated : 02/14/2018
Patricia A. Hartz - updated : 6/24/2014
Marla J. F. O'Neill - updated : 11/2/2012
Patricia A. Hartz - updated : 4/15/2010
Marla J. F. O'Neill - reorganized : 4/14/2010
Marla J. F. O'Neill - updated : 4/14/2010
Ada Hamosh - updated : 1/15/2010
Victor A. McKusick - updated : 5/23/2007
Marla J. F. O'Neill - updated : 4/17/2007
Marla J. F. O'Neill - updated : 3/9/2007
Victor A. McKusick - updated : 11/28/2006
Cassandra L. Kniffin - updated : 11/2/2006
John A. Phillips, III - updated : 10/19/2006
George E. Tiller - updated : 8/31/2006
Marla J. F. O'Neill - updated : 3/17/2006
Marla J. F. O'Neill - updated : 1/25/2006
Victor A. McKusick - updated : 11/19/2003
Ada Hamosh - updated : 9/16/2003
Ada Hamosh - updated : 8/5/2003
Ada Hamosh - updated : 10/18/2001
Ada Hamosh - updated : 8/23/2001
Victor A. McKusick - updated : 6/25/2001
Victor A. McKusick - updated : 2/2/1999
Victor A. McKusick - updated : 6/25/1998
Victor A. McKusick - updated : 4/15/1998
Jennifer P. Macke - updated : 9/2/1997

Creation Date:
Victor A. McKusick : 2/1/1992

Edit History:
alopez : 01/17/2023
alopez : 12/01/2020
carol : 09/02/2020
carol : 07/14/2020
alopez : 04/30/2020
mgross : 02/14/2018
alopez : 08/04/2016
carol : 08/12/2015
carol : 8/12/2015
mgross : 6/24/2014
mgross : 10/7/2013
mgross : 10/7/2013
terry : 4/4/2013
carol : 11/6/2012
terry : 11/2/2012
terry : 7/5/2012
carol : 12/15/2010
alopez : 4/22/2010
mgross : 4/16/2010
terry : 4/15/2010
carol : 4/14/2010
carol : 4/14/2010
wwang : 2/3/2010
alopez : 1/27/2010
alopez : 1/26/2010
terry : 1/15/2010
terry : 9/11/2009
joanna : 9/4/2009
alopez : 5/29/2007
terry : 5/23/2007
wwang : 4/17/2007
wwang : 3/13/2007
terry : 3/9/2007
alopez : 12/8/2006
terry : 11/28/2006
carol : 11/3/2006
ckniffin : 11/2/2006
alopez : 10/19/2006
alopez : 8/31/2006
carol : 3/17/2006
carol : 3/16/2006
carol : 3/16/2006
carol : 3/16/2006
carol : 3/16/2006
wwang : 2/2/2006
terry : 1/25/2006
carol : 12/3/2004
carol : 12/3/2004
tkritzer : 12/30/2003
tkritzer : 12/18/2003
tkritzer : 11/24/2003
terry : 11/19/2003
alopez : 9/16/2003
alopez : 8/7/2003
terry : 8/5/2003
terry : 6/27/2002
carol : 10/18/2001
carol : 8/23/2001
mcapotos : 7/6/2001
mcapotos : 6/29/2001
terry : 6/25/2001
carol : 12/26/2000
mgross : 3/16/1999
carol : 2/15/1999
carol : 2/15/1999
terry : 2/2/1999
dkim : 12/10/1998
alopez : 6/29/1998
terry : 6/25/1998
carol : 4/20/1998
terry : 4/15/1998
mark : 10/20/1997
mark : 10/17/1997
alopez : 10/7/1997
alopez : 10/7/1997
alopez : 10/6/1997
terry : 11/25/1996
terry : 11/12/1996
terry : 11/1/1996
terry : 9/17/1996
marlene : 8/15/1996
terry : 11/2/1995
mark : 8/18/1995
mimadm : 9/24/1994
jason : 7/14/1994
carol : 5/10/1994
carol : 12/7/1993