Anda di halaman 1dari 12

Review

doi: 10.1111/joim.12096

The required beta cell research for improving treatment of type 2 diabetes
B. Thorens
From the Center for Integrative Genomics, University of Lausanne, Lausanne, Switzerland

Abstract. Thorens B (University of Lausanne, Lausanne, Switzerland). The required beta cell research for improving treatment of type 2 diabetes. (Review). J Intern Med 2013; 274: 203 214. In healthy individuals, insulin resistance is associated with physiological conditions such as pregnancy or body weight gain and triggers an increase in beta cell number and insulin secretion capacity to preserve normoglycaemia. Failure of this beta cell compensation capacity is a fundamental cause

of diabetic hyperglycaemia. Incomplete understanding of the molecular mechanisms controlling the plasticity of adult beta cells mechanisms and how these cells fail during the pathogenesis of diabetes strongly limits the ability to develop new beta cell-specic therapies. Here, current knowledge of the signalling pathways controlling beta cell plasticity is reviewed, and possible directions for future research are discussed. Keywords: apoptosis, beta cells, diabetes, GLP-1, insulin secretion, pregnancy.

Introduction Insufcient insulin secretion by pancreatic beta cells to compensate for developing insulin resistance of liver, muscles and adipose tissue is considered to be the cause of overt type 2 diabetes [1]. In insulin-resistant conditions, such as during pregnancy or in response to increased body weight, there is an increase in both beta cell number and glucose competence (i.e. the amount of insulin the cells can secrete in response to a given rise in extracellular glucose concentration). This beta cell plasticity ensures that insulin secretion can precisely match the metabolic requirements of the organism under changing environmental conditions and maintains normoglycaemia throughout life (Fig. 1). The roles of increased cell number and glucose competence have been investigated in animals and humans. Histomorphometric analysis of pancreatic autopsy samples revealed a higher beta cell mass in the pancreas from obese, insulinresistant individuals, compared with samples from normal-weight individuals, but the beta cell mass in the pancreas from individuals with type 2 diabetes was reduced in association with increased signs of apoptosis [2]. These ndings suggest that reduction in beta cell mass may underlie the decreased insulin secretion capacity. However, further analysis showed that the reduction in beta cell mass was proportional to the time from onset of diabetes and that hyperglycaemia probably devel-

oped when beta cell mass was still within normal levels [3]. This implies that reduced glucose competence of a normal number of beta cells may lead to the onset of diabetes. Analysis of the data presented in this last publication also shows that there is no relation between beta cell mass and insulin secretion capacity, with a much higher beta cell mass in the pancreas from some diabetic subjects than from many normal individuals. Thus, there is no direct correlation between beta cell mass and glycaemic control, suggesting that glucose competence of individual beta cells is a major factor in determining pancreatic endocrine function. Similar conclusions have been drawn from the results of animal studies. In particular, in a study of genetically obese and diabetic mice (ob/ob or db/db mice), it was shown that beta cell mass and plasma insulin levels were markedly increased during the progression of obesity and insulin resistance. However, after a few months of diabetic hyperglycaemia, a reduction in beta cell mass and hypoinsulinaemia developed [4, 5]. In a model of nutrition-induced metabolic stress, mice fed a high-fat diet that rapidly developed insulin resistance leading to compensatory insulin secretion capacity [6, 7]. This response is highly inuenced by the genetic background of the mice studied [8 11]. In humans, genome-wide association studies have been used to identify single-nucleotide variants in or in close proximity to more than 50 genes

2013 The Association for the Publication of the Journal of Internal Medicine

203

B. Thorens

Review: Beta cell research for better therapy

Insulin resistance in obesity or pregnancy


Compensatory -cell proliferation Increased glucose competence

Replication of mature -cells Neogenesis from precursors Transdifferentiation


Decreased glucose competence -cell apoptosis


Normal state

Decreased glucose competence

Type 2 Diabetes
-cell apoptosis Reduced glucose competence Decreased -cell mass

Glucose intolerance with normal body weight


Normal -cell mass Reduced glucose competence

Fig. 1 Plasticity of beta cell mass and function. From an initial beta cell mass (normal state), new beta cells can be generated in response to insulin resistance associated with obesity or pregnancy. This compensatory response probably involves replication of mature beta cells and neogenesis from precursors. Independently of insulin resistance and obesity, progressive loss of beta cell glucose competence may also develop because of combined genetic susceptibilities and metabolic stresses. The nal progression to type 2 diabetes is associated with reduction in beta cell mass caused by imbalance between apoptosis and neoformation.

that inuence susceptibility to type 2 diabetes. It is noteworthy that most of these genes are expressed in beta cells and participate either in the glucose signalling pathway that controls insulin secretion or in various transcriptional mechanisms that control beta cell differentiation and proliferation [1215]. Thus, to understand the pathogenesis of type 2 diabetes, the critical signalling or metabolic pathways that control beta cell proliferation and glucose competence must be identied and how the activity of these pathways is inuenced by individual genetic variability must be determined. Here, the evidence for plasticity of adult beta cell mass and function is reviewed, together with the respective signalling pathways involved. In addition, potential future lines of research are explored. Adult beta cell plasticity Over the last decade, a major focus in beta cell research has been the investigation of beta cell generation from precursor cells or from embryonic
204 2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214

stem cells. This effort led to an understanding of the transcriptional control of pancreatic endocrine development during embryogenesis [1618]. It also led to the establishment of the optimal conditions to differentiate mouse and human embryonic stem cells into endodermal progenitors, which could further be differentiated into endocrine cells, including insulin-secreting cells [1921]. More recently, through genetic techniques allowing immortalization of human embryonic beta cells, protocols have been established to generate human beta cell lines capable of unlimited replication whilst preserving glucose-stimulated insulin secretion and other features of adult beta cells, such as sensitivity to the action of gluco-incretin hormones [22]. These cell lines are vital for the study of human beta cell biology; however, whether they can also be used for the treatment of diabetes, for instance using cell transplantation techniques, is still unclear. In comparison with the study of beta cell development and differentiation, progress in

B. Thorens

Review: Beta cell research for better therapy

understanding the molecular control of adult beta cell plasticity and in dening the methods to manipulate it has been more modest. In theory, controlling beta cell number can be achieved by increasing beta cell proliferation, inducing beta cell differentiation from precursors or protecting mature beta cells against apoptosis; this last process is caused by the combination of inammatory cytokines released locally by inammatory or immune cells or secreted by adipose tissue or muscle [23, 24] and high plasma levels of glucose and free fatty acids, that is, glucolipotoxicity [25]. Replication following destruction of adult beta cells In normal physiological conditions, adult beta cell replication, although very difcult to precisely assess (especially in humans), is considered to be very low. Findings from studies in mice and rats suggest that the rate of beta cell renewal is ~3% per day, that is, complete replacement every month [26], and replication appears to be the major pathway to beta cell neoformation [27]. The rate of replication is high in the early postnatal period and declines rapidly in adult animals. In humans, it has been suggested that once the full complement of beta cells has been generated in young adults, almost no replication occurs later in life [28]. Beta cell expansion can be induced experimentally in adult animals in several ways: (i) in response to partial pancreatectomy [29], (ii) in response to destruction of beta cells by diphtheria toxin treatment in transgenic mice expressing the diphtheria toxin receptor in their beta cells [30], or (iii) following induction of an inammatory response caused by wrapping the pancreas with cellophane [31, 32] or by pancreatic duct ligation [33]. The mechanism of beta cell neoformation varies depending on the experimental protocol. Following pancreatic duct ligation, new beta cells are formed from precursor cells recruited from an unknown source to the pancreatic duct. When beta cells are destroyed by diphtheria toxin, their regeneration mostly results from the transdifferentiation of alpha cells [30]. Transdifferentiation was also observed in transgenic mice expressing the transcription factor Pax4 in alpha cells. This lead to a massive increase in beta cell mass, which could be sustained over time because the disappearance of alpha cells resulted in the recruitment to duct and islets of new precursors able to differentiate into Pax-4-expressing alpha cells [34]. Beta cell neoformation can also proceed from the dedifferentiation

of exocrine cells into ductal-like cells, which can then redifferentiate into mature beta cells [3537]. It has also been proposed that beta cells may originate in the pancreatic ducts, in which precursor cells have been located. However, the importance of this pathway for beta cell regeneration in adult mice is still debated [3840]. Thus, there is ample evidence that new beta cells can be generated in adult animals, in response to various experimental conditions and using different mechanisms (Fig. 2). This indicates that total beta cell mass is constantly monitored and that signals are produced to induce new beta cell formation. The nature of the signals and whether they differ under the various regeneration conditions discussed above remain unknown. It is an important challenge of current research to identify genes expressed in these conditions, either by beta cells themselves or possibly also by alpha or duct cells, as well as genes that trigger beta cell neoformation. Beta cell replication in insulin-resistant conditions The mechanisms leading to a compensatory increase in beta cell number in insulin resistance are also not known. In the setting of obesity and insulin resistance, hyperglycaemic episodes may occur during the phase of beta cell compensation. As glucose is one of the most potent stimulators of beta cell proliferation [41, 42], it may induce
Pax4
P

3 EX

-cell -cell Precursor cell Duct cell

1 4

EX Exocrine cell

Fig. 2 Multiple paths to beta cell neoformation. In the adult mouse pancreas, new beta cells can be generated by replication of existing mature beta cells (1). Alpha cells can transdifferentiate into new beta cells either following beta cell destruction or following targeted overexpression of the transcription factor Pax4 in alpha cells, which leads to recruitment of progenitor cells to feed massive transdifferentiation of alpha cells into beta cells (2). Exocrine cells can dedifferentiate into duct-like cells, which can be converted into beta cells (3). Precursors present in pancreatic ducts may also provide a source of new beta cells (4).
2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214 205

B. Thorens

Review: Beta cell research for better therapy

compensatory beta cell growth. Glucose induces beta cell proliferation by a mechanism that requires its metabolism and closure of KATP channels [43, 44] leading to membrane depolarization and insulin granule exocytosis. This leads to the secretion not only of insulin but also of other peptides such as insulin-like growth factor (IGF)-2, which could act as autocrine regulators of the insulin and IGF-1 receptors. As beta cell expansion in genetic models of insulin resistance requires expression of the insulin receptor substrate-2 (IRS2) in beta cells, this supports the hypothesis that regulation of beta cell mass can involve activation of the insulin or IGF-2 receptors [45, 46]. The IGF-1 receptor/IRS-2/Akt pathway has also been linked to increased beta cell glucose competence [47], indicating that proliferation and glucose competence may, in some situations at least, be regulated simultaneously. It has also been proposed that secreted factors, for example released by insulin-resistant muscle, can increase beta cell proliferation [24]. Also, parabiosis experiments carried out between control mice and mice with liver-specic knockout of the insulin receptor, which have massive beta cell compensatory expansion, induce beta cell proliferation in the control animals; this suggests that factors released by the insulin-resistant liver can stimulate beta cell proliferation [48]. Neuronal signals may also be involved. This has been demonstrated in a model of liver insulin resistance induced by activation of the extracellular signal regulated kinase activation of the extracellular signal regulated kinases (Erk1, Erk2),(Erk1/2) kinase pathway specically in this organ [49]. This led to a remarkable increase in beta cell proliferation, which appears to be entirely mediated by a neuronal pathway linking the liver to the endocrine pancreas. Inammation of the endocrine pancreas, with inltration of macrophages and other inammatory cells in the islets, is a hallmark of type 2 diabetes in humans and mice [50]. This is associated with production of cytokines, which not only involves glucose-induced interleukin (IL)-1 production by beta cells and autocrine activation of the Fas pathway but also secretion by activated inammatory cells [5053]. At low levels of IL-1 expression and Fas activation by beta cells, this signal may induce beta cell proliferation, especially when the intracellular signalling molecule Flip is expressed [53, 54]; this pathway may link initial, low-grade inammation to adaptation of beta cell mass.
206 2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214

There is thus very strong evidence for the involvement of metabolic, endocrine and nervous signals in the adaptation of beta cell mass to insulin resistance in liver, muscle and fat. However, the identity of these signals, how they are generated and by which tissue(s), is still far from being understood. Beta cell replication during pregnancy Pregnancy is an insulin-resistant state that develops to ensure sufcient provision of glucose to the foetus. However, to preserve normoglycaemia, the beta cells of the mother undergo multiple functional changes, including increased glucose-stimulated insulin secretion, increased glucose uptake, phosphorylation and oxidation capacity [55] and a large increase in beta cell mass. In mice, a peak of proliferation is observed at day 14 of gestation and the maximum increase in beta cell mass, reaching ~150% of the prepregnancy mass, is observed by day 19 of gestation. Following delivery, a phase of rapid apoptosis ensues to normalize the beta cell mass [56, 57]. An increase in beta cell mass during pregnancy in humans has also been reported [58]. In rodents, beta cell proliferation as well as functional changes leading to increased glucose-stimulated insulin secretion appears to be mostly under the control of prolactin and the placental lactogen acting through activation of the prolactin receptor (PRL-R)/Jak/STAT signalling pathway [55, 59]. This activates the transcription factor FoxM1, which induces the expression of several cell cycle regulators [60] but also suppresses the expression of the multiple endocrine neoplasia 1 gene (menin1), which leads to reduced expression of the cell cycle inhibitors p18 and p27 [6163] (Fig. 3). It is interesting that activation of the PRL-R induces substantial expression of the enzyme tryptophan hydroxylase, leading to serotonin production and autocrine activation of the serotonin receptor 5HTR2B[64, 65]. Further, the role of the cell surface oestradiol receptor GPR30 in inducing beta cell proliferation has recently been demonstrated; its mechanism of signalling involves the silencing of the microRNA mir338-3p, a negative regulator of the IGF-1 receptor signalling pathway [66]. In humans, the normal beta cell mass expansion during pregnancy may be blunted in gestational diabetes mellitus. The cause of this impaired expansion response is not known but is certainly associated with gene variants leading to an

B. Thorens

Review: Beta cell research for better therapy

PRL/PL IGF-1R/IRS2 GPR30 E2 PRL-R

5-HT

5HTR2B

mir338-3p STAT cAMP/PKA Pi-Akt Menin FoxoM1


cdc25A cdc25B cyclinB1 CENP-F Plk-1 AuroraB

Tph1

5-HT

Gq/11

Bclxl

Apoptosis

p18, p27

Proliferation
Fig. 3 Signalling pathways that control beta cell expansion in pregnancy. In mice, the beta cell proliferation rate is maximal at day 14 of gestation, and beta cell mass expansion reaches a peak at day 19. Proliferation is largely controlled by prolactin (PRL) and placental lactogen (PL) activating the PRL receptor (PRL-R)/STAT pathway. This activates the transcription factor FoxM1, which induces the indicated regulators of cell cycle progression; it also suppresses the expression of the multiple endocrine neoplasia 1 gene (Men1), an inducer of cell cycle inhibitors p18 and p27, induces the expression of the anti-apoptotic gene Blcxl and leads to massive induction of tryptophan hydroxylase (Tph1). This results in production of serotonin, an autocrine inducer of proliferation through activation of the serotonin receptor 5HTR2B. Separately, activation of the oestrogen receptor GPR30, through suppression of mir388-3p expression, causes increased expression of the IGF-1 receptor (IGF-1R) and its signalling pathway. All pathways converge to stimulate beta cell proliferation.

improper proliferation response to the pregnancy hormones. As gestational diabetes is associated with increased risk of developing type 2 diabetes later in life [67, 68], this suggests that the adaptive mechanisms that lead to beta cell proliferation in response to the transient insulin resistance of pregnancy also play a role in the life-long adaptation of beta cell mass and function. Identication of the genes conferring susceptibility to gestational diabetes mellitus would be of great interest. Currently available evidence suggests the participation of the already identied type 2 diabetes genes CDKAL1 and MTNR1B [67]. Gluco-incretins and regulation of beta cell mass and function The gluco-incretin hormones glucagon-like peptide-1 (GLP-1) and gastric inhibitory polypeptide (GIP) have direct impact on the function of the pancreatic beta cells by binding to specic

receptors located on their cell surface [69, 70]. Binding triggers intracellular signalling mechanisms initiated by the production of cAMP and activation of protein kinase A and Epac2, a cAMPbinding protein [71]. The immediate effect of these events is the potentiation of glucose-induced insulin secretion [72, 73]; this is an important control mechanism as it is estimated that gluco-incretin action on beta cells is responsible for ~50% of insulin secreted in the absorptive phase [74]. This acute effect of GLP-1, but not GIP, is preserved in patients with type 2 diabetes, although supraphysiological concentrations of GLP-1 are needed to trigger insulin secretion and normalize glucose levels in the blood [75]. Nevertheless, various GLP-1 receptor agonists, as well as inhibitors of the enzyme dipeptidylpeptidase-4 (which rapidly inactivates endogenous GLP-1), have most recently been introduced for the treatment of type 2 diabetes [76].
2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214 207

B. Thorens

Review: Beta cell research for better therapy

Besides this acute insulinotropic effect, both GLP-1 and GIP also have trophic effects leading to increased beta cell proliferation [7779], protection against cytokine- and glucolipotoxicity-induced apoptosis [80, 81] and increased glucose competence [82, 83]. In rodents, these effects combine to effectively increase beta cell mass and even protect beta cells against autoimmune destruction in the NOD mouse model of type 1 diabetes [84, 85]. As shown in Fig. 4, several intracellular signalling pathways are activated downstream of the initial cAMP production, which combine to control the trophic actions of GLP-1. First, the classical cAMP/ protein kinase A/cAMP response element-binding protein (CREBP) pathway controls the expression of many genes. Secondly, activation of the Erk1/2 pathway requires simultaneous Ca2+ uptake (induced by high extracellular glucose) and release of Ca2+ from the endoplasmic reticulum. Thirdly, the IRS-2/Pi3K/Akt pathway plays a major role in protecting beta cells against apoptosis, inducing proliferation and increasing glucose competence [8688]. Fourthly, Cornu et al. [83, 89] showed that the trophic actions of GLP-1 were dependent on increased expression of the IGF-1 receptor and its autocrine activation by IGF-2 produced by the beta cells. This autocrine loop controls beta cell plasticity and transmits the GLP-1 signal. Finally, a role of the Wnt signalling pathway in GLP-1 action has also been suggested. This is activated by stabilization of -catenin secondary to activation of the Akt, Erk1/2 and PKA pathways. This leads to expression of transcription factor 7-like 2 (TCF7L2) [90], a diabetes susceptibility gene [12], which controls glucose-stimulated insulin secretion, in part by regulating GLP-1 receptor expression [91 93]. Whether GLP-1 and GIP have similar trophic effects on human beta cells is unclear. Good evidence supports a role for GLP-1 in protecting against cytokine- and glucolipotoxicity-induced apoptosis [80]. However, attempts to induce human beta cell proliferation in vitro using GLP-1 have so far been disappointing [94, 95], and there is an urgent need to determine whether mouse and human beta cells respond similarly to the action of gluco-incretins. Long-term treatment with GLP-1 receptor agonists is very effective in controlling glycaemia in patients with type 2 diabetes, but diabetes quickly resumes after treatment cessation. This indicates that there
208 2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214

is no long-term improvement of beta cell function [96], although it was very recently reported that beta cell mass was strongly increased in the pancreas of patients with type 2 diabetes treated with GLP-1 agonists or dipeptidyl-peptidase IV inhibitors [97]. These observations need to be conrmed, a task that is, however, particularly difcult in the absence of proper imaging techniques for in vivo assessment of beta cell mass and function. One important observation is that the increase in mouse islet proliferation induced by GLP-1 or other growth factors is usually very modest, with 1% to ~5% of the beta cell population showing signs of progression through the cell cycle. In GLP-1treated cells, this low level of proliferation has been linked to the induction by GLP-1 of multiple mechanisms that limit its own signalling pathway. Indeed, GLP-1-induced proliferation requires activation of the PKA/CREBP, PI3K and Erk1/2 signalling pathways. However, immediately after GLP-1 binding to its receptor, multiple suppressors of these signalling pathways are induced, including RGS2 (an inhibitor of Gsa activation and cAMP production), ICER and CREM (inhibitors of CREBP) and DUSP14 (a dual-specicity phosphatase that inactivates Erk1/2 signalling) [98]. Knockdown of these negative regulators of signalling increases GLP-1-induced beta cell proliferation. Thus, beta cells have evolved mechanisms to limit their proliferative response to growth factors, probably because over secretion of insulin can be lethal. Therefore increasing beta cell mass may not only need to target the pathways that induce proliferation, but also those that prevent over-responsiveness to stimuli. Proliferation, glucose competence and nutrient-regulated enzymes Beta cells are highly sensitive to the levels of circulating nutrients, which control the acute insulin secretion response but also the long-term adaptation of beta cell mass. In recent years, several nutrient-sensing enzymes have been identied that are activated by changing levels of nutrients or of specic metabolites. PAS kinase The serine/threonine protein kinase PAS kinase is a sensor of elevated glucose concentrations that has evolved from a large family of prokaryotic kinases containing the conserved Per-Arnt-Sim

B. Thorens

Review: Beta cell research for better therapy

2
Ca2+ VDCC KATP K+ Glucose GLUT2

1
GLP-1 IGF-1R

IGF-2

RGS2 Ca2+ ATP/ADP Glucose cAMP Epac2 PKA RYR Endoplasmic reticulum Ca2+ Ras/Raf DUSP14 MAPK/Erk1/2 CREM ICER CREBP Akt/PKB IRS-2

IGF-2 Insulin

PI3K

4
-catenin

TCF7L2

Nucleus

GLUT2, Glucokinase, IGF-1R, IRS-2 Insulin, Pdx-1, c-fos, cyclins

Apoptosis

Proliferation

Glucose competence

Fig. 4 Multiple intracellular pathways activated by GLP-1 to increase beta cell functional mass. Activation of the GLP-1 receptor induces several intracellular signalling pathways: (1) the classical cAMP/protein kinase A (PKA) pathway that activates the transcription factor CREBP; (2) the MAP kinase/Erk1/2 signalling pathway that requires interaction with the glucose signalling pathway (green box) to induce Ca2+ release from the endoplasmic reticulum through activation by Ca2+ and Epac2 of the ryanodine receptor (RYR); (3) induction of IGF-1 receptor (IGF-1R) expression, which becomes activated by the autocrine factor IGF-2 cosecreted with insulin; (4) activation of -catenin/TCF7L2 by the combined action of PKA, MAP kinases and AKT. These pathways activate the transcription of the indicated (and other) genes involved in glucosestimulated insulin secretion, beta cell differentiation and proliferation. Of note, GLP-1 signalling also induces the rapid and strong induction of negative regulators of its own signalling: RGS2, which prevents activation of cAMP production, CREM and ICER, which antagonize CREBP activity, and DUSP14, a dual-specicity phosphatase which de-activates the MAP kinase pathway.

(PAS) sensor domain [99]. In beta cells, activation of this kinase induces translocation of the transcription factor Pdx-1 in the nucleus and increases insulin gene transcription and glucosestimulated insulin secretion [100, 101]. Thus, PAS kinase is a regulator of glucose competence, and

its expression is reduced in islets from type 2 diabetic individuals [102]. It is also expressed in alpha cells and studies with gene knockout mice suggest that the role of PAS kinase in these cells is to limit glucagon gene expression and secretion [102].
2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214 209

B. Thorens

Review: Beta cell research for better therapy

Mammalian target of rapamycin (mTOR) mTOR is a serine/threonine kinase that is found in two forms, mTORC1 and mTORC2, with different substrate specicities [103]. mTORC1 has a role in the control of beta cell size and proliferation, in response to branched-chain amino acids, and possibly also glucose, or growth factors that cause the induction of protein kinase Cf [104108]. Constitutive activation of mTORC1 in beta cells by genetic inactivation of the upstream TSC1/2 regulatory genes induces increased beta cell mass and hypoglycaemia, and proliferation is associated with regulation of the cell cycle regulators cyclin D2, cyclin D3 and Cdk4. Because mTORC1 is inhibited by rapamycin, an immunosuppressive drug used in organ transplantations, treatment with this drug negatively inuences beta cell mass and function [109, 110]. Sirt1 Mammalian sirtuins comprise a family of NAD+dependent protein deacetylases including Sirt1, which has been extensively investigated for its role in the control of cellular metabolism and ageing [111, 112]. Sirt1 is activated by fasting, when the intracellular NAD+/NADH ratio increases or by the polyphenol compound resveratrol. Sirt1 thus regulates the activity of enzymes, transcription factors, histones and structural proteins by inducing their deacetylation. In beta cells, activation of Sirt1 leads to a coordinated increased expression of Glut2, glucokinase, Pdx-1, (pancreatic and duodenal homeobox 1), HNF1a, (HNF1a : hepatic transcription factor 1) and Tfam (: transcription factor A, mitochondrial UCP2 : uncoupling protein 2), and suppression of UCP2 expression, resulting in increased ATP production and glucose-stimulated insulin secretion [113115]. An important action of Sirt1 is to deacetylate the tumour suppressor gene LKB1, an upstream regulator of AMP kinase. When deacetylated, LKB1 phosphorylates and activates AMP kinase and several AMP kinase-related kinases [116]. Genetic inactivation of LKB1 induces a massive increase in beta cell mass and loss of cellular polarity [117 119], effects that are most probably due to inactivation of several kinases as genetic inactivation of AMP kinase does not induce beta cell proliferation. Of interest, a mutation in Sirt1 was recently identied in a family with a history of type 1 diabetes. Cellular studies of the Sirt1 mutant showed that its expression in beta cells caused
210 2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214

increased nitric oxide and cytokine production, suggesting a possible role in beta cell destruction in these patients [120]. AMP kinase This is an evolutionarily conserved kinase that acts as a sensor of low-nutrient conditions and is particularly activated during hypoglycaemia [121, 122]. It is a trimeric protein composed of one of two a subunits (a1 or a2), one of two b subunits (b1 or b2) and one of three c subunits (c1, c2 or c3). Activation of AMP kinase depends on an increase in the intracellular AMP/ATP ratio, but full activity requires further phosphorylation of the a subunit on threonine 172 by the upstream kinase LKB1, itself regulated by deacetylation by Sirt1, or by CamKK1, a protein kinase activated by Ca2+. AMP kinase is also activated by the antidiabetic drug metformin, and therefore, it is important to understand its role in beta cell function. Unfortunately, this role is currently debated with several studies demonstrating that activation of AMP kinase increases glucose-stimulated insulin secretion, whereas others show the opposite, as comprehensively reviewed recently [123]. One difculty in studying the physiological role of AMP kinase in beta cell biology is that this enzyme is activated when glucose levels fall well below the normoglycaemic level. It is thus difcult to understand how it can acutely regulate glucose-stimulated insulin secretion. It may rather be an important sensor of hypoglycaemia or of nutrient deprivation that affects long-term adaptation of beta cells to these challenging conditions. Although an important sensor of energy status, its precise role in beta cell biology remains to be understood. Summary and future challenges Beta cells can display a marked plasticity under physiological conditions, with modulation of both number and glucose competence. Type 2 diabetes results when this plasticity fails to compensate for the developing insulin resistance, possibly initiated by a defect in glucose competence followed by a decrease in beta cell number. There is now extensive knowledge of the pathways controlling beta cell proliferation, yet insufcient to develop rational ways to increase beta cell mass. In particular, the diversity of mechanisms that limit beta cell proliferation remains poorly understood. It is striking that all the stimuli that have been reported to increase beta cell proliferation have similar modest effect, suggesting that the mechanisms

B. Thorens

Review: Beta cell research for better therapy

limiting proliferation are very potent. More investigations of these mechanisms are required to enable manipulation of beta cell mass. It is also important to note that most of our present knowledge is derived from the study of rodent beta cells and it is not clear that human beta cells will behave in exactly the same way. It is thus critical to study human beta cells from normal individuals and patients with diabetes. Recently generated human beta cell lines can also provide increased understanding of human beta cell biology. Reliable imaging techniques, which would allow in vivo visualization of beta cells and assessment of their secretion capacity, are still lacking to study the pathogenesis of type 2 diabetes and the response to therapeutic treatments. Intensive research activities are ongoing to develop multiple modes of beta cell imaging, and some lines of investigations are already producing interesting results as discussed in recent excellent reviews [124126]. Finally, when the signalling pathways controlling beta cell proliferation and glucose competence are fully elucidated, two challenges will remain to understand (i) how individual genetic variability impacts on beta cell cell function and susceptibility to deregulation by various metabolic stresses and ageing, and (ii) how these pathways can be targeted by pharmacological intervention, nutrition or exercise. Based on the great advances made in recent years and the importance of current challenges to improve health, there is clearly a need for strong commitments from the research community and funding bodies to better support adult beta cell research to design rational and long-term ways to preserve the insulin secretion capacity of the endocrine pancreas. Conict of interest statement No conicts of interest to declare. Acknowledgements Work in the authors laboratory has been supported by grants from the Swiss National Science Foundation (3100A0-113525), the National Center of Competence in Research Frontiers in Genetics and the Innovative Medicine Initiative Joint Undertaking under grant agreement no. 155005 (IMIDIA), resources of which are composed of

nancial contribution from the European Unions Seventh Framework Programme (FP7/20072013) and EFPIA companies in kind contribution and European Unions Seventh Framework Programme Integrated Project BetaBat.

References
1 Prentki M, Nolan CJ. Islet beta cell failure in type 2 diabetes. J Clin Invest 2006; 116: 180212. 2 Butler AE, Janson J, Bonner-Weir S, Ritzel R, Rizza RA, Butler PC. Beta-cell decit and increased beta-cell apoptosis in humans with type 2 diabetes. Diabetes 2003; 52: 10210. 3 Rahier J, Goebbels RM, Henquin JC. Cellular composition of the human diabetic pancreas. Diabetologia 1983; 24: 366 71. 4 Coleman DL, Hummel KP. The inuence of genetic background on the expression of the obese (Ob) gene in the mouse. Diabetologia 1973; 9: 28793. 5 Coleman DL. Obese and diabetes: two mutant genes causing diabetes-obesity syndrome in mice. Diabetologia 1978; 14: 1418. 6 Surwit RS, Kuhn CM, Cochrane C, McCubbin JA, Feinglos MN. Diet-induced type II diabetes in C57BL/6J mice. Diabetes 1988; 37: 11637. 7 Surwit RS, Feinglos MN, Rodin J et al. Differential effects of fat and sucrose on the development of obesity and diabetes in C57BL/6J and A/J mice. Metabolism 1995; 44: 64551. 8 Andrikopoulos S, Massa CM, Aston-Mourney K et al. Differential effect of inbred mouse strain (C57BL/6, DBA/2, 129T2) on insulin secretory function in response to a high fat diet. J Endocrinol 2005; 187: 4553. 9 Kooptiwut S, Zraika S, Thorburn AW et al. Comparison of insulin secretory function in two mouse models with different susceptibility to beta-cell failure. Endocrinology 2002; 143: 208592. 10 Rossmeisl M, Rim JS, Koza RA, Kozak LP. Variation in type 2 diabetesrelated traits in mouse strains susceptible to diet-induced obesity. Diabetes 2003; 52: 195866. 11 Parks BW, Nam E, Org E et al. Genetic control of obesity and gut microbiota composition in response to high-fat, high-sucrose diet in mice. Cell Metab 2013; 17: 14152. 12 Sladek R, Rocheleau G, Rung J et al. A genome-wide association study identies novel risk loci for type 2 diabetes. Nature 2007; 445: 8815. 13 Groop L, Lyssenko V. Genetic basis of beta-cell dysfunction in man. Diabetes Obes Metab 2009; 11(Suppl 4): 14958. 14 Bonnefond A, Froguel P, Vaxillaire M. The emerging genetics of type 2 diabetes. Trends Mol Med 2010; 16: 40716. 15 McCarthy MI, Zeggini E. Genome-wide association studies in type 2 diabetes. Curr Diabetes Rep 2009; 9: 16471. 16 Noguchi H. Production of pancreatic beta-cells from stem cells. Curr Diabetes Rev 2010; 6: 18490. 17 Rojas A, Khoo A, Tejedo JR, Bedoya FJ, Soria B, Martin F. Islet cell development. Adv Exp Med Biol 2010; 654: 5975. 18 Van Hoof D, DAmour KA, German MS. Derivation of insulin-producing cells from human embryonic stem cells. Stem Cell Res 2009; 3: 7387. 19 Kroon E, Martinson LA, Kadoya K et al. Pancreatic endoderm derived from human embryonic stem cells generates 2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214 211

B. Thorens

Review: Beta cell research for better therapy

20

21

22

23

24

25 26

27

28

29

30

31

32

33

34

35 36

37

38

glucose-responsive insulin-secreting cells in vivo. Nat Biotechnol 2008; 26: 44352. DAmour KA, Bang AG, Eliazer S et al. Production of pancreatic hormone-expressing endocrine cells from human embryonic stem cells. Nat Biotechnol 2006; 24: 1392401. DAmour KA, Agulnick AD, Eliazer S, Kelly OG, Kroon E, Baetge EE. Efcient differentiation of human embryonic stem cells to denitive endoderm. Nat Biotechnol 2005; 23: 153441. Ravassard P, Hazhouz Y, Pechberty S et al. A genetically engineered human pancreatic beta cell line exhibiting glucose-inducible insulin secretion. J Clin Invest 2011; 121: 358997. Donath MY, Ehses JA, Maedler K et al. Mechanisms of beta-cell death in type 2 diabetes. Diabetes 2005; 54(Suppl 2): S10813. Bouzakri K, Plomgaard P, Berney T, Donath MY, Pedersen BK, Halban PA. Bimodal effect on pancreatic beta-cells of secretory products from normal or insulin-resistant human skeletal muscle. Diabetes 2011; 60: 111121. Poitout V, Robertson RP. Glucolipotoxicity: fuel excess and {beta}-cell dysfunction. Endocr Rev 2008; 29: 35166. Finegood DT, Scaglia L, Bonner-Weir S. Dynamics of beta-cell mass in the growing rat pancreas. Estimation with a simple mathematical model. Diabetes 1995; 44: 24956. Dor Y, Brown J, Martinez OI, Melton DA. Adult pancreatic beta-cells are formed by self-duplication rather than stem-cell differentiation. Nature 2004; 429: 416. Cnop M, Hughes SJ, Igoillo-Esteve M et al. The long lifespan and low turnover of human islet beta cells estimated by mathematical modelling of lipofuscin accumulation. Diabetologia 2010; 53: 32130. Lee HC, Bonner-Weir S, Weir GC, Leahy JL. Compensatory adaption to partial pancreatectomy in the rat. Endocrinology 1989; 124: 15715. Thorel F, Nepote V, Avril I et al. Conversion of adult pancreatic alpha-cells to beta-cells after extreme beta-cell loss. Nature 2010; 464: 114954. Rafaeloff R, Qin XF, Barlow SW, Rosenberg L, Vinik AI. Identication of differentially expressed genes induced in pancreatic islet neogenesis. FEBS Lett 1996; 378: 21923. Rosenberg L, Duguid WP, Brown RA, Vinik AI. Induction of nesidioblastosis will reverse diabetes in Syrian golden hamster. Diabetes 1988; 37: 33441. Xu X, DHoker J, Stange G et al. Beta cells can be generated from endogenous progenitors in injured adult mouse pancreas. Cell 2008; 132: 197207. Collombat P, Xu X, Ravassard P et al. The ectopic expression of Pax4 in the mouse pancreas converts progenitor cells into alpha and subsequently beta cells. Cell 2009; 138: 44962. Minami K, Seino S. Pancreatic acinar-to-beta cell transdifferentiation in vitro. Front Biosci 2008; 13: 582437. Minami K, Okuno M, Miyawaki K et al. Lineage tracing and characterization of insulin-secreting cells generated from adult pancreatic acinar cells. Proc Natl Acad Sci USA 2005; 102: 1511621. Baeyens L, Bouwens L. Can beta-cells be derived from exocrine pancreas? Diabetes Obes Metab 2008; 10(Suppl 4): 1708. Kushner JA, Weir GC, Bonner-Weir S. Ductal origin hypothesis of pancreatic regeneration under attack. Cell Metab 2010; 11: 23.

39 Solar M, Cardalda C, Houbracken I et al. Pancreatic exocrine duct cells give rise to insulin-producing beta cells during embryogenesis but not after birth. Dev Cell 2009; 17: 84960. 40 Inada A, Nienaber C, Katsuta H et al. Carbonic anhydrase II-positive pancreatic cells are progenitors for both endocrine and exocrine pancreas after birth. Proc Natl Acad Sci USA 2008; 105: 199159. 41 Sjoholm A. Intracellular signal transduction pathways that control pancreatic beta-cell proliferation. FEBS Lett 1992; 311: 8590. 42 Scharfmann R, Basmaciogullari A, Czernichow P. Effect of growth hormone and glucose on rat islet cells replication using 5-bromo-2-deoxyuridine incorporation. Diabetes Res 1990; 15: 13741. 43 Popiela H, Moore W. Tolbutamide stimulates proliferation of pancreatic beta cells in culture. Pancreas 1991; 6: 4649. 44 Porat S, Weinberg-Corem N, Tornovsky-Babaey S et al. Control of pancreatic beta cell regeneration by glucose metabolism. Cell Metab 2011; 13: 4409. 45 Kubota N, Tobe K, Terauchi Y et al. Disruption of insulin receptor substrate 2 causes type 2 diabetes because of liver insulin resistance and lack of compensatory beta-cell hyperplasia. Diabetes 2000; 49: 18809. 46 Withers DJ, Sanchez Gutierrez J, Towery H et al. Disruption of IRS-2 causes type 2 diabetes in mice. Nature 1998; 391: 9004. 47 Kulkarni RN, Holzenberger M, Shih DQ et al. beta-cell-specic deletion of the Igf1 receptor leads to hyperinsulinemia and glucose intolerance but does not alter beta-cell mass. Nat Genet 2002; 31: 1115. 48 El Ouaamari A, Kawamori D, Dirice E et al. Liver-derived systemic factors drive beta cell hyperplasia in insulin-resistant states. Cell Rep 2013; 3: 40110. 49 Imai J, Katagiri H, Yamada T et al. Regulation of pancreatic beta cell mass by neuronal signals from the liver. Science 2008; 322: 12504. 50 Ehses JA, Perren A, Eppler E et al. Increased number of islet-associated macrophages in type 2 diabetes. Diabetes 2007; 56: 235670. 51 Maedler K, Sergeev P, Ris F et al. Glucose-induced beta cell production of IL-1beta contributes to glucotoxicity in human pancreatic islets. J Clin Invest 2002; 110: 85160. 52 Maedler K, Spinas GA, Lehmann R et al. Glucose induces beta-cell apoptosis via upregulation of the Fas receptor in human islets. Diabetes 2001; 50: 168390. 53 Maedler K, Schumann DM, Sauter N et al. Low concentration of interleukin-1beta induces FLICE-inhibitory protein-mediated beta-cell proliferation in human pancreatic islets. Diabetes 2006; 55: 271322. 54 Maedler K, Fontana A, Ris F et al. FLIP switches Fas-mediated glucose signaling in human pancreatic beta cells from apoptosis to cell replication. Proc Natl Acad Sci USA 2002; 99: 823641. 55 Weinhaus AJ, Stout LE, Sorensen RL. Glucokinase, hexokinase, glucose transporter 2, and glucose metabolism in islets during pregnancy and prolactin-treated islets in vitro: mechanisms for long term up-regulation of islets. Endocrinology 1996; 137: 16409. 56 Sorenson RL, Brelje TC. Adaptation of islets of Langerhans to pregnancy: beta-cell growth, enhanced insulin secretion and the role of lactogenic hormones. Horm Metab Res 1997; 29: 3017.

212

2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214

B. Thorens

Review: Beta cell research for better therapy

57 Rieck S, Kaestner KH. Expansion of beta-cell mass in response to pregnancy. Trends Endocrinol Metab 2010; 21: 1518. 58 Butler AE, Cao-Minh L, Galasso R et al. Adaptive changes in pancreatic beta cell fractional area and beta cell turnover in human pregnancy. Diabetologia 2010; 53: 216776. 59 Moldrup A, Petersen ED, Nielsen JH. Effects of sex and pregnancy hormones on growth hormone and prolactin receptor gene expression in insulin-producing cells. Endocrinology 1993; 133: 116572. 60 Zhang H, Zhang J, Pope CF et al. Gestational diabetes mellitus resulting from impaired beta-cell compensation in the absence of FoxM1, a novel downstream effector of placental lactogen. Diabetes 2010; 59: 14352. 61 Hughes E, Huang C. Participation of Akt, menin, and p21 in pregnancy-induced beta-cell proliferation. Endocrinology 2011; 152: 84755. 62 Karnik SK, Chen H, McLean GW et al. Menin controls growth of pancreatic beta-cells in pregnant mice and promotes gestational diabetes mellitus. Science 2007; 318: 8069. 63 Zhang H, Li W, Wang Q et al. Glucose-mediated repression of menin promotes pancreatic beta-cell proliferation. Endocrinology 2012; 153: 60211. 64 Kim H, Toyofuku Y, Lynn FC et al. Serotonin regulates pancreatic beta cell mass during pregnancy. Nat Med 2010; 16: 8048. 65 Schraenen A, Lemaire K, de Faudeur G et al. Placental lactogens induce serotonin biosynthesis in a subset of mouse beta cells during pregnancy. Diabetologia 2010; 53: 258999. 66 Jacovetti C, Abderrahmani A, Parnaud G et al. MicroRNAs contribute to compensatory beta cell expansion during pregnancy and obesity. J Clin Invest 2012; 122: 354151. 67 Kwak SH, Jang HC, Park KS. Finding genetic risk factors of gestational diabetes. Genomics Inform 2012; 10: 23943. 68 Baptiste-Roberts K, Barone BB, Gary TL et al. Risk factors for type 2 diabetes among women with gestational diabetes: a systematic review. Am J Med 2009; 122: 20714.e4. 69 Thorens B. Expression cloning of the pancreatic beta cell receptor for the gluco-incretin hormone glucagon-like peptide I. Proc Natl Acad Sci USA 1992; 89: 86415. 70 Usdin TB, Mezey E, Button DC, Brownstein MJ, Bonner TI. Gastric inhibitory polypeptide receptor, a member of the secretin-vasoactive intestinal peptide receptor family, is widely distributed in peripheral organs and the brain. Endocrinology 1993; 133: 286170. 71 Ozaki N, Shibasaki T, Kashima Y et al. cAMP-GEFII is a direct target of cAMP in regulated exocytosis. Nat Cell Biol 2000; 2: 80511. 72 Holst JJ, rskov C, Vagn Nielsen O, Schwartz TW. Truncated glucagon-like peptide I, an insulin-releasing hormone from the distal gut. FEBS Lett 1987; 211: 16974. 73 Mojsov S, Weir GC, Habener JF. Insulinotropin: glucagon-like peptide 1(7-37) co-encoded in the glucagon gene is a potent stimulator of insulin release in the perfused rat pancreas. J Clin Invest 1987; 79: 6169. 74 Preitner F, Ibberson M, Franklin I et al. Gluco-incretins control insulin secretion at multiple levels as revealed in mice lacking GLP-1 and GIP receptors. J Clin Invest 2004; 113: 63545. 75 Nauck MA, Heimesaat MM, Orskov C, Holst JJ, Ebert R, Creutzfeldt W. Preserved incretin activity of glucagon-like peptide 1 (7-36)amide but not of synthetic human gastric

76 77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

inhibitory polypeptide in patients with type-2 diabetes mellitus. J Clin Invest 1993; 91: 3017. Lovshin JA, Drucker DJ. Incretin-based therapies for type 2 diabetes mellitus. Nat Rev Endocrinol 2009; 5: 2629. Perfetti R, Zhou J, Doyle ME, Egan JM. Glucagon-like peptide-1 induces cell proliferation and pancreatic-duodenum homeobox-1 expression and increases endocrine cell mass in the pancreas of old, glucose-intolerant rats. Endocrinology 2000; 141: 46005. Buteau J, Roduit R, Susini S, Prentki M. Glucagon-like peptide-1 promotes DNA synthesis, activates phosphatidylinositol 3-kinase and increases transcription factor pancreatic and duodenal homeobox gene 1 (PDX-1) DNA binding activity in beta (INS-1)-cells. Diabetologia 1999; 42: 85664. Buteau J, Foisy S, Joly E, Prentki M. Glucagon-like peptide 1 induces pancreatic beta-cell proliferation via transactivation of the epidermal growth factor receptor. Diabetes 2003; 52: 12432. Buteau J, El-Assaad W, Rhodes CJ, Rosenberg L, Joly E, Prentki M. Glucagon-like peptide-1 prevents beta cell glucolipotoxicity. Diabetologia 2004; 47: 80615. Li Y, Hansotia T, Yusta B, Ris F, Halban PA, Drucker DJ. Glucagon-like peptide-1 receptor signaling modulates beta cell apoptosis. J Biol Chem 2003; 278: 4718. Hinke SA, Hellemans K, Schuit FC. Plasticity of the beta cell insulin secretory competence: preparing the pancreatic beta cell for the next meal. J Physiol 2004; 558: 36980. Cornu M, Modi H, Kawamori D, Kulkarni RN, Joffraud M, Thorens B. Glucagon-like peptide-1 increases beta-cell glucose competence and proliferation by translational induction of insulin-like growth factor-1 receptor expression. J Biol Chem 2010; 285: 1053845. Liu MJ, Shin S, Li N et al. Prolonged remission of diabetes by regeneration of beta cells in diabetic mice treated with recombinant adenoviral vector expressing glucagon-like peptide-1. Mol Ther 2007; 15: 8693. Zhang J, Tokui Y, Yamagata K et al. Continuous stimulation of human glucagon-like peptide-1 (7-36) amide in a mouse model (NOD) delays onset of autoimmune type 1 diabetes. Diabetologia 2007; 50: 19009. Jhala US, Canettieri G, Screaton RA et al. cAMP promotes pancreatic beta-cells survival via CREB-mediated induction of IRS2. Genes Dev 2003; 17: 157580. Park S, Dong X, Fisher TL et al. Exendin-4 uses Irs2 signaling to mediate pancreatic beta cell growth and function. J Biol Chem 2006; 281: 115968. Buteau J, Spatz ML, Accili D. Transcription factor FoxO1 mediates glucagon-like peptide-1 effects on pancreatic beta-cell mass. Diabetes 2006; 55: 11906. Cornu M, Yang JY, Jaccard E, Poussin C, Widmann C, Thorens B. Glp-1 Protects Beta-Cells Against Apoptosis By Increasing The Activtiy Of An Igf-2/Igf1-Receptor Autocrine Loop. Diabetes 2009; 58: 181625. Liu Z, Habener JF. Glucagon-like peptide-1 activation of TCF7L2-dependent Wnt signaling enhances pancreatic beta cell proliferation. J Biol Chem 2008; 283: 872335. da Silva Xavier G, Loder MK, McDonald A et al. TCF7L2 regulates late events in insulin secretion from pancreatic islet beta-cells. Diabetes 2009; 58: 894905. da Silva Xavier G, Mondragon A, Sun G et al. Abnormal glucose tolerance and insulin secretion in pancreas-specic Tcf7 l2-null mice. Diabetologia 2012; 55: 266776.

2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214

213

B. Thorens

Review: Beta cell research for better therapy

93 Shu L, Matveyenko AV, Kerr-Conte J, Cho JH, McIntosh CH, Maedler K. Decreased TCF7L2 protein levels in type 2 diabetes mellitus correlate with downregulation of GIP- and GLP-1 receptors and impaired beta-cell function. Hum Mol Genet 2009; 18: 238899. 94 Rutti S, Sauter NS, Bouzakri K, Prazak R, Halban PA, Donath MY. In vitro proliferation of adult human beta-cells. PLoS One 2012; 7: e35801. 95 Parnaud G, Bosco D, Berney T et al. Proliferation of sorted human and rat beta cells. Diabetologia 2008; 51: 91100. 96 Bunck MC, Diamant M, Corner A et al. One-year treatment with exenatide improves beta-cell function, compared with insulin glargine, in metformin-treated type 2 diabetic patients: a randomized, controlled trial. Diabetes Care 2009; 32: 7628. 97 Butler AE, Campbell-Thompson M, Gurlo T, Dawson DW, Atkinson M, Butler PC. Marked expansion of exocrine and endocrine pancreas with incretin therapy in humans with increased exocrine pancreas dysplasia and the potential for glucagon-producing neuroendocrine tumors. Diabetes 2013; PMID: 23524641. [Epub ahead of print]. 98 Klinger S, Poussin C, Debril MB, Dolci W, Halban PA, Thorens B. Increasing GLP-1-induced beta-cell proliferation by silencing the negative regulators of signaling cAMP response element modulator-alpha and DUSP14. Diabetes 2008; 57: 58493. 99 Rutter J, Michnoff CH, Harper SM, Gardner KH, McKnight SL. PAS kinase: an evolutionarily conserved PAS domain-regulated serine/threonine kinase. Proc Natl Acad Sci USA 2001; 98: 89916. 100 da Silva Xavier G, Rutter J, Rutter GA. Involvement of Per-Arnt-Sim (PAS) kinase in the stimulation of preproinsulin and pancreatic duodenum homeobox 1 gene expression by glucose. Proc Natl Acad Sci USA 2004; 101: 831924. 101 An R, da Silva Xavier G, Hao HX, Semplici F, Rutter J, Rutter GA. Regulation by Per-Arnt-Sim (PAS) kinase of pancreatic duodenal homeobox-1 nuclear import in pancreatic beta-cells. Biochem Soc Trans 2006; 34: 7913. 102 da Silva Xavier G, Farhan H, Kim H et al. Per-arnt-sim (PAS) domain-containing protein kinase is downregulated in human islets in type 2 diabetes and regulates glucagon secretion. Diabetologia 2011; 54: 81927. 103 Cornu M, Albert V, Hall MN. mTOR in aging, metabolism, and cancer. Curr Opin Genet Dev 2013; 23: 5362. 104 Rachdi L, Aiello V, Duvillie B, Scharfmann R. L-leucine alters pancreatic beta-cell differentiation and function via the mTor signaling pathway. Diabetes 2012; 61: 40917. 105 Xie J, Herbert TP. The role of mammalian target of rapamycin (mTOR) in the regulation of pancreatic beta-cell mass: implications in the development of type-2 diabetes. Cell Mol Life Sci 2012; 69: 1289304. 106 Velazquez-Garcia S, Valle S, Rosa TC et al. Activation of protein kinase C-zeta in pancreatic beta-cells in vivo improves glucose tolerance and induces beta-cell expansion via mTOR activation. Diabetes 2011; 60: 254659. 107 Bartolome A, Guillen C, Benito M. Role of the TSC1-TSC2 complex in the integration of insulin and glucose signaling involved in pancreatic beta-cell proliferation. Endocrinology 2010; 151: 308494. 108 Mori H, Inoki K, Opland D et al. Critical roles for the TSC-mTOR pathway in beta-cell function. Am J Physiol Endocrinol Metab 2009; 297: E101322.

109 Yang SB, Lee HY, Young DM et al. Rapamycin induces glucose intolerance in mice by reducing islet mass, insulin content, and insulin sensitivity. J Mol Med (Berl) 2012; 90: 57585. 110 Zahr E, Molano RD, Pileggi A et al. Rapamycin impairs beta-cell proliferation in vivo. Transplant Proc 2008; 40: 4367. 111 Baur JA, Ungvari Z, Minor RK, Le Couteur DG, de Cabo R. Are sirtuins viable targets for improving healthspan and lifespan? Nat Rev Drug Discov 2012; 11: 44361. 112 Houtkooper RH, Pirinen E, Auwerx J. Sirtuins as regulators of metabolism and healthspan. Nat Rev Mol Cell Biol 2012; 13: 22538. 113 Vetterli L, Brun T, Giovannoni L, Bosco D, Maechler P. Resveratrol potentiates glucose-stimulated insulin secretion in INS-1E beta-cells and human islets through a SIRT1-dependent mechanism. J Biol Chem 2011; 286: 604960. 114 Bordone L, Motta MC, Picard F et al. Sirt1 regulates insulin secretion by repressing UCP2 in pancreatic beta cells. PLoS Biol 2006; 4: e31. 115 Moynihan KA, Grimm AA, Plueger MM et al. Increased dosage of mammalian Sir2 in pancreatic beta cells enhances glucose-stimulated insulin secretion in mice. Cell Metab 2005; 2: 10517. 116 Lan F, Cacicedo JM, Ruderman N, Ido Y. SIRT1 modulation of the acetylation status, cytosolic localization, and activity of LKB1. Possible role in AMP-activated protein kinase activation. J Biol Chem 2008; 283: 2762835. 117 Sun G, Tarasov AI, McGinty JA et al. LKB1 deletion with the RIP2.Cre transgene modies pancreatic beta-cell morphology and enhances insulin secretion in vivo. Am J Physiol Endocrinol Metab 2010; 298: E126173. 118 Granot Z, Swisa A, Magenheim J et al. LKB1 regulates pancreatic beta cell size, polarity, and function. Cell Metab 2009; 10: 296308. 119 Fu A, Ng AC, Depatie C et al. Loss of Lkb1 in adult beta cells increases beta cell mass and enhances glucose tolerance in mice. Cell Metab 2009; 10: 28595. 120 Biason-Lauber A, Boni-Schnetzler M, Hubbard BP et al. Identication of a SIRT1 Mutation in a Family with Type 1 Diabetes. Cell Metab 2013; 17: 44855. 121 Hardie DG, Sakamoto K. AMPK: a key sensor of fuel and energy status in skeletal muscle. Physiology (Bethesda) 2006; 21: 4860. 122 Kahn BB, Alquier T, Carling D, Hardie DG. AMP-activated protein kinase: ancient energy gauge provides clues to modern understanding of metabolism. Cell Metab 2005; 1: 1525. 123 Fu A, Eberhard CE, Screaton RA. Role of AMPK in pancreatic beta cell function. Mol Cell Endocrinol 2013; 366: 12734. 124 Andralojc K, Srinivas M, Brom M et al. Obstacles on the way to the clinical visualisation of beta cells: looking for the Aeneas of molecular imaging to navigate between Scylla and Charybdis. Diabetologia 2012; 55: 124757. 125 Arin DR, Bulte JW. Imaging of pancreatic islet cells. Diabetes Metab Res Rev 2011; 27: 7616. 126 Veluthakal R, Harris P. In vivo beta-cell imaging with VMAT 2 ligandscurrent state-of-the-art and future perspective. Curr Pharm Des 2010; 16: 156881. Correspondence: Bernard Thorens, Center for Integrative Genomics, University of Lausanne, Genopode Building, CH-1015 Lausanne, Switzerland. (fax: + 4121 692 3985; e-mail: Bernard.Thorens@unil.ch).

214

2013 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2013, 274; 203214

Anda mungkin juga menyukai