ArticlePDF AvailableLiterature Review

Molecular Physiology of Mammalian Glucokinase

Authors:

Abstract

The glucokinase (GCK) gene was one of the first candidate genes to be identified as a human "diabetes gene". Subsequently, important advances were made in understanding the impact of GCK in the regulation of glucose metabolism. Structure elucidation by crystallography provided insight into the kinetic properties of GCK. Protein interaction partners of GCK were discovered. Gene expression studies revealed new facets of the tissue distribution of GCK, including in the brain, and its regulation by insulin in the liver. Metabolic control analysis coupled to gene overexpression and knockout experiments highlighted the unique impact of GCK as a regulator of glucose metabolism. Human GCK mutants were studied biochemically to understand disease mechanisms. Drug development programs identified small molecule activators of GCK as potential antidiabetics. These advances are summarized here, with the aim of offering an integrated view of the role of GCK in the molecular physiology and medicine of glucose homeostasis.
Review
Molecular Physiology of Mammalian Glucokinase
P. B. Iynedjian
Department of Cell Physiolgy and Metabolism, University of Geneva School of Medicine,
CMU 1 Rue Michel-Servet, 1211 Geneva 4 (Switzerland), Fax: +41 22/379 55 43,
e-mail: patrick.iynedjian@medecine.unige.ch
Received 11 June 2008; received after revision 18 July 2008; accepted 30 July 2008
Online First 26 August 2008
Abstract. The glucokinase (GCK) gene was one of the
first candidate genes to be identified as a human
“diabetes gene”. Subsequently, important advances
were made in understanding the impact of GCK in the
regulation of glucose metabolism. Structure elucida-
tion by crystallography provided insight into the
kinetic properties of GCK. Protein interaction part-
ners of GCK were discovered. Gene expression
studies revealed new facets of the tissue distribution
of GCK, including in the brain, and its regulation by
insulin in the liver. Metabolic control analysis coupled
to gene overexpression and knockout experiments
highlighted the unique impact of GCK as a regulator
of glucose metabolism. Human GCK mutants were
studied biochemically to understand disease mecha-
nisms. Drug development programs identified small
molecule activators of GCK as potential antidiabetics.
These advances are summarized here, with the aim of
offering an integrated view of the role of GCK in the
molecular physiology and medicine of glucose homeo-
stasis.
Keywords. Glucokinase, diabetes, glucose metabolism, insulin, hexokinase, islet of Langerhans, hepatocyte,
MODY.
Introduction
This review focuses on the enzyme known as mamma-
lian glucokinase. Somewhat ironically, this enzyme is
actually not a glucokinase in the strict sense, denoting
a glucose phosphorylating enzyme with stringent
substrate specificity for glucose, such as exists in
lower organisms. Mammalian glucokinase is able to
phosphorylate hexoses like mannose or fructose in
addition to glucose, a property shared with three other
hexokinases present at various levels and in various
combinations in human, rat, and mouse tissues [1].
The four mammalian isoenzymes display extensive
sequence identities, leaving no doubt about a close
phylogenetic relationship [2], and arguing for the
terminology hexokinases I IV (or A D), in which
hexokinase IV (or D) designates mammalian gluco-
kinase. In spite of these considerations, the term
glucokinase has become entrenched in tradition,
probably because it conveys a sense of the very unique
role of this enzyme in the regulation of glucose
homeostasis. This tradition is followed in the current
article, with the understanding that glucokinase
(GCK) stands for hexokinase IV (or D). Whether
the name should be reconsidered in the future is an
open question, because a gene encoding a functional
ADP-dependent glucokinase reminiscent of primitive
organisms was recently identified in the mouse and
human genomes [3]. The level of activity and physio-
logical role of this enzyme in mammalian tissues are
currently unknown.
Glucokinase was identified as an enzyme of rat liver in
1962. In 1992, two studies using the candidate gene
approach reported a genetic linkage between a form
Cell. Mol. Life Sci. 66 (2009) 27 42
1420-682X/09/010027-16
DOI 10.1007/s00018-008-8322-9
Birkhuser Verlag, Basel, 2008
Cellular and Molecular Life Sciences
of maturity onset diabetes of the young, named
MODY 2, and the GCK gene [4, 5] . This discovery
came as the culmination of three decades of research
in biochemistry, physiology and molecular biology,
which collectively demonstrated the preferential ex-
pression of GCK in hepatocytes and b-cells of the
pancreas, and suggested a key role in glucose metab-
olism. These investigations, which had conferred to
GCK the rank of prime candidate as a “diabetes
gene”, were the subject of two reviews published in
1993 [6, 7].
In recent years our perception of the pivotal role of
GCK in the regulation of glucose metabolism has
been reinforced by important results in a variety of
experimental areas. Structural investigations by X-
ray crystallography shed new light on the very
special enzyme kinetics of GCK. Gene expression
studies revealed novel facets of the tissue distribu-
tion of GCK, notably in brain, and of the insulin
regulation of expression in the liver. Genetic
manipulations in cells and mice led to major
advances in our understanding of the glucose sensor
function of GCK at the cellular level and its impact
on whole body glucose homeostasis. This article was
written to assemble knowledge from very diverse
lines of investigation, with the aim of offering an
integrated view of the unique place of GCK in
physiology and medicine.
Three-dimensional structure of GCK: cooperative
glucose kinetics explained
Glucose taken up by mammalian cells has to be
converted into glucose 6-phosphate as a prelude to
further utilization for glycolysis, the pentose phos-
phate pathway or glycogen synthesis. Although some
contribution to overall glucose phosphorylation by the
recently identified ADP-dependent glucokinase [3]
cannot be ruled out, the bulk of glucose phosphor-
ylation occurs at the expense of ATP in the generic
reaction catalyzed by the hexokinases: hexose +ATP
!hexose 6-phosphate +ADP [8].
Among the hexokinases, GCK displays unique en-
zyme kinetics in two respects. First, GCK has an
affinity for glucose (S0.5 6 mM) more than 20 times
lower than that of the next ranking hexokinase,
hexokinase II. Second, the rate of the GCK reaction
displays a sigmoid rather than hyperbolic dependence
on glucose concentration [1]. These two kinetic
properties, low affinity and cooperativity with respect
to glucose, allow GCK to operate as an ultrasensitive
physiological glucose sensor in cells endowed with
non-limiting glucose transport activity at the plasma
membrane (see below).
The cooperativity of GCK with glucose cannot be
explained by the classical mechanisms of allostery for
multisubunit proteins, because GCK is catalytically
active in the monomeric state [9]. Moreover, GCK
harbours a single glucose binding site (the active site)
[10], excluding the possibility of multisite allostery in
which two molecules of glucose would bind simulta-
neously to the enzyme. Deviation from hyperbolic
kinetics is sometimes noted for monomeric enzymes
in case of random addition of substrates, but this is not
the case for the GCK catalytic cycle, which involves an
ordered addition of substrates in which glucose binds
first and ATP second [11]. A mechanism of positive
cooperativity applicable to monomeric enzymes,
named the mnemonic model, was proposed for GCK
on the basis of early enzymological investigations [12] .
Predictions from the mnemonic model were con-
firmed by recent studies using pre-steady-state meas-
urements of the intrinsic fluorescence of GCK during
binding of glucose and other ligands [13, 14]. The
mnemonic mechanism of cooperativity includes the
following features: i) equilibrium between two con-
formational states of GCK with vastly different
glucose affinities; ii) large predominance of the low
affinity conformation in the absence of glucose; iii)
slow interconversion between the conformational
states for glucose-free GCK; iv) conversion from the
low affinity to the high affinity conformational state
strongly accelerated upon glucose binding to GCK ; v)
only the high affinity form of the enzyme is catalyti-
cally competent, and the rate of the catalytic step(s) is
very fast compared to the conformational transition
rate.
Explicit support for the model came recently from
structural data on human GCK obtained by Kamata
and colleagues, who solved the structure of GCK by
X-ray crystallography [15]. Crystals were obtained
from GCK with short truncations at the NH2-terminal
end (11 or 15 amino acids) under two distinct
crystallization conditions: in the presence of glucose
plus a small-molecule activator of GCK (see below),
or in the absence of any ligand. The overall structure
of the protein was comprised of a large and a small
globular domain connected by a hinge made up of
three flexible loops. With glucose and the activator
present, the space between the two domains assumed
the shape of a narrow, deep cleft containing the
glucose binding pocket. The binding site for the
activator was also localized in the hinge region, at a
distance from the glucose binding pocket. The struc-
ture was very similar to that of the C-terminal half of
hexokinase I crystallized in the presence of glucose,
called the closed form of hexokinase, and was there-
fore designated the “closed” form of GCK. By
contrast, GCK crystals formed without ligands dis-
28 P. B. Iynedjian Molecular physiology of glucokinase
played the so-called “super-open” conformation. In
this configuration, the cleft space between the two
domains of GCK was much broader, as the result of
two conformational changes compared to the closed
form: tilt and rotation of the small domain with
respect to the large domain, and extensive rearrange-
ment of the secondary structural elements of the small
domain itself [15].
The two X-ray structures of GCK could readily be
fitted into the mnemonic model of enzyme mecha-
nism. The super-open conformation represents the
low glucose affinity, catalytically inactive form of
enzyme. The closed conformation corresponds to the
high affinity form of GCK with glucose and ATP
bound at the active site during catalysis. An additional
conformation of GCK, intermediary between the
super-open and the closed states, was postulated for
the high affinity form of the enzyme free of substrates
or products. This putative conformation was desig-
nated the “open” form. The crystal structure of GCK
in the open form remains to be established. However,
X-ray diffraction studies of crystals of the C-terminal
half of hexokinase I formed in the absence of glucose
can serve as a model for the open form. Compared to
the closed form, the open form displayed a wider
interdomain cleft due to slight twist of the small
domain relative to the large domain, without change
in the internal topology of the small domain core [16] .
The transition between open and closed conforma-
tions would be very fast and easily reversible, in
contrast to the more complex molecular reorganiza-
tion involved in the slow transition from the super-
open to the open conformations. Indeed, the predic-
tion of an intermediary open state of GCK was borne
out in a recent study using targeted molecular
dynamics simulation to dissect the transition from
the closed to the super-open configurations of GCK
[17].
Regulation of glucokinase activity by protein-protein
interaction
Interaction with hepatic glucokinase regulatory pro-
tein. The best understood mechanism of rapid regu-
lation of GCK activity relies on protein-protein
interaction between GCK and a partner termed
GCK regulatory protein (GCKR). The association
between the two proteins is ligand-dependent and
inhibitory to GCK activity. Ligands for GCKR are
fructose-6-phosphate and fructose-1-phosphate,
which are mutual competitors. Binding of fructose 6-
phosphate to GCKR favors the GCKR-GCK inter-
action with a negative effect on enzyme activity, while
binding of fructose-1-phosphate weakens the GCKR-
GCK interaction and releases active GCK. Intra-
hepatic fructose-1-phosphate rises postprandially
after intestinal absorption of fructose and its con-
version to fructose-1-phosphate by liver fructokinase.
Conversely, inhibition of GCK by fructose-6-phos-
phate liganded to GCKR would provide a mechanism
for indirect negative feedback, since fructose 6-
phosphate is in equilbrium with glucose 6-phosphate
(the product of the GCK reaction) through the
phosphohexose isomerase step of glycolysis [18].
The conformational state of GCK preferred for
binding to GCKR is the super-open form. This was
suggested by experiments in a cell-free assay with
purified proteins, in which the association of GCK
with GCKR was estimated by co-immunoprecipita-
tion. The amount of GCK associated with GCKR was
reduced under conditions favoring the closed con-
formational state of GCK, that is during incubation at
elevated glucose concentrations in the presence of
ATP, or in the presence of glucose together with a
small molecule activator of GCK. This suggested that
the interface for binding to GCKR would be disrupted
by the conformational transition from the super-open
to the closed forms of GCK [19].
The reversible association of GCK with GCKR does
more than simply regulate the catalytic activity of
GCK, it also appears to underlie the intracellular
trafficking of GCK between cytoplasm and nucleus. In
transfected rat hepatocytes, GCK tagged with green
fluorescent protein was localized in the nucleus at low
glucose and was released to the cytoplasm at high
glucose, but GCK mutants with reduced affinity for
GCKR were predominantly cytoplasmic at low as well
as high glucose [20]. In cell lines devoid of GCK and
GCKR, such as COS-1, HeLa and HEK293T cells,
forced expression of GCK alone resulted in cytoplas-
mic localization, whereas co-expression with GCKR
resulted in nuclear accumulation of both proteins [20
22]. In a rat islet b-cell line with no endogenous
GCKR, doxycycline-inducible expression of GCKR
resulted in nuclear accumulation of GCKR together
with translocation of endogenous GCK from an
extranuclear localization to the nuclei [22].
A model (Fig. 1) to account for these results proposes
that GCKR in the free state would shuttle between
nucleus and cytoplasm. Under low glucose conditions,
the assembly of GCKR-GCK complex would result in
nuclear import and trapping of both proteins by
molecular mechanisms not yet completely elucidated.
Masking of a nuclear export sequence (NES) between
amino acids 300 and 310 of GCK sequence would be
part of this mechanism [21]. In metabolic states
accompanied by high glucose or fructose-1-phosphate
concentrations, and sufficient ATP levels [23], the
GCK-GCKR complex would dissociate and the NES
Cell. Mol. Life Sci. Vol.66, 2009 Review Article 29
of GCK would become functional, allowing for the
rapid export of GCK from nuclei. In the isolated
perfused liver system and in primary cultures of rat
hepatocytes, GCK and GCKR were indeed mainly
localized to the nuclei at glucose concentrations
around 5 mM. Higher glucose levels or fructose
addition to the medium caused a rapid translocation
of GCK to a predominant cytoplasmic localization
[24, 25]. This was accompanied to a lesser degree by
redistribution of GCKR to the cytoplasm [26, 27]. As
predicted by this model, mice with homozygous
inactivation of the GCKR gene had an extranuclear
localization of hepatic GCK under all metabolic
conditions [28, 29]. Interestingly, the steady state
amount of GCK in the liver was reduced about 50 % in
the GCKR knock-out animals, suggesting the possi-
bility that the interaction with GCKR might stabilize
GCK against degradation.
Recently, a minor fraction of GCK and GCKR in
hepatocytes was reported to be associated with the
mitochondrial fraction, raising the suggestion of
distinct pools of cytoplasmic GCK with specialized
metabolic functions [30]. However, this notion should
be considered with caution, because other authors
failed to detect GCK activity or immunoreactivity in
isolated mouse liver mitochondria [31].
Interaction with other proteins. The impact of the
GCKR model led several investigators to search for
new protein partners that might affect GCK activity or
localization, notably in the islet b-cells in which
GCKR is essentially absent. A list of GCK interacting
proteins with their major properties is given in
Table 1. For virtually all of these proteins other than
GCKR, additional investigations are necessary for
definitive proof that the proposed interaction with
GCK has an impact on GCK activity and/or cell
function.
The bifunctional enzyme 6-phosphofructo-2-kinase/
fructose-2,6-bisphosphatase (PFKFB) is responsible
for the regulated formation and degradation of
fructose 2,6-bisphosphate, a key allosteric activator
of the enzyme 6-phosphofructo-1-kinase, and thereby
of glycolysis. An interaction of the fructose-2,6-bi-
sphosphatase domain of PFKBP with GCK was
demonstrated in a yeast two-hybrid system [34].
Overexpression of PFKBP in insulinoma b-cells
resulted in increased GCK specific activity without
increase in the cellular content of GCK protein [35,
36]. Although provocative, the published studies did
not directly demonstrate an association of GCK with
PFKFB in mammalian cells, nor did they provide
insight into a possible mechanism for the apparent
increase in GCK activity in cells co-expressing
PFKBP. These issues should be further investigated
before a conclusive model of GCK regulation by
association with PFKBP becomes widely accepted.
A model in which neuronal nitric oxide synthase
(NOS) would serve to anchor GCK to the periphery of
insulin secretory granules in the basal state in bTC3
insulinoma cells was proposed by Rizzo and Piston
[37]. Autocrine activation of NOS by insulin during
Figure 1. Regulation of GCK activity and subcellular localization by interaction with GCKR. The scheme depicts hepatocytes after an
overnight fast (postabsorptive) and during the ingestive phase after a carbohydrate containing meal (postprandial). The nuclear
compartment (circle) communicates with the cytoplasmic space by nuclear pores. The blue and cyan ovals represent the large and small
lobes of GCK respectively. After carbohydrate ingestion, the glucose concentration in plasma and hepatocytes rises, as does the
concentration of fructose-1-phosphate. As a result,the GCKR (red)-GCK interaction is loosened, allowing GCK to bind glucose, adopt the
closed (catalytically active) conformation and exit from the nucleus to generate glucose-6-phosphate for glycogen synthesis and glycolysis.
30 P. B. Iynedjian Molecular physiology of glucokinase
glucose stimulated secretion would stimulate the
nitrosylation of GCK on a specific cysteine residue,
and its release from the granules to the cytosol. The
general validity of this model remains uncertain,
however, because glucose-induced translocation of
GCK from the secretory granules to the cytosol was
not confirmed in MIN6 insulinoma cells [38].
Human GCK was recently found to be polyubiquiti-
nated at multiple Lys residues in a cell-free system.
Additionally, a sequence motif compatible with an
ubiquitin interacting consensus was identified in the
C-terminal a-helix of human GCK. This motif was the
target for non-covalent binding of free pentaubiquitin
chains in a cell-free system. Ubiquitin chains added to
a GCK enzyme assay mixture resulted in an increased
Kcat [39]. The occurrence and physiological implica-
tion of this interaction in living cells were not
investigated.
Multiprotein assembly at the outer mitochondrial
membrane. In investigations of the function of the
BH3-only proapoptotic protein BAD, Danial and
colleagues described a large multimolecular assembly
comprising BAD, GCK and other proteins in a highly
purified mitochondrial fraction from the mouse liver
[40]. Analysis of the supramolecular complex using
immunochemistry and LC-tandem MS peptide anal-
ysis revealed that the complex contained GCK and
BAD, as well as the Wiskott–Aldrich family member
WAVE, the catalytic subunits of protein phospatase 1
(PP1) and of protein kinase A (PKA). The essential
protein in nucleating the complex appeared to be
BAD, since the complex was absent in BAD knock-
out (BAD-/-) mice. The oxidative metabolism of
glucose was reduced in hepatocytes from BAD-/-
mice, suggesting a function of the complex in meta-
bolic regulation. However, caution is warranted,
because GCK associated with mitochondria was a
very small fraction of the total [40], if present at all
[31]. Complicating the matter further, the role and
regulation of the protein kinase and phosphatase
associated with the complex remain elusive.
The relevance of the Danial complex was recently
extended to the b-cells of the islets of Langerhans [41] .
Perifused islets from BAD-/- mice displayed glucose
unresponsive insulin secretion. Total assayable GCK
activity in islet homogenates from BAD-/- mice was
reduced to less than 30% of the wild-type level.
Glucose-sensitive insulin release was restored after
treatment of islets from BAD-/- animals with cell-
permeable synthetic peptides corresponding in se-
quence to a BH3 a-helix of BAD. Peptides harboring a
phosphomimetic Ser 155 to Asp mutation were active,
while a peptide with an Ala mutation was ineffective.
Peptides active in the insulin secretion assay also
induced strong increases in GCK activity when added
to INS-1 insulinoma cells in culture. Furthermore,
photoactivatable BAD peptides were shown to cross-
link with GCK in homogenates of INS-1 cells. It is
worth pointing out that a much larger fraction of
cellular GCK appeared to be under BAD control in
the islets than in the liver. Whether most or all of the
GCK interacting with BAD resided at the outer
mitochondrial membrane in islet cells was not exper-
imentally demonstrated.
Table 1. Protein interaction partners of glucokinase.
Name Method of identification Biochemical
effects Cellular effects References
glucokinase regulatory
protein (GCKR) protein purification from liver extract inhibits GCK
activity regulates GCK activity and
localization in hepatocytes reviewed
in [18]
glucokinase-associated-
phosphatase yeast two-hybrid system increases GCK
activty unknown [32]
propionyl-CoA
carboxylase b-subunit yeast two-hybrid system increases GCK
affinity and
activity
unknown [33]
6-phosphofructo-2-
kinase/fructose-2,6-
bisphosphatase
screening of random peptide display phage
library for binding to GCK no effect in GCK
activity assay overexpression in insulinoma
cells increases GCK activity [34 36]
neuronal nitric oxide
synthase co-immunoprecipitation nitrosylates
GCK anchors GCK to the periphery
of insulin secretory granules [37]
pentaubiquitin identification of ubiquitin interacting motif
in GCK increases GCK
affinity and
activity
unknown [39]
bad peptide analysis of multiprotein complex in
liver mitochondrial fraction,
immunochemistry
increases GCK
activity regulates GCK activity and
localization in hepatocytes and
b-cells
[40, 41]
Cell. Mol. Life Sci. Vol.66, 2009 Review Article 31
Regulation of glucokinase gene expression
Tissue-specific promoters. A defining feature of the
GCK gene in mammalian species is the presence of
two alternative promoters. The promoters are sepa-
rated by approximately 30 kbp of genomic DNA and
responsible for transcription initiation in a mutually
exclusive manner in distinct tissues. The upstream
promoter and adjacent leader exon drive the synthesis
of GCK mRNA in non-hepatic tissues, that is islet
cells, enteroendocrine cells, specialized glucose-sensi-
tive neurons in the central nervous system (CNS), and
additional scattered cells in various tissues [42]. This
promoter is denominated the neuroendocrine pro-
moter. By contrast, the downstream so-called liver
promoter and its associated leader exon serve for
transcription initiation in hepatocytes only. Each of
the leader exons specifies the 5 untranslated regions
of the mRNAs as well as the first 15 codons for the
GCK protein sequences, which therefore differ at
their NH2-terminal ends in liver and other tissues.
During splicing of the primary transcripts, the donor
splice sites at the end of the neuroendocrine or liver
leader exons are ligated to the acceptor site of a
common exon 2 and further splicing incorporates
common exons 3 10 in the two types of mRNAs.
Thus, the GCK enzymes in neuroendocrine cells and
hepatocytes have a common amino acid sequence
over most of their length. There is currently no known
functional difference in the GCK enzymes arising
from the short distinct leader sequences.
Alternative promoters allow for the control of tran-
scription initiation by distinct sets of cis-acting regu-
latory elements in different tissues. In early experi-
ments, Magnuson and colleagues showed that a 294 bp
rat DNA fragment of the neuroendocrine promoter
activated the expression of luciferase in insulinoma
cells, but not 3T3-preadipocytes [43]. In transgenic
mice, the same DNA fragment drove the expression of
a human growth hormone reporter gene in pancreatic
islet cells, intestinal cells and brain cells [42]. Two
major functional cis-acting elements were delineated
in this promoter fragment [43]. One element bound
proteins which did not match known transcription
factors [44], whereas the other element was identified
as a binding site for Beta2/NeuroD, a basic helix-loop-
helix zipper (bHLHz) transcriptional activator im-
portant in the ontogeny of the endocrine pancreas
[45]. Another key transcription factor of early pan-
creas development, Pdx-1, was proposed to activate
the neuroendocrine GCK promoter [46], but this
notion was subsequently dispelled [47].
There are no published investigations using transgenic
mice to define the regulatory sequences responsible
for the hepatocyte-specific activity of the liver GCK
promoter. In transient transfection assays, a 200 bp
liver promoter fragment from the rat gene was
competent to stimulate the expression of a luciferase
reporter in primary hepatocytes and hepatoma culture
cells, but not in insulinoma cells [48, 49]. More
importantly, a 300 bp enhancer active only in primary
hepatocytes was mapped between nucleotides -1000
and -700 with respect to the start of transcription of the
rat gene and at a similar position in the human gene.
This sequence exhibited 6 DNAse I protection foot-
prints, but the identity of the binding proteins was not
elucidated [48].
Transcriptional induction of gene expression by
insulin in liver. In the liver, expression of GCK is
strictly dependent on the presence of insulin. Thus,
GCK mRNA and protein disappear from the livers of
insulino-deficient rats and are restored after insulin
treatment [50]. In cultured rat hepatocytes, the effect
of insulin as inducer of GCK was shown to be
primarily at a transcriptional level, triggering a 15-
to 30-fold increase in GCK mRNA in three hours, and
taking place in glucose-free as well as glucose-
containing medium without any synergy by addition
of glucose [6]. Induction of GCK by insulin and
repression by glucagon were also documented in
hepatocytes isolated from the human liver [51].
Insulin induction of GCK in hepatocytes was shown to
be suppressed by the inhibitors of phosphoinositide 3-
kinase (PI3-kinase) wortmannin and LY294002, im-
plying that activation of PI3-kinase following insulin
binding to its receptor was essential. Protein kinase B
(PKB, also called Akt) is a major protein kinase
activated as a result of PI3-kinase stimulation. To test
for a possible role of PKB in mediating GCK
induction, hepatocytes were transduced for expres-
sion of an estrogen receptor-PKB chimeric protein
that could be conditionally activated by tamoxifen.
The transduced hepatocytes reponded to the addition
of tamoxifen by a rise of GCK mRNA, mimicking the
effect of insulin [52]. In separate experiments, insulin
induction of GCK mRNAwas abrogated by inhibitors
of PKB activation [53, 54] . Additionally, GCK mRNA
and protein were induced in the livers of streptozo-
tocin diabetic mice following in vivo transduction of
adenoviruses encoding kinase active versions of
PFKFB, a maneuver that unexpectedly triggered the
activation of PKB [55]. A role of PKB in insulin
induction of GCK was questioned by Matsumoto and
colleagues [56], who observed that they could inhibit
insulin-mediated induction of GCK mRNA in hep-
atocytes by expressing a dominant negative NH2-
terminal fragment of insulin receptor substrate-1
(IRS-1), while PKB activation apparently remained
unaffected. An explanation for this negative result
32 P. B. Iynedjian Molecular physiology of glucokinase
might be that GCK induction is mediated by a minor
pool of PKB functionally associated with IRS-1,
making a small contribution in assays of total cellular
PKB activity and therefore difficult to quantify.
The transcriptional regulators which respond to the
insulin activation of PKB (and perhaps of accessory
signaling cascades) and stimulate transcription at the
liver GCK promoter remain elusive. The sterol-
regulatory element binding protein SREBP1c, a
master regulator of lipogenic enzymes such as fatty
acid synthase (FAS) and acetyl-CoA carboxylase
(ACC), was proposed to be a mediator of insulin
induction of GCK. Briefly, SREBP1c itself was shown
to be induced at the transcriptional level by insulin in
hepatocytes [57, 58]. Overexpression of SREBP1c in
primary hepatocytes was accompanied by an elevation
of GCK mRNA and, conversely, a dominant negative
mutant of SREBP1c inhibited insulin-dependent in-
duction of GCK mRNA [59]. More recently, inves-
tigators identified a tandem DNA element around
position -200 in the rat liver GCK promoter, which
could bind bacterially expressed mature SREBP1c in
DNAse protection assay and electrophoretic mobility
shift assay (EMSA) [60]. However, in other inves-
tigations, inducible expression of mature SREBP1c in
hepatocytes failed to produce a significant increase of
GCK mRNA [61]. In SREBP1c ko mice, GCK
mRNA was induced normally during the fasting-
refeeding transition [62]. Moreover, liver transduc-
tion of adenovirus encoding active PKB (i.e. mimick-
ing insulin signaling) resulted in strong induction of
hepatic GCK mRNA in normal as well as SREBP1 ko
mice [63]. In our own experiments in insulin stimu-
lated hepatocytes, SREBP1c failed to bind to the liver
GCK promoter, while its binding to the FAS promoter
was readily discerned by chromatin immunoprecipi-
tation [64]. Also, insulin produced a more than 10-fold
augmentation of GCK mRNA amount in a time frame
in which there was strictly no increase in transcrip-
tionally mature SREBP1c protein. Vice-versa, activa-
tion of the nuclear receptor liver X receptor (LXR)
with a synthetic ligand produced a strong induction of
SREBP1c and its target gene FAS without any rise of
GCK mRNA [64]. Together, these results seem to
virtually rule out a determinant role of SREBP1c in
the rapid induction of liver GCK by insulin in rat
hepatocytes.
Binding sites for the transcriptional activators hep-
atocyte nuclear factor-4 (HNF-4) and hypoxia-indu-
cible factor-1a (HIF-1a) were identified within the
first 100 bp of the liver GCK promoter [65, 66]. Co-
expression of HNF-4 and HIF-1awith GCK reporter
plasmids in primary rat hepatocytes resulted in
additive transactivation of the GCK promoter,
which was further augmented by expression of the
co-activator p300. More importantly, co-transfection
of plasmids encoding constitutively active PI3K or
PKB stimulated GCK promoter activity, as long as an
intact binding site for HIF-1awas present in the
promoter DNA fragment. Together, these findings
suggested that co-operation between HNF-4, HIF-1a
and p300 at the liver GCK promoter might be
important for induction of the gene in response to
acute insulin signaling [66]. The proposal that HIF-1a
might be a mediator in insulin signaling is significant,
because hypoxia and insulin were previously shown to
synergize in inducing GCK gene expression in cul-
tured hepatocytes [67]. This type of synergy might in
part account for the enrichment of insulin inducible
enzymes such as GCK in the underoxygenated
perivenous region of the liver lobule, compared to
the periportal region [68]. Based on findings in
another system, insulin might hypothetically induce
HIF-1atarget genes by relieving HIF-1a from co-
repression by a forkhead box class O (FOXO) factor,
through a mechanism involving PKB-mediated phos-
phorylation and functional inactivation of FOXO
[69]. However, arguing against this possibility, levels
of hepatic GCK mRNA in adult mice with liver
specific knockout of FOXO1 were similar to those of
control mice in both the fasted and fed states [70].
Hormone and nutrient regulation of glucokinase gene
expression in islets of Langerhans. Glucose refeeding
after a fast is an experimental paradigm often used to
study acute hyperinsulinemia in the intact animal.
Consistent with data in cultured hepatocytes, GCK
mRNA was virtually absent from the livers of fasted
rats and accumulated massively within hours of
feeding a glucose-enriched diet [71]. By contrast,
GCK mRNA and protein levels in islets of Langer-
hans were not significantly depressed in fasted rats
compared to freely fed animals and, more importantly,
were not acutely induced during the fasting-refeeding
transition [72]. Thus, constitutive expression of GCK
in the pancreatic islets contrasted with inducible GCK
gene expression in the liver.
In the MIN6 insulinoma cell line, GCK mRNA was
higher during culture at 30 mM than at 3 mM glucose
concentration, or following the addition of subnano-
molar concentrations of insulin to the medium [73].
However, luciferase reporter plasmids driven by the
neuroendocrine GCK promoter were unaffected after
a rise in glucose or insulin concentrations [74],
suggesting that, in insulinoma cells, the major effects
of glucose or insulin might be to stabilize GCK mRNA
against degradation rather than stimulate transcrip-
tion from the neuroendocrine promoter. In rat INS-1
insulinoma cells, GCK mRNA and protein levels were
increased about 50% after 24 hours of culture in the
Cell. Mol. Life Sci. Vol.66, 2009 Review Article 33
presence of insulin-like growth factor-1 (IGF-1),
consistent with a modest transcriptional effect of
IGF-1. An inhibitory cis-acting DNA element target-
ed by FOXO1 was identified at position 500 in the
neuroendocrine promoter. Thus, FOXO1 might act in
this system as a repressor, itself negatively regulated
by IGF-1 [75].
Metabolic role of hepatic glucokinase
Regulatory role of glucokinase at the hepatocellular
level. The liver is equipped with an insulin-independ-
ent facilitative glucose transport system (now known
to be mostly by GLUT2) with apparent affinity
constant of 17 mM and high transport rate, allowing
extremely rapid equilibration of the glucose concen-
tration across the hepatocyte plasma membrane [76].
Unlimiting glucose transport, and the unique glucose
affinity of GCK (which is the predominant hexokinase
isoenzyme in the liver), ensure that fluctuations of the
plasma glucose concentration are rapidly translated
into changes in the rate of glucose phosphorylation
inside the hepatocyte. If, in turn, the GCK reaction
controls the substrate flux along the entire length of
the pathways from glucose to final products, the
overall metabolic rate in these pathways will be
regulated autonomously in response to changes in
the plasma glucose concentration.
In hepatocytes with adenovirus mediated overexpres-
sion of GCK, compared to hepatocytes transduced
with control vectors, rates of glycogen synthesis were
augmented in proportion to GCK activity [77, 78].
This effect was attributed to an increase in the cellular
content of glucose-6-phosphate, a product of the GCK
reaction and an allosteric activator of glycogen
synthase [79]. Overexpression of GCK also resulted
in proportionate stimulation of glycolysis and glucose
oxidation [80, 81]. As stated above, the fact that GCK
exerts dominant control on glycogen synthesis and
glycolysis has a crucial physiological implication: it
confers to hepatocytes the ability for cell autonomous
regulation of glucose metabolism in response to
fluctuations of the plasma glucose.
Regulatory role of liver glucokinase at the organismal
level. The impact of altering liver glucokinase activity
on whole body glucose homeostasis was illustrated in
transgenic mice with an extracopy of the GCK gene
locus. These animals had an approximately 50%
increase in hepatic GCK activity over the wild-type
level and a 25% reduction in plasma glucose in the
postabsorptive state, reflecting increased glucose
clearance. During an hyperglycemic clamp, overall
glucose utilization was similar in control and trans-
genic mice, in spite of plasma insulin levels that were
40 % lower in the transgenics. Liver glycogen synthesis
over the period of hyperglycemia was increased about
3.5 times in transgenic mice compared to wild-type
[82]. The level of immunodetectable GCK was lower
in the pancreatic islets of transgenic mice with addi-
tional GCK gene copies than in wild-type animals,
perhaps owing to destabilization of the GCK protein
in islets at low glycemic levels.
Another mouse model with overexpression of liver
GCK was produced by transgenesis of GCK cDNA
fused to the phosphoenolpyruvate carboxykinase
(PEPCK) promoter. This construct allowed for high
GCK activity to be maintained in the livers of trans-
genic mice made insulino-deficient by streptozotocin
treatment, while virtual suppression of hepatic GCK
was noted in nontransgenic littermates [83]. Transgenic
mice were largely protected against streptozotocin-
induced hyperglycemia. They maintained normal levels
of liver glycogen and lactate, compared to severe
depletion of these metabolites in the livers of non-
transgenic diabetic mice, consistent with a determinant
role of GCK in hepatic glycogen synthesis and glycol-
ysis.
The effects of loss of hepatic GCK function on glucose
homeostasis were investigated using a gene targeting
system allowing for liver-specific inactivation of the
GCK gene [84]. Mice with homozygous GCK gene
deletion in the hepatocytes survived to adult age with
virtually complete absence of GCK enzyme activity in
liver tissue. In the freely fed state, the mice exhibited
modest hyperglycemia with relative hyperinsulinemia
compared to controls, indicative of an insulin resistant
state. Liver glycogen in the freely fed state was
normal, implying that gluconeogenesis (and perhaps
high affinity hexokinases) produced sufficient glu-
cose-6-phosphate in the basal state to ensure glycogen
formation and compensate for the ablation of GCK.
However, during a hyperglycemic clamp, the rate of
liver glycogen synthesis failed to rise and whole body
glucose turnover was lower than in control animals.
Interestingly, glucose-induced insulin secretion during
hyperglycemia was markedly diminished in liver
knockout mice, possibly reflecting functional damage
to the b-cells of the islets of Langerhans due to chronic
hyperglycemia and secretory overstimulation [84].
The role of liver GCK in restraining hepatic glucose
output in response to hyperglycemia was also high-
lighted in mice with global heterozygous inactivation
of the GCK gene [85]. These mice exhibited a 50%
reduction of GCK activity in the livers, and presum-
ably in islets and other neuroendocrine cells as well.
Glucose metabolism was investigated by the glucose
clamp technique combined with pancreatic clamping
with somatostatin, in order to maintain circulating
34 P. B. Iynedjian Molecular physiology of glucokinase
insulin concentrations at the same fixed level in
knockout and wild-type mice. Hyperglycemia caused
a 10% reduction of the rate of hepatic glucose
production in knockout mice, compared to 40% in
controls, demonstrating that liver GCK is instrumen-
tal in insulin-independent inhibition of hepatic glu-
cose production in response to high glucose [85]. Since
GCK gene inactivation was ubiquitous in these mice,
it cannot be excluded that decreased GCK activity at
extrahepatic sites, for example in the hypothalamus
(see below), might have compromised a neurally-
mediated mechanism contributing to the suppression
of hepatic glucose production during hyperglycemia.
Metabolic role of islet b-cell glucokinase
Regulatory role of glucokinase in the b-cell. Insulin
secretion elicited by hyperglycemia requires stimula-
tion of glucose metabolism in the b-cells. Coupling
between metabolism and secretion relies on the
generation of metabolic products, most prominently
ATP generated by glycolysis and glucose oxidation,
which act as upstream signals in cascades leading to
the stimulation of the exocytic machinery [86, 87]. For
instance, a rise in the cellular ATP/AMP ratio causes
the inhibition of ATP-sensitive K+channels in the
plasma membrane, leading to membrane depolariza-
tion and activation of voltage-sensitive Ca++ channels.
The influx of extracellular Ca++ ions then somehow
promotes the exocytosis of insulin secretory granules.
The ability of the b-cells to increase their rate of
glucose metabolism in response to a rise in the
extracellular glucose concentration is basic to the
regulation of insulin secretion. The biochemical
properties which allow glucose to regulate its own
metabolism, and thereby insulin secretion, are the
same in the b-cells and in hepatocytes. The first
property is unrestricted glucose transport to ensure
rapid glucose equilibration across the plasma mem-
brane [88, 89]. The second aspect is that glucose
phosphorylation is predominantly catalyzed by GCK
[90], which provides for concentration-dependent
adjustment of the rate of the initial reaction of glucose
metabolism. The third property is a metabolic organ-
ization in which flux along the entire pathways of
glycolysis and glucose oxidation is mostly controlled
at the initial reaction, that catalyzed by GCK.
The extent to which glucose phosphorylation does
control the rate of glycolysis in insulin secreting cells
was investigated in INS-1 derived insulinoma cells
that had been engineered for doxycycline inducible
GCK expression. Moderate increases in GCK activity
elicited by small doses of doxycycline were accom-
panied by quantitatively comparable fractional in-
creases in glycolytic rate at physiologically relevant
extracellular glucose concentration. The fact that the
ratio of fractional increase in glycolytic flux on
fractional increase in GCK activity was close to
unity showed that GCK exerted overwhelming con-
trol over the glycolytic rate [91]. Not surprisingly, at
very high levels of GCK overexpression and especially
at high glucose levels, a limit to the increase in
glycolytic flux was attained, indicating that one or
several steps in the lower part of the glycolytic
pathway became limiting [91, 92]. The glyceralde-
hyde-phosphate dehydrogenase reaction was incrimi-
nated as such a step, due to NAD+depletion [93].
Regulatory role of b-cell glucokinase at the organ-
ismal level. A strategy targeting the neuroendocrine
promoter region of the GCK gene was used to
generate mice with GCK deficiency in the b-cells of
the islets of Langerhans (and other cell-types utilizing
the same promoter) [94]. Mice homozygous for the
deleted allele developed severe diabetes and died
during the first week after birth. Glucose-stimulated
insulin secretion was completely absent in islets from
newborn null mice. Mice heterozygous for the deleted
allele survived to adulthood with slight hyperglyce-
mia, intolerance to a glucose load and reduced
glucose-stimulated hyperinsulinemia. As expected,
GCK activity in islets was reduced to about 50% of
the wild-type level in the heterozygotes. Isolated islets
form heterozygous mice exhibited impaired glucose
stimulated insulin release, most apparent at a glucose
concentration around 10 mM [94]. Similar findings
were reported in mice made deficient in islet GCK
activity by a LoxP strategy using Cre recombinase
expressed from the rat insulin II gene promoter. The
mice with heterozygous inactivation of b-cell GCK
were markedly impaired in ability to raise plasma
insulin during a hyperglycemic clamp in comparison
to controls [84]. Collectively, these data underlined
the role of islet GCK in glucose-induced insulin
secretion and maintenance of glucose homeostasis in
the whole animal. Whether the ablation of GCK in
other neuroendocrine or insulin expressing cells might
have contributed to the observed phenotype in the
above experiments remains to be determined.
Glucokinase in the brain
Transgenic mice harboring a chimeric human growth
hormone (hGH) structural gene transcribed from the
rat neuroendocrine GCK promoter had abundant
hGH in their pancreatic islets as expected, and more
surprisingly, also displayed strong immunoreactivity
to hGH in isolated hypothalamic cells [42]. Following
Cell. Mol. Life Sci. Vol.66, 2009 Review Article 35
this discovery, Jetton and colleagues detected GCK
mRNA by reverse transcription-polymerase chain
reaction (RT-PCR) in several regions of the rat brain
and localized GCK transcripts as well as immunor-
eactive protein in situ in the hypothalamus [42].
Subsequently, the neuroendocrine form of GCK
mRNA was selectively identified by RT-PCR assay
as well as in situ hybridization in the ventromedial and
arcuate nuclei of the rat hypothalamus [95]. At the
protein level, immunoblotting and enzyme assays
confirmed the presence of GCK in rat hypothalamic
extracts [96].
In the ventromedial hypothalamus, GCK mRNA was
shown to be expressed in neurons synthesizing pre-
proopiomelanocortin (POMC) and neuropeptide Y
(NPY)/agouti-related peptide (AgRP), which play
critical roles in neural pathways involved in the
regulation of food intake and energy expenditure.
Experiments using primary cultures of rat hypothala-
mic neurons incubated with metabolic inhibitors such
as mannoheptulose or alloxan suggested that GCK
activity was instrumental for the glucose response of
glucose-excited and glucose-inhibited neurons as well
[97]. Transcript profiling at the single cell level
confirmed the presence of GCK mRNA in approx-
imately 50 % of glucose-excited and glucose-inhibited
neurons in the ventromedial area of the hypothal-
amus. Some of these neurons also contained mRNAs
for subunits of ATP-sensitive K+channels [98]. In
primary cultures of neurons from the rat ventromedial
hypothalamus, treatment with specific small interfer-
ing RNA (siRNA) to knock-down GCK resulted in
the virtual disappearance of excitatory as well as
inhibitory responses to glucose elevation [99]. Col-
lectively, these findings supported the idea that the
glucose sensor of at least a fraction of ventromedial
hypothalamic neurons was GCK, functioning as the
rate determining enzyme of glucose metabolism and
controlling neuronal excitability. However, this pro-
posal raises puzzling questions. First, the interstitial
glucose concentration in brain in the postabsorptive
state is around 1.5 mM, well below the optimal range
of concentrations where GCK displays ultrasensitive
kinetics with respect to glucose concentration. The
second key issue is that of coupling between glucose
metabolism and neuronal activity. Many neurons are
equipped with ATP-sensitive K+channels, and the
excitability of such neurons might then be controlled
by the rate of glucose metabolism and the attending
fluctuations in cellular ATP content, in ways reminis-
cent of the islet b-cells [100]. Other glucose-derived
metabolites with a signaling role might be intermedi-
ates of the lipogenic pathway [101].
Molecular medicine of glucokinase
Maturity onset diabetes of the young (MODY2).
Approximately 200 mutations of the human GCK
gene, including missense, non-sense, and splice site
mutations, were described in the literature. The
majority of these mutations cause a form of familial
hyperglycemia or mild diabetes, transmitted as an
autosomal dominant trait and often diagnosed before
25 years of age. Based on these characteristics,
diabetes associated with GCK mutations was included
in the group of maturity onset diabetes of the young
(MODY), and was named MODY2. Clinically,
MODY2 was defined as a syndrome of impaired
fasting glucose level or mild diabetes, with hyper-
glycemia maintained at stable level for many years,
little propensity for late neurovascular complications
and often adequate glycemic control with diet alone
[102, 103] . Several cases of MODY2 were diagnosed in
women with a history of gestational diabetes [104].
Recently, a new nomenclature was proposed for the
MODY syndromes, MODY2 being replaced by the
descriptive name “familial mild fasting hyperglyce-
mia” [105]. Rare cases of infants inheriting GCK
mutations from two heterozygous parents were de-
scribed in the literature. They presented an autosomal
recessive form of permanent neonatal diabetes melli-
tus (PNDM) [106].
Clinical investigations in patients with MODY2 re-
vealed disorders of liver metabolism and pancreatic
islet function predicted from basic studies in cells and
animals. Net synthesis of glycogen in the liver after
each of three daily meals was decreased by about 50 %
in patients compared to controls [107]. Additionally,
the suppression of hepatic glucose output by hyper-
glycemia was deficient in patients with a GCK
mutation [108]. Patients with a variety of missense
or nonsense GCK mutations exhibited postprandial
plasma insulin levels comparable to controls, although
glucose concentrations were higher in the patients, a
finding compatible with a loss of sensitivity of the b-
cells to glucose. More specifically, during a ramping
glucose infusion, the glucose dose-response curve for
insulin secretion was found to be shifted to the right in
the MODY2 patients [109].
The biochemical properties of many GCK mutants
identified in MODY2 individuals were investigated
using bacterially produced protein. A wide range of
abnormalities affecting the turn-over number of the
enzyme, affinity for the substrates, and cooperativity
with respect to glucose were described [110 112].
Mathematical models based on the kinetics properties
were used to predict the threshold of glucose-stimu-
lated insulin secretion by b-cells harboring a given
heterozygous mutation. On average, prediction for a
36 P. B. Iynedjian Molecular physiology of glucokinase
number of mutants was in agreement with the level of
fasting hyperglycemia noted in cohorts of MODY2
patients [111]. Some mutants were found to have
unaltered kinetic properties compared to the wild-
type enzyme, but to display accelerated temperature-
dependent loss of activity in the test tube, in one case
correlating with decreased protein half-life in trans-
duced cells [111, 113].
Several mouse lines with heritable hyperglycemia
were produced through phenotype-driven mutagene-
sis programs using the potent mutagen N-ethyl-N-
nitrosourea (ENU). Interestingly, the majority of
hyperglycemic lines that were genetically character-
ized so far harbored mutations in the GCK gene [114,
115], several of which corresponded to mutations that
had been described in pedigrees with MODY2 or in
one case in PNDM [116]. These mice displayed
spontaneous hyperglycemia in the freely fed state,
and glucose intolerance accompanied by relative
hypoinsulinemia after an oral glucose load, similar
to abnormalities found in human MODY2 patients
[116].
Persistent hyperinsulinemic hypoglycemia of infancy
(PHHI). A small number of heterozygous missense
GCK mutations were identified as probable cause of
hyperinsulinism in infancy, a rare syndrome charac-
terized by chronic low fasting plasma glucose and
bouts of symptomatic hypoglycemia [117]. Kinetic
studies of the mutant enzymes revealed an increase in
glucose affinity with or without increased turnover
number (kcat) [118 120] . A gene knock-in strategy
was used to produce mice with a GCK allele encoding
a hyperactive mutant GCK with increased affinity for
glucose. The mice were hyperinsulinemic relative to
their plasma glucose levels. However, the severity of
hypoglycemia was less than expected, suggesting a
compensatory mechanism in the liver. Paradoxically,
liver GCK activity and protein content in these mice
were lower than normal, suggesting the possibility of a
shorter half-life of the mutant enzyme due to impaired
interaction with GCKR [121].
Small-molecule activators of glucokinase as potential
antidiabetic drugs
Type 2 diabetes melltus (T2D) is a common disease of
multifactorial etiology, in which both a genetic pre-
disposition determined by multiple genes, and lifestyle
factors such as excess calorie intake and low physical
activity, play causative roles. The increasing preva-
lence of T2D, and its serious consequences on health
at both the individual and public levels, create
incentives for the development of novel antidiabetic
drugs. Physiopathologically, T2D consists of defective
insulin action in target tissue, including reduced
insulin suppression of liver glucose production, and
defective insulin secretion and other functions in the
b-cells of the islets of Langerhans [122]. The central
role of GCK in facilitating glucose disposal by the liver
on the one hand, and insulin secretion by the islets of
Langerhans on the other hand, provided a strong
rationale for the search of small molecule activators of
GCK in drug discovery programs aimed at developing
a new class of antidiabetic drugs. Such programs were
implemented by several major drug companies, and a
limited number of molecules acting as GCK activators
are currently in early stages of clinical trials.
Small molecule activators of GCK emerged from a
high throughput screening strategy based on the
capacity of chemicals to increase the activity of
recombinant GCK in an enzyme assay mixture con-
taining GCKR. The first promising compound was
described in 2003 by Grippo and colleagues [123] and
shown actually to be a direct activator of GCK in the
absence of GCKR. This compound, designated RO-
28-1675, had the dual effects of increasing both the
affinity for glucose and the turnover number of GCK,
that is, it shifted the glucose S0.5 of GCK from 8.5 mM
to 0.8 mM and increased the kcat by approximately
80% at maximal concentration. When added to
isolated rat islets of Langerhans, RO-28-1675 stimu-
lated the rate of glycolysis, lowered the threshold for
glucose-stimulated insulin secretion and increased the
maximal secretory response. A single oral adminis-
tration of RO-28-1675 to normal or obese mice had a
clear-cut hypoglycemic effect accompanied by hyper-
insulinemia. Treated mice displayed improved glucose
tolerance after an oral glucose load. Endogenous
glucose production was reduced during a hyperglyce-
mic hyperinsulinemic clamp in RO-28-1675-treated
rats compared to controls, consistent with an effect of
the drug on hepatic GCK activity [123]. A variety of
chemically distinct small molecule activators of GCK
were subsequently discovered in several laboratories.
The biochemical and cellular effects, as well as
biological actions in experimental animals, were
essentially similar to those of the prototype RO-28-
1675 [19, 124 126]. An expert review of the current
status of GCK activators in drug development was
recently published [127].
An implicit hope in developing activators of the
physiological glucose sensor was the idea of restoring
“physiological” glucose-sensing, particularly in the
islet b-cells. However, the strong increase in the
glucose affinity of GCK induced by all activators
known to date might drastically lower the threshold
for insulin release and cause maximal insulin secretion
at low plasma glucose levels. Whether the design of
Cell. Mol. Life Sci. Vol.66, 2009 Review Article 37
drug preparations without liability of inducing serious
hypoglycemia is possible remains to be seen. Another
concern with GCK activators is related to the activa-
tion of hepatic lipogenesis leading to hepatosteatosis
(and associated insulin resistance), as was noted in
transgenic mice with hepatic overexpression of GCK
[128, 129]. Finally, in a more general vein, the long
term benefit of using insulinotropic agents in the
treatment of T2D (at least early in the disease) was
recently questioned, because of concerns about the
possibility of precipitating islet failure through hyper-
secretory activity [130].
Conclusions and perspectives
The evidence presented in this review should leave
little doubt about the status of GCK as an eminent
player in the regulation of glucose metabolism and
homeostasis. This status relies on the mechanism
whereby GCK operates as an ultrasensitive glucose
sensor, recently illuminated by structural data. The
localization of such a sensor inside, rather than on the
surface, of cells is striking. It is increasingly clear that
the traffic of GCK between distinct subcellular sites,
regulated by reversible interactions with diverse
scaffold proteins, might contribute to shape the
metabolic activity of cells, but by which exact molec-
ular mechanisms remains to be explained, notably in
the pancreatic b-cells. In hypothalamic nuclei and
other brain areas, it will be important to resolve the
apparent paradox of GCK operating under less than
optimal glucose concentrations for substrate regula-
tion of the enzyme reaction. Some of the most
intriguing unanswered questions pertain to the regu-
lation of GCK gene expression in hepatocytes. The
transcriptional induction of GCK is a dramatic effect
of insulin in the liver, perhaps the best example of
induction of any gene in any cell-type by insulin.
However, details about the regulatory proteins of
transcription which mediate this effect are still a black
box. The burstlike pattern of regulation of GCK
transcription in rat liver with very large, rapid and
transient inductive effects of insulin for a protein with
relatively slow turnover raises puzzling questions with
respect to the biological advantage that might arise
from this type of regulation. In the diabetes field, the
development of GCK activators through the preclin-
ical phase has generated exciting results, but also
illustrated some liabilities. If these can be overcome,
drugs with a novel mechanism of action will become
available as an additional option in the panoply of oral
antidiabetics.
Acknowledgements. The long term support of research in the
authors laboratory by the Swiss National Science Foundation is
gratefully acknowledged.
1 Cardenas, M. L., Cornish-Bowden, A. and Ureta, T. (1998)
Evolution and regulatory role of the hexokinases. Biochim.
Biophys. Acta 1401, 242– 264.
2 Kawai, S., Mukai, T., Mori, S., Mikami, B. and Murata, K.
(2005) Hypothesis: structures, evolution, and ancestor of
glucose kinases in the hexokinase family. J. Biosci. Bioeng. 99,
320330.
3 Ronimus, R. S. and Morgan, H. W. (2004) Cloning and
biochemical characterization of a novel mouse ADP-depend-
ent glucokinase. Biochem. Biophys. Res. Commun. 315, 652
658.
4 Froguel, P., Vaxillaire, M., Sun, F., Velho, G., Zouali, H. ,
Butel, M. O., Lesage, S., Vionnet, N., Clement, K., Fouger-
ousse, F. and et al. (1992) Close linkage of glucokinase locus
on chromosome 7p to early-onset non-insulin-dependent
diabetes mellitus. Nature 356, 162– 164.
5 Hattersley, A. T., Turner, R. C., Permutt, M. A., Patel, P.,
Tanizawa, Y., Chiu, K. C., ORahilly, S., Watkins, P. J. and
Wainscoat, J. S. (1992) Linkage of type 2 diabetes to the
glucokinase gene. Lancet 339, 13071310.
6 Iynedjian, P. B. (1993) Mammalian glucokinase and its gene.
Biochem. J. 293 (Pt 1), 1– 13.
7 Printz, R. L., Magnuson, M. A. and Granner, D. K. (1993)
Mammalian glucokinase. Annu. Rev. Nutr. 13, 463– 496.
8 Wilson, J. E. (2003) Isozymes of mammalian hexokinase:
structure, subcellular localization and metabolic function. J
Exp Biol 206, 20492057.
9 Cardenas, M. L., Rabajille, E. and Niemeyer, H. (1978)
Maintenance of the monomeric structure of glucokinase
under reacting conditions. Arch. Biochem. Biophys. 190, 142
148.
10 Moukil, M. A. and Van Schaftingen, E. (2001) Analysis of the
cooperativity of human beta-cell glucokinase through the
stimulatory effect of glucose on fructose phosphorylation.
J.Biol. Chem. 276, 3872– 3878.
11 Monasterio, O. and Cardenas, M. L. (2003) Kinetic studies of
rat liver hexokinase D (glucokinase) in non-co-operative
conditions show an ordered mechanism with MgADP as the
last product to be released. Biochem. J. 371, 29 38.
12 Storer, A. C. and Cornish-Bowden, A. (1977) Kinetic evi-
dence for a mnemonical mechanism for rat liver glucokinase.
Biochem. J. 165, 61– 69.
13 Heredia, V. V., Thomson, J., Nettleton, D. and Sun, S. (2006)
Glucose-induced conformational changes in glucokinase
mediate allosteric regulation: transient kinetic analysis. Bio-
chemistry 45, 75537562.
14 Kim, Y. B., Kalinowski, S. S. and Marcinkeviciene, J. (2007) A
pre-steady state analysis of ligand binding to human gluco-
kinase: evidence for a preexisting equilibrium. Biochemistry
46, 14231431.
15 Kamata, K., Mitsuya, M., Nishimura, T., Eiki, J. and Nagata,
Y. (2004) Structural basis for allosteric regulation of the
monomeric allosteric enzyme human glucokinase. Structure
12, 429438.
16 Aleshin, A. E., Zeng, C., Bartunik, H. D., Fromm, H. J. and
Honzatko, R. B. (1998) Regulation of hexokinase I: crystal
structure of recombinant human brain hexokinase complexed
with glucose and phosphate. J. Mol. Biol. 282, 345357.
17 Zhang, J., Li, C., Chen, K., Zhu, W., Shen, X. and Jiang, H.
(2006) Conformational transition pathway in the allosteric
process of human glucokinase. Proc. Natl. Acad. Sci. USA
103, 1336813373.
18 Van Schaftingen, E., Detheux, M. and Veiga da Cunha, M.
(1994) Short-term control of glucokinase activity: role of a
regulatory protein. Faseb. J. 8, 414– 419.
19 Futamura, M., Hosaka, H., Kadotani, A., Shimazaki, H.,
Sasaki, K., Ohyama, S., Nishimura, T., Eiki, J. and Nagata,Y.
38 P. B. Iynedjian Molecular physiology of glucokinase
(2006) An allosteric activator of glucokinase impairs the
interaction of glucokinase and glucokinase regulatory protein
and regulates glucose metabolism. J. Biol. Chem. 281, 37668–
37674.
20 de la Iglesia, N., Veiga-da-Cunha, M., Van Schaftingen, E.,
Guinovart, J. J. and Ferrer, J. C. (1999) Glucokinase regula-
tory protein is essential for the proper subcellular localisation
of liver glucokinase. FEBS Lett. 456, 332338.
21 Shiota, C. , Coffey, J., Grimsby, J., Grippo, J. F. and Magnuson,
M. A. (1999) Nuclear import of hepatic glucokinase depends
upon glucokinase regulatory protein, whereas export is due to
a nuclear export signal sequence in glucokinase. J. Biol.
Chem. 274, 3712537130.
22 Bosco, D., Meda, P. and Iynedjian, P. B. (2000) Glucokinase
and glucokinase regulatory protein: mutual dependence for
nuclear localization. Biochem. J. 348 Pt 1, 215– 222.
23 Guigas, B., Bertrand, L., Taleux, N., Foretz, M., Wiernsperg-
er, N., Vertommen, D., Andreelli, F., Viollet, B. and Hue, L.
(2006) 5-Aminoimidazole-4-carboxamide-1-beta-D-ribofur-
anoside and metformin inhibit hepatic glucose phosphoryla-
tion by an AMP-activated protein kinase-independent effect
on glucokinase translocation. Diabetes 55, 865874.
24 Toyoda, Y., Miwa, I. , Kamiya, M., Ogiso, S., Nonogaki, T.,
Aoki, S. and Okuda, J. (1994) Evidence for glucokinase
translocation by glucose in rat hepatocytes. Biochem. Bio-
phys. Res. Commun. 204, 252 –256.
25 Brown, K. S., Kalinowski, S. S., Megill, J. R., Durham, S. K.
and Mookhtiar, K. A. (1997) Glucokinase regulatory protein
may interact with glucokinase in the hepatocyte nucleus.
Diabetes 46, 179186.
26 Toyoda, Y., Miwa, I. , Satake, S., Anai, M. and Oka, Y. (1995)
Nuclear location of the regulatory protein of glucokinase in
rat liver and translocation of the regulator to the cytoplasm in
response to high glucose. Biochem. Biophys. Res. Com-
mun. 215, 467473.
27 Mukhtar, M., Stubbs, M. and Agius, L. (1999) Evidence for
glucose and sorbitol-induced nuclear export of glucokinase
regulatory protein in hepatocytes. FEBS Lett. 462, 453 –458.
28 Farrelly, D., Brown, K. S., Tieman, A., Ren, J., Lira, S. A.,
Hagan, D., Gregg, R., Mookhtiar, K. A. and Hariharan, N.
(1999) Mice mutant for glucokinase regulatory protein exhibit
decreased liver glucokinase: a sequestration mechanism in
metabolic regulation. Proc. Natl. Acad. Sci. USA 96, 14511
14516.
29 Grimsby, J., Coffey, J. W. , Dvorozniak, M. T., Magram, J., Li,
G., Matschinsky,F. M. , Shiota, C., Kaur, S., Magnuson, M. A.
and Grippo, J. F. (2000) Characterization of glucokinase
regulatory protein-deficient mice. J. Biol. Chem. 275, 7826–
7831.
30 Arden, C., Baltrusch, S. and Agius, L. (2006) Glucokinase
regulatory protein is associated with mitochondria in hepa-
tocytes. FEBS Lett. 580, 2065– 2070.
31 Bustamante, E., Pediaditakis, P., He, L. and Lemasters, J. J.
(2005) Isolated mouse liver mitochondria are devoid of
glucokinase. Biochem. Biophys. Res. Commun. 334, 907 910.
32 Munoz-Alonso, M. J., Guillemain, G., Kassis, N., Girard, J.,
Burnol, A. F. and Leturque, A. (2000) A novel cytosolic dual
specificity phosphatase, interacting with glucokinase, increas-
es glucose phosphorylation rate. J. Biol. Chem. 275, 32406
32412.
33 Shiraishi, A., Yamada, Y. , Tsuura, Y., Fijimoto, S., Tsukiya-
ma, K., Mukai, E. , Toyoda, Y., Miwa, I. and Seino, Y. (2001) A
novel glucokinase regulator in pancreatic beta cells : precursor
of propionyl-CoA carboxylase beta subunit interacts with
glucokinase and augments its activity. J. Biol. Chem. 276,
23252328.
34 Baltrusch, S., Lenzen, S., Okar, D. A., Lange, A. J. and
Tiedge, M. (2001) Characterization of glucokinase-binding
protein epitopes by a phage-displayed peptide library. Iden-
tification of 6-phosphofructo-2-kinase/fructose-2,6-bisphos-
phatase as a novel interaction partner. J Biol Chem 276,
4391543923.
35 Massa, L. , Baltrusch, S., Okar, D. A., Lange, A. J., Lenzen, S.
and Tiedge, M. (2004) Interaction of 6-phosphofructo-2-
kinase/fructose-2,6-bisphosphatase (PFK-2/FBPase-2) with
glucokinase activates glucose phosphorylation and glucose
metabolism in insulin-producing cells. Diabetes 53, 1020–
1029.
36 Baltrusch, S., Langer, S. , Massa, L., Tiedge, M. and Lenzen, S.
(2006) Improved metabolic stimulus for glucose-induced
insulin secretion through GK and PFK-2/FBPase-2 coexpres-
sion in insulin-producing RINm5F cells. Endocrinology 147,
57685776.
37 Rizzo, M. A. and Piston, D. W. (2003) Regulation of beta cell
glucokinase by S-nitrosylation and association with nitric
oxide synthase. J. Cell. Biol. 161, 243248.
38 Arden, C. , Harbottle, A. , Baltrusch, S. , Tiedge, M. and Agius,
L. (2004) Glucokinase is an integral component of the insulin
granules in glucose-responsive insulin secretory cells and does
not translocate during glucose stimulation. Diabetes 53,
23462352.
39 Bjorkhaug, L., Molnes, J., Sovik, O., Njolstad, P. R. and
Flatmark, T. (2007) Allosteric activation of human glucoki-
nase by free polyubiquitin chains and its ubiquitin-dependent
cotranslational proteasomal degradation. J. Biol. Chem. 282,
2275722764.
40 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. and Korsmeyer, S. J.
(2003) BAD and glucokinase reside in a mitochondrial
complex that integrates glycolysis and apoptosis. Nature
424, 952956.
41 Danial, N. N., Walensky, L. D., Zhang, C.Y., Choi, C. S.,
Fisher, J.K., Molina, A.J., Datta, S.R., Pitter, K.L., Bird,
G. H., Wikstrom, J. D., Deeney, J. T., Robertson, K., Morash,
J., Kulkarni, A., Neschen, S., Kim, S., Greenberg, M. E.,
Corkey, B. E., Shirihai, O. S., Shulman, G. I., Lowell, B. B. and
Korsmeyer, S.J. (2008) Dual role of proapoptotic BAD in
insulin secretion and beta cell survival. Nat. Med. 14, 144– 153.
42 Jetton, T. L., Liang, Y., Pettepher, C. C., Zimmerman, E. C.,
Cox, F. G., Horvath, K., Matschinsky, F. M. and Magnuson,
M. A. (1994) Analysis of upstream glucokinase promoter
activity in transgenic mice and identification of glucokinase in
rare neuroendocrine cells in the brain and gut. J. Biol.
Chem. 269, 36413654.
43 Shelton, K. D., Franklin, A. J., Khoor, A., Beechem, J. and
Magnuson, M. A. (1992) Multiple elements in the upstream
glucokinase promoter contribute to transcription in insulino-
ma cells. Mol. Cell. Biol. 12, 4578– 4589.
44 Moates, J. M. and Magnuson, M. A. (2004) The Pal elements
in the upstream glucokinase promoter exhibit dyad symmetry
and display cell-specific enhancer activity when multimerised.
Diabetologia 47, 16321640.
45 Moates, J. M., Nanda, S., Cissell, M. A., Tsai, M. J. and Stein,
R. (2003) BETA2 activates transcription from the upstream
glucokinase gene promoter in islet beta-cells and gut endo-
crine cells. Diabetes 52, 403– 408.
46 Watada, H., Kajimoto, Y., Umayahara, Y., Matsuoka, T.,
Kaneto, H., Fujitani, Y., Kamada, T., Kawamori, R. and
Yamasaki, Y. (1996) The human glucokinase gene beta-cell-
type promoter: an essential role of insulin promoter factor 1/
PDX-1 in its activation in HIT-T15 cells. Diabetes 45, 1478
1488.
47 Chakrabarti, S. K., James, J. C. and Mirmira, R. G. (2002)
Quantitative assessment of gene targeting in vitro and in vivo
by the pancreatic transcription factor, Pdx1. Importance of
chromatin structure in directing promoter binding. J. Biol.
Chem. 277, 1328613293.
48 Iynedjian, P. B., Marie, S., Wang, H., Gjinovci, A. and
Nazaryan, K. (1996) Liver-specific enhancer of the glucoki-
nase gene. J. Biol. Chem. 271, 29113– 29120.
49 Iynedjian, P. B. (1998) Identification of upstream stimulatory
factor as transcriptional activator of the liver promoter of the
glucokinase gene. Biochem. J. 333 (Pt 3), 705712.
Cell. Mol. Life Sci. Vol.66, 2009 Review Article 39
50 Iynedjian, P. B., Gjinovci, A. and Renold, A. E. (1988)
Stimulation by insulin of glucokinase gene transcription in
liver of diabetic rats. J. Biol. Chem. 263, 740– 744.
51 Iynedjian, P. B., Marie, S., Gjinovci, A. , Genin, B., Deng, S. P. ,
Buhler, L., Morel, P. and Mentha, G. (1995) Glucokinase and
cytosolic phosphoenolpyruvate carboxykinase (GTP) in the
human liver. Regulation of gene expression in cultured
hepatocytes. J. Clin. Invest. 95, 19661973.
52 Iynedjian, P. B., Roth, R. A., Fleischmann, M. and Gjinovci,
A. (2000) Activation of protein kinase B/cAkt in hepatocytes
is sufficient for the induction of expression of the gene
encoding glucokinase. Biochem. J. 351 Pt 3, 621 –627.
53 Ribaux, P. G. and Iynedjian, P. B. (2003) Analysis of the role of
protein kinase B (cAKT) in insulin-dependent induction of
glucokinase and sterol regulatory element-binding protein 1
(SREBP1) mRNAs in hepatocytes. Biochem. J. 376, 697 –705.
54 Iynedjian, P. B. (2005) Lack of evidence for a role of TRB3/
NIPK as an inhibitor of PKB-mediated insulin signalling in
primary hepatocytes. Biochem. J. 386, 113– 118.
55 Wu, C., Okar, D. A., Stoeckman, A. K., Peng, L. J., Herrera,
A. H., Herrera, J. E., Towle, H. C. and Lange, A. J. (2004) A
potential role for fructose-2,6-bisphosphate in the stimulation
of hepatic glucokinase gene expression. Endocrinology 145,
650658.
56 Matsumoto, M., Ogawa, W., Teshigawara, K., Inoue, H.,
Miyake, K., Sakaue, H. and Kasuga, M. (2002) Role of the
insulin receptor substrate 1 and phosphatidylinositol 3-kinase
signaling pathway in insulin-induced expression of sterol
regulatory element binding protein 1c and glucokinase genes
in rat hepatocytes. Diabetes 51, 1672– 1680.
57 Foretz, M., Pacot, C., Dugail, I., Lemarchand, P., Guichard,
C., Le Liepvre, X., Berthelier-Lubrano, C., Spiegelman, B.,
Kim, J. B., Ferre, P. and Foufelle, F. (1999) ADD1/SREBP-1c
is required in the activation of hepatic lipogenic gene
expression by glucose. Mol. Cell. Biol. 19, 3760 –3768.
58 Fleischmann, M. and Iynedjian, P. B. (2000) Regulation of
sterol regulatory-element binding protein 1 gene expression
in liver: role of insulin and protein kinase B/cAkt. Biochem. J.
349, 1317.
59 Foretz, M., Guichard, C., Ferre, P. and Foufelle, F. (1999)
Sterol regulatory element binding protein-1c is a major
mediator of insulin action on the hepatic expression of
glucokinase and lipogenesis-related genes. Proc. Natl. Acad.
Sci. USA 96, 1273712742.
60 Kim, S. Y., Kim, H. I., Kim, T. H., Im, S. S., Park, S. K., Lee,
I. K., Kim, K. S. and Ahn, Y. H. (2004) SREBP-1c mediates
the insulin-dependent hepatic glucokinase expression. J. Biol.
Chem. 279, 3082330829.
61 Stoeckman, A. K. and Towle, H. C. (2002) The role of
SREBP-1c in nutritional regulation of lipogenic enzyme
gene expression. J. Biol. Chem. 277, 27029– 27035.
62 Liang, G., Yang, J., Horton, J. D., Hammer, R. E. , Goldstein,
J. L. and Brown, M. S. (2002) Diminished hepatic responseto
fasting/refeeding and liver X receptor agonists in mice with
selective deficiency of sterol regulatory element-binding
protein-1c. J. Biol. Chem. 277, 9520 9528.
63 Ono, H., Shimano, H., Katagiri, H., Yahagi, N., Sakoda, H.,
Onishi, Y. , Anai, M., Ogihara, T. , Fujishiro, M. , Viana, A. Y.,
Fukushima, Y., Abe, M., Shojima, N., Kikuchi, M., Yamada,
N., Oka, Y. and Asano, T. (2003) Hepatic Akt activation
induces marked hypoglycemia, hepatomegaly, and hyper-
triglyceridemia with sterol regulatory element binding pro-
tein involvement. Diabetes 52, 29052913.
64 Hansmannel, F., Mordier, S. and Iynedjian, P. B. (2006)
Insulin induction of glucokinase and fatty acid synthase in
hepatocytes: analysis of the roles of sterol-regulatory-ele-
ment-binding protein-1c and liver X receptor. Biochem. J.
399, 275283.
65 Roth, U., Jungermann, K. and Kietzmann, T. (2002) Activa-
tion of glucokinase gene expression by hepatic nuclear factor
4alpha in primary hepatocytes. Biochem. J. 365, 223– 228.
66 Roth, U., Curth, K., Unterman, T. G. and Kietzmann, T.
(2004) The transcription factors HIF-1 and HNF-4 and the
coactivator p300 are involved in insulin-regulated glucokinase
gene expression via the phosphatidylinositol 3-kinase/protein
kinase B pathway. J. Biol. Chem. 279, 2623– 2631.
67 Kietzmann, T., Roth, U., Freimann, S. and Jungermann, K.
(1997) Arterial oxygen partial pressures reduce the insulin-
dependent induction of the perivenously located glucokinase
in rat hepatocyte cultures: mimicry of arterial oxygen
pressures by H2O2. Biochem. J. 321 (Pt 1), 17– 20.
68 Jungermann, K. and Kietzmann, T. (2000) Oxygen: modu-
lator of metabolic zonation and disease of the liver. Hepatol-
ogy 31, 255260.
69 Emerling, B. M., Weinberg, F., Liu, J. L., Mak, T. W. and
Chandel, N. S. (2008) PTEN regulates p300-dependent hypo-
xia-inducible factor 1 transcriptional activity through Fork-
head transcription factor 3a (FOXO3a). Proc. Natl. Acad. Sci.
USA 105, 26222627.
70 Matsumoto, M., Pocai, A., Rossetti, L., Depinho, R. A. and
Accili, D. (2007) Impaired regulation of hepatic glucose
production in mice lacking the forkhead transcription factor
foxo1 in liver. Cell Metab. 6, 208– 216.
71 Iynedjian, P. B., Ucla, C. and Mach, B. (1987) Molecular
cloning of glucokinase cDNA. Developmental and dietary
regulation of glucokinase mRNA in rat liver. J. Biol.
Chem. 262, 60326038.
72 Iynedjian, P. B., Pilot, P. R., Nouspikel, T., Milburn, J. L.,
Quaade, C., Hughes, S., Ucla, C. and Newgard, C. B. (1989)
Differential expression and regulation of the glucokinase
gene in liver and islets of Langerhans. Proc. Natl. Acad. Sci.
USA 86, 78387842.
73 Da Silva Xavier, G., Qian, Q., Cullen, P. J. and Rutter, G. A.
(2004) Distinct roles for insulin and insulin-like growth factor-
1 receptors in pancreatic beta-cell glucose sensing revealed by
RNA silencing. Biochem. J. 377, 149158.
74 da Silva Xavier, G., Varadi, A., Ainscow, E. K. and Rutter,
G. A. (2000) Regulation of gene expression by glucose in
pancreatic beta -cells (MIN6) via insulin secretion and
activation of phosphatidylinositol 3-kinase. J. Biol.
Chem. 275, 3626936277.
75 Yoshida, K., Murao, K. , Imachi,H. , Cao, W. M. , Yu,X. , Li, J.,
Ahmed, R. A., Kitanaka, N., Wong, N. C., Unterman, T. G.,
Magnuson, M. A. and Ishida, T. (2007) Pancreatic glucoki-
nase is activated by insulin-like growth factor-I. Endocrinol-
ogy 148, 29042913.
76 Williams, T. F. , Exton, J. H., Park, C. R. and Regen, D. M.
(1968) Stereospecific transport of glucose in the perfused rat
liver. Am. J. Physiol. 215, 1200– 1209.
77 Seoane, J., Gomez-Foix, A. M. , ODoherty, R. M., Gomez-
Ara, C., Newgard, C. B. and Guinovart, J. J. (1996) Glucose 6-
phosphate produced by glucokinase, but not hexokinase I,
promotes the activation of hepatic glycogen synthase. J. Biol.
Chem. 271, 2375623760.
78 Agius, L., Peak, M. , Newgard, C. B., Gomez-Foix, A. M. and
Guinovart, J. J. (1996) Evidence for a role of glucose-induced
translocation of glucokinase in the control of hepatic glycogen
synthesis. J. Biol. Chem. 271, 3047930486.
79 Gomis, R. R., Ferrer, J. C. and Guinovart, J. J. (2000) Shared
control of hepatic glycogen synthesis by glycogen synthase
and glucokinase. Biochem. J. 351 Pt 3, 811816.
80 Takeuchi, H., Inoue, Y., Ishihara, H. and Oka, Y. (1996)
Overexpression of either liver type or pancreatic beta cell
type glucokinase via recombinant adenovirus enhances
glucose oxidation in isolated rat hepatocytes. FEBS
Lett. 393, 60 64.
81 Seoane, J., Barbera, A., Telemaque-Potts, S., Newgard, C. B.
and Guinovart, J. J. (1999) Glucokinase overexpression
restores glucose utilization and storage in cultured hepato-
cytes from male Zucker diabetic fatty rats. J. Biol. Chem. 274,
3183331838.
82 Niswender, K. D., Shiota, M., Postic, C., Cherrington, A. D.
and Magnuson, M. A. (1997) Effects of increased glucokinase
40 P. B. Iynedjian Molecular physiology of glucokinase
gene copy number on glucose homeostasis and hepatic
glucose metabolism. J. Biol. Chem. 272, 22570– 22575.
83 Ferre, T., Pujol, A., Riu, E., Bosch, F. and Valera, A. (1996)
Correction of diabetic alterations by glucokinase. Proc. Natl.
Acad. Sci. USA 93, 72257230.
84 Postic, C., Shiota, M., Niswender, K. D., Jetton, T. L., Chen,
Y., Moates, J. M., Shelton, K. D., Lindner, J., Cherrington,
A. D. and Magnuson, M. A. (1999) Dual roles for glucokinase
in glucose homeostasis as determined by liver and pancreatic
beta cell-specific gene knock-outs using Cre recombinase. J.
Biol. Chem. 274, 305 –315.
85 Rossetti, L. , Chen,W. , Hu, M., Hawkins, M., Barzilai, N. and
Efrat, S. (1997) Abnormal regulation of HGP by hyper-
glycemia in mice with a disrupted glucokinase allele. Am. J.
Physiol. 273, E743750.
86 Henquin, J. C., Ravier, M. A., Nenquin, M., Jonas, J. C. and
Gilon, P. (2003) Hierarchy of the beta-cell signals controlling
insulin secretion. Eur. J. Clin. Invest. 33, 742 750.
87 MacDonald, M. J., Fahien, L. A., Brown, L. J., Hasan, N. M.,
Buss, J. D. and Kendrick, M. A. (2005) Perspective: emerging
evidence for signaling roles of mitochondrial anaplerotic
products in insulin secretion. Am. J. Physiol. Endocrinol.
Metab. 288, E115.
88 Whitesell, R. R., Powers, A. C., Regen, D. M. and Abumrad,
N. A. (1991) Transport and metabolism of glucose in an
insulin-secreting cell line, beta TC-1. Biochemistry 30, 11560
11566.
89 Tal, M., Liang, Y., Najafi, H., Lodish, H. F. and Matschinsky,
F. M. (1992) Expression and function of GLUT-1 and GLUT-
2 glucose transporter isoforms in cells of cultured rat
pancreatic islets. J. Biol. Chem. 267, 17241– 17247.
90 Schuit, F., Moens, K., Heimberg, H. and Pipeleers, D. (1999)
Cellular origin of hexokinase in pancreatic islets. J. Biol.
Chem. 274, 3280332809.
91 Wang, H. and Iynedjian, P. B. (1997) Modulation of glucose
responsiveness of insulinoma beta-cells by graded overex-
pression of glucokinase. Proc. Natl. Acad. Sci. USA94, 4372
4377.
92 Wang, H. and Iynedjian, P. B. (1997) Acute glucose intoler-
ance in insulinoma cells with unbalanced overexpression of
glucokinase. J. Biol. Chem. 272, 25731 –25736.
93 Berman, H. K. and Newgard, C. B. (1998) Fundamental
metabolic differences between hepatocytes and islet beta-
cells revealed by glucokinase overexpression. Biochemistry
37, 45434552.
94 Terauchi, Y., Sakura, H. , Yasuda,K. , Iwamoto, K., Takahashi,
N., Ito, K., Kasai, H. , Suzuki, H., Ueda, O., Kamada, N. and et
al. (1995) Pancreatic beta-cell-specific targeted disruption of
glucokinase gene. Diabetes mellitus due to defective insulin
secretion to glucose. J. Biol. Chem. 270, 3025330256.
95 Yang, X. J., Kow, L. M. , Funabashi, T. and Mobbs, C. V.
(1999) Hypothalamic glucose sensor: similarities to and
differences from pancreatic beta-cell mechanisms. Diabetes
48, 17631772.
96 Roncero, I., Alvarez, E., Vazquez, P. and Blazquez, E. (2000)
Functional glucokinase isoforms are expressed in rat brain. J.
Neurochem. 74, 1848 –1857.
97 Dunn-Meynell, A. A., Routh, V. H. , Kang, L., Gaspers, L.
and Levin, B. E. (2002) Glucokinase is the likely mediator of
glucosensing in both glucose-excited and glucose-inhibited
central neurons. Diabetes 51, 2056– 2065.
98 Kang, L. , Routh, V. H. , Kuzhikandathil, E. V., Gaspers, L. D.
and Levin, B. E. (2004) Physiological and molecular charac-
teristics of rat hypothalamic ventromedial nucleus glucosens-
ing neurons. Diabetes 53, 549– 559.
99 Kang, L. , Dunn-Meynell, A. A. , Routh, V. H. , Gaspers, L. D.,
Nagata, Y., Nishimura, T., Eiki, J., Zhang, B. B. and Levin,
B. E. (2006) Glucokinase is a critical regulator of ventrome-
dial hypothalamic neuronal glucosensing. Diabetes 55, 412
420.
100 Parton, L. E., Ye, C. P., Coppari, R., Enriori, P. J., Choi, B.,
Zhang, C. Y., Xu, C., Vianna, C. R. , Balthasar, N., Lee, C. E. ,
Elmquist, J. K., Cowley, M. A. and Lowell, B. B. (2007)
Glucose sensing by POMC neurons regulates glucose homeo-
stasis and is impaired in obesity. Nature 449, 228– 232.
101 Wolfgang, M. J. and Lane, M. D. (2006) The role of hypo-
thalamic malonyl-CoA in energy homeostasis. J. Biol.
Chem. 281, 3726537269.
102 Velho, G., Blanche, H., Vaxillaire, M., Bellanne-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. and Froguel, P. (1997) Identification of 14 new
glucokinase mutations and description of the clinical profile
of 42 MODY-2 families. Diabetologia 40, 217224.
103 Fajans, S. S., Bell, G. I. and Polonsky, K. S. (2001) Molecular
mechanisms and clinical pathophysiology of maturity-onset
diabetes of the young. N. Engl. J. Med. 345, 971– 980.
104 Ellard, S., Beards, F., Allen, L. I., Shepherd, M. , Ballantyne,
E., Harvey, R. and Hattersley, A. T. (2000) A high prevalence
of glucokinase mutations in gestational diabetic subjects
selected by clinical criteria. Diabetologia 43, 250253.
105 Murphy, R., Ellard, S. and Hattersley, A. T. (2008) Clinical
implications of a molecular genetic classification of mono-
genic beta-cell diabetes. Nat Clin Pract Endocrinol. Metab. 4,
200213.
106 Njolstad, P. R., Sagen, J. V. , Bjorkhaug, L., Odili, S., Sheha-
deh, N., Bakry, D., Sarici, S. U., Alpay, F. , Molnes, J., Molven,
A., Sovik, O. and Matschinsky, F. M. (2003) Permanent
neonatal diabetes caused by glucokinase deficiency: inborn
error of the glucose-insulin signaling pathway. Diabetes 52,
28542860.
107 Velho, G., Petersen, K. F., Perseghin, G., Hwang, J. H. ,
Rothman, D. L., Pueyo, M. E., Cline, G. W., Froguel, P. and
Shulman, G. I. (1996) Impaired hepatic glycogen synthesis in
glucokinase-deficient (MODY-2) subjects. J. Clin. Invest. 98,
17551761.
108 Tappy, L., Dussoix, P., Iynedjian, P., Henry, S., Schneiter, P.,
Zahnd, G., Jequier, E. and Philippe, J. (1997) Abnormal
regulation of hepatic glucose output in maturity-onset dia-
betes of the young caused by a specific mutation of the
glucokinase gene. Diabetes 46, 204208.
109 Byrne, M. M., Sturis, J., Clement, K., Vionnet, N., Pueyo,
M. E., Stoffel, M., Takeda, J., Passa, P., Cohen, D. , Bell, G. I.
and et al. (1994) Insulin secretory abnormalities in subjects
with hyperglycemia due to glucokinase mutations. J. Clin.
Invest. 93, 1120 1130.
110 Miller, S. P., Anand, G. R. , Karschnia, E. J., Bell, G. I.,
LaPorte, D. C. and Lange, A. J. (1999) Characterization of
glucokinase mutations associated with maturity-onset diabe-
tes of the young type 2 (MODY-2): different glucokinase
defects lead to a common phenotype. Diabetes 48, 1645
1651.
111 Davis, E. A., Cuesta-Munoz, A., Raoul, M., Buettger, C.,
Sweet, I., Moates, M., Magnuson, M. A. and Matschinsky,
F. M. (1999) Mutants of glucokinase cause hypoglycaemia-
and hyperglycaemia syndromes and their analysis illuminates
fundamental quantitative concepts of glucose homeostasis.
Diabetologia 42, 11751186.
112 Galan, M. , Vincent, O., Roncero, I., Azriel, S., Boix-Pallares,
P., Delgado-Alvarez, E. , Diaz-Cadorniga, F., Blazquez, E.
and Navas, M. A. (2006) Effects of novel maturity-onset
diabetes of the young (MODY)-associated mutations on
glucokinase activity and protein stability. Biochem. J. 393,
389396.
113 Burke, C. V., Buettger, C. W., Davis, E. A. , McClane, S. J.,
Matschinsky, F. M. and Raper, S. E. (1999) Cell-biological
assessment of human glucokinase mutants causing maturity-
onset diabetes of the young type 2 (MODY-2) or glucokinase-
linked hyperinsulinaemia (GK-HI). Biochem. J. 342 (Pt 2),
345352.
114 Toye, A. A., Moir, L. , Hugill, A., Bentley,L., Quarterman, J.,
Mijat, V., Hough, T., Goldsworthy, M., Haynes, A., Hunter,
A. J., Browne, M., Spurr, N. and Cox, R. D. (2004) A new
mouse model of type 2 diabetes, produced by N-ethyl-
Cell. Mol. Life Sci. Vol.66, 2009 Review Article 41
nitrosourea mutagenesis, is the result of a missense mutation
in the glucokinase gene. Diabetes 53, 15771583.
115 Aigner, B., Rathkolb, B., Herbach, N., Hrabe de Angelis, M.,
Wanke, R. and Wolf, E. (2008) Diabetes models by screen for
hyperglycemia in phenotype-driven ENU mouse mutagenesis
projects. Am. J. Physiol. Endocrinol. Metab. 294, E232– 240.
116 Inoue, M., Sakuraba, Y., Motegi, H., Kubota, N., Toki, H.,
Matsui, J., Toyoda, Y., Miwa, I., Terauchi, Y., Kadowaki, T.,
Shigeyama, Y., Kasuga, M., Adachi, T., Fujimoto, N.,
Matsumoto, R., Tsuchihashi, K., Kagami, T., Inoue, A.,
Kaneda, H., Ishijima, J., Masuya, H. , Suzuki, T., Wakana, S.,
Gondo, Y., Minowa, O., Shiroishi, T. and Noda, T. (2004) A
series of maturity onset diabetes of the young, type 2
(MODY2) mouse models generated by a large-scale ENU
mutagenesis program. Hum. Mol. Genet. 13, 1147 –1157.
117 Glaser, B., Kesavan, P., Heyman, M., Davis, E., Cuesta, A.,
Buchs, A., Stanley, C. A. , Thornton, P. S., Permutt, M. A. ,
Matschinsky, F. M. and Herold, K. C. (1998) Familial hyper-
insulinism caused by an activating glucokinase mutation. N.
Engl. J. Med. 338, 226230.
118 Gloyn, A. L., Noordam, K., Willemsen, M. A. , Ellard, S.,
Lam, W. W., Campbell, I. W., Midgley, P., Shiota, C., Buettg-
er, C., Magnuson, M. A., Matschinsky, F. M. and Hattersley,
A. T. (2003) Insights into the biochemical and genetic basis of
glucokinase activation from naturally occurring hypoglyce-
mia mutations. Diabetes 52, 2433– 2440.
119 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.
and Laakso, M. (2004) Severe persistent hyperinsulinemic
hypoglycemia due to a de novo glucokinase mutation.
Diabetes 53, 21642168.
120 Wabitsch, M., Lahr, G., Van de Bunt, M., Marchant, C.,
Lindner, M., von Puttkamer, J., Fenneberg, A., Debatin,
K. M. , Klein, R., Ellard, S., Clark, A. and Gloyn, A. L. (2007)
Heterogeneity in disease severity in a family with a novel
G68V GCK activating mutation causing persistent hyper-
insulinaemic hypoglycaemia of infancy. Diabet. Med. 24,
13931399.
121 Pino, M. F., Kim, K. A., Shelton, K. D., Lindner, J., Odili, S.,
Li, C., Collins, H. W., Shiota, M., Matschinsky, F. M. and
Magnuson, M. A. (2007) Glucokinase thermolability and
hepatic regulatory protein binding are essential factors for
predicting the blood glucose phenotype of missense muta-
tions. J. Biol. Chem. 282, 1390613916.
122 Muoio, D. M. and Newgard, C. B. (2008) Mechanisms of
disease: molecular and metabolic mechanisms of insulin
resistance and beta-cell failure in type 2 diabetes. Nat. Rev.
Mol. Cell. Biol. 9, 193 –205.
123 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. and Grippo, J. F. (2003) Allosteric
activators of glucokinase: potential role in diabetes therapy.
Science 301, 370373.
124 Brocklehurst, K. J., Payne, V. A., Davies, R. A., Carroll, D.,
Vertigan, H. L., Wightman, H. J., Aiston, S., Waddell, I. D.,
Leighton, B., Coghlan, M. P. and Agius, L. (2004) Stimulation
of hepatocyte glucose metabolism by novel small molecule
glucokinase activators. Diabetes 53, 535– 541.
125 Efanov, A. M. , Barrett, D. G., Brenner, M. B., Briggs, S. L.,
Delaunois, A., Durbin, J. D. , Giese, U., Guo, H., Radloff, M.,
Gil, G. S. , Sewing, S., Wang, Y., Weichert, A., Zaliani, A. and
Gromada, J. (2005) A novel glucokinase activator modulates
pancreatic islet and hepatocyte function. Endocrinology 146,
36963701.
126 Fyfe, M. C., White, J. R., Taylor, A., Chatfield, R., Wargent,
E., Printz, R. L., Sulpice, T., McCormack, J. G., Procter,
M. J., Reynet, C., Widdowson, P. S. and Wong-Kai-In, P.
(2007) Glucokinase activator PSN-GK1 displays enhanced
antihyperglycaemic and insulinotropic actions. Diabetologia
50, 12771287.
127 Coghlan, M. and Leighton, B. (2008) Glucokinase activators
in diabetes management. Expert Opin. Investig. Drugs 17,
145167.
128 ODoherty, R. M., Lehman, D. L., Telemaque-Potts, S. and
Newgard, C. B. (1999) Metabolic impact of glucokinase
overexpression in liver: lowering of blood glucose in fed rats
is accompanied by hyperlipidemia. Diabetes 48, 2022 2027.
129 Ferre, T., Riu, E., Franckhauser, S., Agudo, J. and Bosch, F.
(2003) Long-term overexpression of glucokinase in the liver
of transgenic mice leads to insulin resistance. Diabetologia 46,
16621668.
130 Aston-Mourney, K., Proietto, J., Morahan, G. and Andriko-
poulos, S. (2008) Too much of a good thing: why it is bad to
stimulate the beta cell to secrete insulin. Diabetologia 51,
540545.
To access this journal online:
http://www.birkhauser.ch/CMLS
42 P. B. Iynedjian Molecular physiology of glucokinase
... In beta cells, G6PC2 opposes the action of glucokinase, creating a futile substrate cycle that determines the rate of glycolytic flux and, in turn, the sensitivity of glucose-stimulated insulin secretion (GSIS) to glucose [6][7][8][9]. This model is supported by experiments showing reduced G6Pase activity [7] and glucose cycling [10] and elevated glycolysis in isolated G6pc2 knockout (KO) relative to wild type (WT) islets [8]. ...
Article
Full-text available
G6PC2 encodes a glucose-6-phosphatase (G6Pase) catalytic subunit, primarily expressed in pancreatic islet beta cells, that modulates the sensitivity of insulin secretion to glucose and thereby regulates fasting blood glucose (FBG). Mutational analyses were conducted to validate an AlphaFold2 (AF2)-predicted structure of human G6PC2 in conjunction with a novel method to solubilize and purify human G6PC2 from a heterologous expression system. These analyses show that residues forming a predicted intramolecular disulfide bond are essential for G6PC2 expression and that residues forming part of a type 2 phosphatidic acid phosphatase (PAP2) motif are critical for enzyme activity. Additional mutagenesis shows that residues forming a predicted substrate cavity modulate enzyme activity and substrate specificity and residues forming a putative cholesterol recognition amino acid consensus (CRAC) motif influence protein expression or enzyme activity. This CRAC motif begins at residue 219, the site of a common G6PC2 non-synonymous single nucleotide polymorphism (SNP), rs492594 (Val219Leu), though the functional impact of this SNP is disputed. In microsomal membrane preparations, the L219 variant has greater activity than the V219 variant, but this difference disappears when G6PC2 is purified in detergent micelles. We hypothesize that this was due to a differential association of the two variants with cholesterol. This concept was supported by the observation that the addition of cholesteryl hemi-succinate to the purified enzymes decreased the Vmax of the V219 and L219 variants ~8-fold and ~3 fold, respectively. We anticipate that these observations should support the rational development of G6PC2 inhibitors designed to lower FBG.
... This decline preceded the current resurgence of interest in metabolic regulation, and ironically, the discovery of new molecular mechanisms that could contribute to differential enzyme regulation according to cell location. For example, we now know that the nuclear accumulation of a metabolic enzyme can depend on its binding to a protein inhibitor of catalytic activity (nuclear glucokinase inhibition by glucokinase regulatory protein) [6]. Furthermore, it is now clear that the catalytic activity of some nucleocytosolic enzymes is controlled by post-translational modification. ...
Article
Full-text available
Some metabolic enzymes normally occur in the nucleus and cytoplasm. These compartments differ in molecular composition. Since post-translational modification and interaction with allosteric effectors can tune enzyme activity, it follows that the behavior of an enzyme as a catalyst may differ between the cytoplasm and nucleus. We explored this possibility for the glycolytic enzyme glyceraldehyde 3-phosphate dehydrogenase (GAPDH). Homogenates of pristine nuclei and cytoplasms isolated from Xenopus laevis oocytes were used for whole compartment activity profiling in a near-physiological buffer. Titrations of NAD+ revealed similar whole compartment activity profiles for GAPDH in nuclear and cytoplasmic homogenates. Surprisingly however GAPDH in these compartments did not have the same behavior in assays of the dependence of initial velocity (v0) on G3P concentration. First, the peak v0 for nuclear GAPDH was up to 2.5-fold higher than the peak for cytoplasmic GAPDH. Second, while Michaelis Menten-like behavior was observed in all assays of cytoplasm, the v0 versus [G3P] plots for nuclear GAPDH typically exhibited a non-Michaelis Menten (sigmoidal) profile. Apparent Km and Vmax (G3P) values for nuclear GAPDH activity were highly variable, even between replicates of the same sample. Possible sources of this variability include in vitro processing of a metabolite that allosterically regulates GAPDH, turnover of a post-translational modification of the enzyme, and fluctuation of the state of interaction of GAPDH with other proteins. Collectively these findings are consistent with the hypothesis that the environment of the nucleus is distinct from the environment of the cytoplasm with regard to GAPDH activity and its modulation. This finding warrants further comparison of the regulation of nuclear and cytoplasmic GAPDH, as well as whole compartment activity profiling of other enzymes of metabolism with cytosolic and nuclear pools.
Article
As a sensor, glucokinase (GK) controls glucose homeostasis, which progressively declines in patients with diabetes. GK maintains the equilibrium of glucose levels and regulates the homeostatic system set points. Endocrine and hepatic cells can both respond to glucose cooperatively when GK is activated. GK has been under study as a therapeutic target for decades due to the possibility that cellular GK expression and function can be recovered, hence restoring glucose homeostasis in patients with type 2 diabetes. Five therapeutic compounds targeting GK are being investigated globally at the moment. They all have distinctive molecular structures and have been clinically shown to have strong antihyperglycemia effects. The mechanics, classification, and clinical development of GK activators are illustrated in this review. With the recent approval and marketing of the first GK activator (GKA), dorzagliatin, GKA's critical role in treating glucose homeostasis disorder and its long‐term benefits in diabetes will eventually become clear. image
Article
Full-text available
For glucose-stimulated-insulin-secretion (GSIS) by pancreatic β-cells in animals, it is believed that ATP generated from glucose metabolism is primarily responsible for GSIS. However, this ignores two well established aspects in literature: (a) intracellular ATP generation from other sources resulting in an overall pool of ATP, regardless of the original source, and (b) that intracellular glucose transport is 10 to 100-fold higher than intracellular glucose phosphorylation in β-cells. The latter especially provides an earlier unaddressed, but highly appealing, observation pertaining to (at least transient) presence of intracellular glucose molecules. Could these intracellular glucose molecules be responsible for specificity of GSIS to glucose (instead of the widely believed ATP production from its metabolism). In this work, we provide a comprehensive compilation of literature on glucose and GSIS using various cellular systems - all studies focus only on extracellular role of glucose in GSIS. Further, we carried out a comprehensive analysis of differential gene expression in Mouse Insulinoma 6 (Min6) cells, exposed to low and high extracellular-glucose concentrations (EGC), from the existing whole transcriptome data. Expression of other genes involved in Glycolysis, Krebs cycle, and electron transport chain were found to be unaffected by EGC except Gapdh, Atp6v0a4 & Cox20. Remarkably 3 upregulated genes (Atp6v0a4, Cacnb4, Kif11) in high EGC were identified to have an association with cellular secretion. Using glucose as a possible ligand for the 3 proteins, computational investigations were carried out (that will require future 'wet validation', both in vitro and in vivo, e.g., using primary islets and animal models). The glucose-affinity/binding scores (in kcal/mol) obtained were also compared with glucose binding scores for positive controls (GCK and GLUT2), along with negative controls (RPA1, KU70-80, POLA1, ACAA1A, POLR1A). Binding affinity scores of glucose molecules for the 3 proteins were found to be closer to positive controls. Therefore, we report glucose binding ability of 3 secretion-related proteins, and a possible direct role of intracellular glucose molecules in GSIS.
Preprint
Full-text available
Objective Blood glucose metabolism is a crucial risk factor for the onset of osteoarthritis (OA). Our aim is to assess the causal relationship between blood glucose levels and the incidence of osteoarthritis, as well as the impact of antidiabetic medications on osteoarthritis outcomes, using Mendelian randomization (MR) methods. Methods This is a two-sample Mendelian randomization design. Osteoarthritis was chosen as the outcome, and causal relationships with three exposures—glycated hemoglobin (HbA1c), fasting blood glucose (FG), and postprandial two-hour blood glucose (2hGlu)—were evaluated. Additionally, this study examined the relationship between common antidiabetic medication target genes (GCK, HFE, HK1, HKDC1, TCF7L2, and INS) and osteoarthritis. The inverse variance-weighted method (IVW) is primarily employed for effect estimation. Results HbA1c levels were negatively correlated with hip OA (OR 0.725, 95% CI 0.558–0.944) but not associated with knee OA or OA at any site. FG levels were negatively correlated with hip OA (OR 0.675, 95% CI 0.547–0.834) and OA at any site (OR 0.889, 95% CI 0.801–0.986). Gene predictions revealed a negative correlation between GCK and hip OA (OR 0.359, 95% CI 0.238–0.542), HFE and hip OA(OR 0.332, 95% CI 0.181–0.607), HK1 and knee OA (OR 0.735, 95% CI 0.565–0.955) and OA at any site (OR 0.743, 95% CI 0.636–0.868), HKDC1 and knee OA(OR 0.731, 95% CI 0.555–0.963) and OA at any site (OR 0.771, 95% CI 0.655–0.908), while INS shows a positive correlation with knee OA (OR 17.277, 95% CI 4.408–67.722) and OA at any site (OR 4.168, 95% CI 1.844–9.422). Conclusion MR analysis revealed a correlation between reduced HbA1c levels and an increased risk of hip OA, and FG was associated with an increased risk of hip OA and OA at any site. Gene predictions suggest that activating GCK, HFE, HK1, and HKDC1 genes with antidiabetic medications may reduce the risk of developing osteoarthritis. Consideration of antidiabetic medications for treating osteoarthritis is plausible, but clinical validation is necessary.
Chapter
Full-text available
Recent studies have emphasized the multiple aspects of non-coding micro-RNAs in the regulation of pancreatic ß cells in type 2 diabetic patients. Thus, highlighting the significance of non-coding regions of the genome in regulating pancreatic endocrine cells. Functional dysregulation of pancreatic endocrine cells increases the incidence of metabolic disorders in otherwise healthy individuals. A precise understanding of the molecular biology of metabolic dysregulation is important from cellular and clinical perspectives. The current chapter will highlight the important recent findings from type 2 diabetic human patients and aims to enhance our current understanding of ß cell pathophysiology from a clinical perspective for the development of novel therapeutic approaches to control this global incidence.
Article
Many synthetic glucokinase activators (GKAs), modulating glucokinase (GK), an important therapeutic target in diabetes have failed to clear clinical trials. In this study, an in silico structural similarity search with differing scaffolds of reference GKAs have been used to identify derivatives from natural product databases. Ten molecules with good binding score and similar interactions to that in the co‐crystallized GK as well good activation against recombinant human GK experimentally were identified. Tetrahydropalmatine, an alkaloid present in formulations and drugs from medicinal plants, has not been explored as an antidiabetic agent and no information regarding its mechanism of action or GK activation exists. Tetrahydropalmatine activates GK with EC 50 value of 71.7 ± 17.9 μM while lowering the S 0.5 (7.1 mM) and increasing V max (9.22 μM/min) as compared to control without activator ( S 0.5 = 10.37 mM; V max = 4.8 μM/min). Kinetic data ( α and β values) suggests it to act as mixed, nonessential type activator. Using microscale thermophoresis, K d values of 3.8 μM suggests a good affinity for GK. In HepG2 cell line, the compound potentiated the uptake of glucose and maintained glucose homeostasis by increasing the expression of GK, glycogen synthase, and insulin receptor genes and lowering the expression of glucokinase regulatory protein (GKRP) and glucagon. Tetrahydropalmatine at low concentrations could elicit a good response by reducing expression of GKRP, increasing expression of GK while also activating it. Thus, it could be used alone or in combination as therapeutic drug as it could effectively modulate GK and alter glucose homeostasis.
Article
Full-text available
Mexico Mitochondria are the powerhouse of the cell and dynamically control fundamental biological processes including cell reprogramming, pluripotency, and lineage specification. Although remarkable progress in induced pluripotent stem cell (iPSC)-derived cell therapies has been made, very little is known about the role of mitochondria and the mechanisms involved in somatic cell reprogramming into iPSC and directed reprogramming of iPSCs in terminally differentiated cells. Reprogramming requires changes in cellular characteristics, genomic and epigenetic regulation, as well as major mitochondrial metabolic changes to sustain iPSC self-renewal, pluripotency, and proliferation. Differentiation of autologous iPSC into terminally differentiated β-like cells requires further metabolic adaptation. Many studies have characterized these alterations in signaling pathways required for the generation and differentiation of iPSC; however, very little is known regarding the metabolic shifts that govern pluripotency transition to tissue-specific lineage differentiation. Understanding such metabolic transitions and how to modulate them is essential for the optimization of differentiation processes to ensure safe iPSC-derived cell therapies. In this review, we summarize the current understanding of mitochondrial metabolism during somatic cell reprogramming to iPSCs and the metabolic shift that occurs during directed differentiation into pancreatic β-like cells.
Article
Full-text available
The importance of the insulin receptor (IR) and the insulin-like growth factor-1 receptor (IGF-lR) for glucose-regulated insulin secretion and gene expression in pancreatic islet,beta-cells is at present unresolved. Here, we have used small interfering RNAs (siRNAs) to silence the expression of each receptor selectively in clonal MIN6 beta-cells. Reduction of IR levels by > 90% completely inhibited glucose (30 mM compared with 3 mM)-induced insulin secretion, but had no effect on depolarization-stimulated secretion. IR depletion also blocked the accumulation of preproinsulin (PPI), pancreatic duodenum homoeobox-1 (PDX-1) and glucokinase (GK) mRNAs at elevated glucose concentrations, as assessed by quantitative real-time PCR analysis (TaqMan((R))). Similarly, depletion of IGF-1R inhibited glucose-induced insulin secretion but, in contrast with the effects of IR silencing, had little impact on the regulation of gene expression by glucose. More-over, loss of IGF-i=lR, but not IR, markedly inhibited glucose-stimulated increases in cytosolic and mitochondrial ATP, suggesting a role for IGF-1R in the maintenance of oxidative metabolism and in the generation of mitochondrial coupling factors. RNA silencing thus represents a useful tool for the efficient and selective inactivation of receptor tyrosine kinases in isolated,B-cells. By inhibiting glucose-stimulated insulin secretion through the inactivation of IGF-1R, this approach also demonstrates the existence of insulin-independent mechanisms whereby elevated glucose concentrations regulate PPI, PDX-1 and GK gene expression in beta-cells.
Article
Full-text available
The low affinity glucose-phosphorylating enzyme glucokinase shows the phenomenon of intracellular translocation in beta cells of the pancreas and the liver. To identify potential binding partners of glucokinase by a systematic strategy, human beta cell glucokinase was screened by a 12-mer random peptide library displayed by the M13 phage. This panning procedure revealed two consensus motifs with a high binding affinity for glucokinase. The first consensus motif, LSAXXVAG, corresponded to the glucokinase regulatory protein of the liver. The second consensus motif, SLKVWT, showed a complete homology to the bifunctional enzyme 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase (PFK-2/FBPase-2), which acts as a key regulator of glucose metabolism. Through yeast two-hybrid analysis it became evident that the binding of glucokinase to PFK-2/FBPase-2 is conferred by the bisphosphatase domain, whereas the kinase domain is responsible for dimerization. 5′-Rapid amplification of cDNA ends analysis and Northern blot analysis revealed that rat pancreatic islets express the brain isoform of PFK-2/FBPase-2. A minor portion of the islet PFK-2/FBPase-2 cDNA clones comprised a novel splice variant with 8 additional amino acids in the kinase domain. The binding of the islet/brain PFK-2/ FBPase-2 isoform to glucokinase was comparable with that of the liver isoform. The interaction between glucokinase and PFK-2/FBPase-2 may provide the rationale for recent observations of a fructose-2,6-bisphosphate level-dependent partial channeling of glycolytic intermediates between glucokinase and glycolytic enzymes. In pancreatic beta cells this interaction may have a regulatory function for the metabolic stimulus-secretion coupling. Changes in fructose-2,6-bisphosphate levels and modulation of PFK-2/FBPase-2 activities may participate in the physiological regulation of glucokinase-mediated glucose-induced insulin secretion.
Article
Full-text available
Neonatal diabetes can be either permanent or transient. We have recently shown that permanent neonatal diabetes can result from complete deficiency of glucokinase activity. Here we report three new cases of glucokinase-related permanent neonatal diabetes. The probands had intrauterine growth retardation (birth weight <1,900 g) and insulin-treated diabetes from birth (diagnosis within the first week of life). One of the subjects was homozygous for the missense mutation Ala378Val (A378V), which is an inactivating mutation with an activity index of only 0.2% of wild-type glucokinase activity. The second subject was homozygous for a mutation in the splice donor site of exon 8 (intervening sequence 8 [IVS8] + 2T-->G), which is predicted to lead to the synthesis of an inactive protein. The third subject (second cousin of subject 2) was a compound heterozygote with one allele having the splice-site mutation IVS8 + 2T-->G and the other the missense mutation Gly264Ser (G264S), a mutation with an activity index of 86% of normal activity. The five subjects with permanent neonatal diabetes due to glucokinase deficiency identified to date are characterized by intrauterine growth retardation, permanent insulin-requiring diabetes from the first day of life, and hyperglycemia in both parents. Autosomal recessive inheritance and enzyme deficiency are features typical for an inborn error of metabolism, which occurred in the glucose-insulin signaling pathway in these subjects.
Article
Full-text available
Inhibition studies of glucokinase were carried out with the products of the reaction, glucose 6-phosphate and MgADP-, as well as with ADP3-, Mg2+ and ATP4-. The results of these, together with those of kinetic studies of the uninhibited reaction described previously [Storer & Cornish-Bowden (1976) Biochem. J. 159, 7-14], indicate that the enzyme obeys a 'mnemonical' mechanism. This implies that the co-operativity observed with glucose as substrate arises because glucose binds differentially to two forms of the free enzyme that are not in equilibrium under steady-state conditions. The mechanism predicts the decrease in glucose co-operativity observed at low concentrations of MgATP2-. The product-inhibition results suggest that glucose 6-phosphate is released first and that it is possibly displaced by MgATP2- in a concerted reaction.
Article
Nature 356, 162-164 (1992) THE following was omitted from the Acknowledgements section of this Letter: "This work was supported in part by a NIH grant to M.A.P. Y.T. was a recipient of a mentor based fellowship award of the American Diabetes Association."
Article
Kinetic characteristics of glucose transport and glucose phosphorylation were studied in the islet cell line beta-TC-1 to explore the roles of these processes in determining the dependence of glucose metabolism and insulin secretion on external glucose. The predominant glucose transporter present was the rat brain/erythrocyte type (Glut1), as determined by RNA and immunoblot analysis. The liver/islet glucose transporter (Glut2) RNA was not detected. The functional parameters of zero-trans glucose entry were K(m) = 9.5 +/- 2 mM and V(max) = 15.2 +/- 2 nmol min-1 (mu-L of cell water)-1. Phosphorylation kinetics of two hexokinase activities were characterized in situ. A low-K(m) (0.036 mM) hexokinase with a V(max) of 0.40 nmol min-1 (mu-L of cell water)-1 was present along with a high-K(m) (10 mM) hexokinase, which appeared to conform to a cooperative model with a Hill coefficient of about 1.4 and a V(max) of 0.3 nmol min-1 (mu-L of cell water)-1. Intracellular glucose at steady state was about 80% of the extracellular glucose from 3 to 15 mM, and transport did not limit metabolism in this range. In this static (nonperifusion) system, 2-3 times more immunoreactive insulin was secreted into the medium at 15 mM glucose than at 3 mM. The dependence of insulin secretion on external glucose roughly paralleled the dependence of glucose metabolism on external glucose. Simulations with a model demonstrated the degree to which changes in transport activity would affect intracellular glucose levels and the rate of the high-K(m) hexokinase (with the potential to affect insulin release).
Article
Recently, the description of glucokinase mRNA in certain neuroendocrine cells has opened new ways to characterize this enzyme in the rat brain. In this study, we found glucokinase mRNA and a similar RNA splicing pattern of the glucokinase gene product in rat hypothalamus and pancreatic islets; the mRNA that codes for B1 isoform was the most abundant, with minor amounts of those coding for the B2, P1, P2, P1/B2, and P2/B2 isoforms. Glucokinase gene expression in rat brain gave rise to a protein of 52 kDa with a high apparent Km for glucose and no product inhibition by glucose 6-phosphate, with a contribution to the total glucose phosphorylating activity of between 40 and 14%; the hypothalamus and cerebral cortex were the regions of maximal activity. Low and high Km hexokinases were characterized by several criteria. Also, using RT-PCR analysis we found a glucokinase regulatory protein mRNA similar to that previously reported in liver. These findings indicate that the glucokinase present in rat brain should facilitate the adaptation of this organ to fluctuations in blood glucose concentrations, and the expression of glucokinase and GLUT-2 in the same hypothalamic neurons suggests a role in glucose sensing.
Article
Adenovirus-mediated overexpression of the glucose phosphorylating enzyme glucokinase causes large changes in glycolytic flux and glucose storage in isolated rat hepatocytes, but not in pancreatic islets. We have used the well-differentiated insulinoma cell line INS-1 to investigate the basis for these apparent cell-type specific differences. We find that 2- or 5-[3H]glucose usage is increased at low (</=5 mM) but not high glucose concentrations in INS-1 cells treated with a recombinant adenovirus containing the glucokinase cDNA (AdCMV-GKI), while glucose usage is increased at both low and high glucose concentrations in similarly treated hepatocytes. Utilization of 2-[3H]glucose in INS-1 cells is suppressed in glucokinase overexpressing INS-1 cells in a rapid, glucose concentration-dependent, and reversible fashion, while such regulation is largely absent in hepatocytes. Levels of hexose phosphates (glucose-6-phosphate, fructose-6-phosphate, and fructose-1,6-bisphosphate) were profoundly and rapidly elevated following the switch to high glucose in either AdCMV-GKI-treated INS-1 cells or hepatocytes relative to controls. In contrast, triose phosphate levels (glyceraldehyde-3-phosphate + dihydroxyacetone phosphate) were much higher in AdCMV-GKI-treated INS-1 cells than in similarly treated hepatocytes, suggesting limited flux throught the glyceraldehyde-3-phosphate dehydrogenase (G3PDH) step in the former cells. Hepatocytes were found to contain approximately 62 times more lactate dehydrogenase (LDH) activity than INS-1 cells, and this was reflected in a 3-fold increase in lactate production in AdCMV-GKI-treated hepatocytes relative to similarly treated INS-1 cells. Since the amounts of G3PDH activity in INS-1 and hepatocyte extracts are similar, we suggest that flux through this step in INS-1 cells is limited by failure to regenerate NAD in the LDH reaction and that a fundamental difference between hepatocytes and islet beta-cells is the limited capacity of the latter to metabolize glycolytic intermediates beyond the G3PDH step.
Article
Glucokinase is a monomeric protein under native and denaturating conditions yet presents a sigmoidal saturation function for glucose. These peculiarities suggested the possibility that polymerization occurs under assay conditions. Thus the apparent molecular weight of glucokinase was determined by gel filtration at 4 °C and at 30 °C in the presence of substrates and products, singly and in combination, creating during the filtration similar conditions as used in the assay. Gel filtration was performed also in the presence of N-acetylglucosamine, which is a competitive inhibitor and shifts to an hyperbole the saturation function for glucose. The same elution behavior, that is, a single symmetrical peak, was observed in every system used. This persistent monomeric form of glucokinase excludes the possibility that the sigmoidal function is the result of the interaction of different subunits. The possibility of an association-dissociation equilibrium in which the kinetic properties of the enzyme depend on the particular molecular weight species may also be rejected.
Article
Maturity-onset diabetes of the young (MODY) is a subtype of type 2 diabetes that presents from the second decade and has an autosomal dominant mode of inheritance. We have investigated the glucokinase gene, a candidate gene for diabetes, in two MODY pedigrees. In a large 5-generation pedigree (BX) with 15 diabetic members, use of a microsatellite polymorphism revealed linkage of diabetes to the glucokinase locus on chromosome 7p. A peak lod score of 4.60 was obtained at a recombination fraction (theta) of zero. This finding suggests that a defective glucokinase gene contributes to the diabetes phenotype in this pedigree. This is not universal in MODY since linkage to the glucokinase locus was excluded in a second pedigree M (lod score = -7.36 at theta = 0). The affected members in pedigree BX were diagnosed either when young (in pregnancy or on screening) or when they presented symptomatically in middle and old age; most of them were treated by diet alone. Defects in the glucokinase gene may play an important part in the pathogenesis of type 2 diabetes.