Review Open Access
Copyright ©The Author(s) 2015. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Diabetes. May 15, 2015; 6(4): 598-612
Published online May 15, 2015. doi: 10.4239/wjd.v6.i4.598
Type 2 diabetes mellitus: From a metabolic disorder to an inflammatory condition
Iqra Hameed, Mudasar Nabi, Khalid Ghazanfar, Department of Biochemistry, University of Kashmir, Srinagar 190006, India
Shariq R Masoodi, Shahnaz A Mir, Department of Endocrinology, Sher-I-Kashmir Institute of medical Sciences, Srinagar 190006, India
Shariq R Masoodi, Division of Endocrinology, Diabetes and Nutrition, University of Maryland School of Medicine, Baltimore, MD 21201, United States
Bashir A Ganai, Centre for Research and Development, University of Kashmir, Hazratbal, Srinagar 190006, India
Author contributions: Hameed I drafted the manuscript; Masoodi SR and Ganai BA conceived and designed the manuscript; Mir SA, Nabi M and Ghazanfar K acquired data, formatted figures/table and revised the manuscript.
Supported by Department of Science and Technology, Government of India to Iqra Hameed, No. Wos-A LS 509/2012.
Conflict-of-interest: All the authors declare that they have no conflict of interest.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Dr. Bashir A Ganai, Professor, Director, Centre for Research and Development, University of Kashmir, Hazratbal Rd, Hazratbal, Srinagar 190006, India. bbcganai@gmail.com
Telephone: +91-979-7247851 Fax: +91-194-2415357
Received: August 29, 2014
Peer-review started: August 30, 2014
First decision: September 30, 2014
Revised: October 14, 2014
Accepted: December 29, 2014
Article in press: December 31, 2014
Published online: May 15, 2015

Abstract

Diabetes mellitus is increasing at an alarming rate and has become a global challenge. Insulin resistance in target tissues and a relative deficiency of insulin secretion from pancreatic β-cells are the major features of type 2 diabetes (T2D). Chronic low-grade inflammation in T2D has given an impetus to the field of immuno-metabolism linking inflammation to insulin resistance and β-cell dysfunction. Many factors advocate a causal link between metabolic stress and inflammation. Numerous cellular factors trigger inflammatory signalling cascades, and as a result T2D is at the moment considered an inflammatory disorder triggered by disordered metabolism. Cellular mechanisms like activation of Toll-like receptors, Endoplasmic Reticulum stress, and inflammasome activation are related to the nutrient excess linking pathogenesis and progression of T2D with inflammation. This paper aims to systematically review the metabolic profile and role of various inflammatory pathways in T2D by capturing relevant evidence from various sources. The perspectives include suggestions for the development of therapies involving the shift from metabolic stress to homeostasis that would favour insulin sensitivity and survival of pancreatic β-cells in T2D.

Key Words: Diabetes mellitus, Inflammation, Insulin resistance, β-cell dysfunction, Adipose tissue

Core tip: Immuno-metabolism, the confluence of metabolism and immune system has emerged as a chief breakthrough especially in the field of diabetes mellitus; a metabolic disorder of great magnitude. Activation of immune system by metabolic stress has opened new insights in the pathogenesis and progression of type 2 diabetes (T2D). The link between metabolic overload and activation of the immune system form the core tip of this review. Metabolic stress can cause pathologic activation of the immune system, thus metabolic disorders like T2D manifest and progress as an inflammatory disorder with severe consequences thereof.



INTRODUCTION

Diabetes mellitus, a life style disease affecting 8.3% of the adult population of the world and increasing at an alarming rate, is one of the most common non-communicable diseases of current era[1]. The burden of this disease is immense owing to transition in lifestyle and dietary habits, ageing of the population and urbanization in the setting of a genetically predisposed environment[2]. The fact that the number of subjects with diabetes mellitus has doubled over the past three decades has made this disease a global challenge[3]. The number of diabetes mellitus patients is projected to increase from 382 million in 2013 to 592 million by 2035, denoting a net increase of 55%[1]. The predominant form is type 2 diabetes (T2D) which accounts for nearly 90% of all diabetes cases.

Diabetes mellitus-not so sweet

T2D is a metabolic disorder characterized by insulin resistance and pancreatic β-cell dysfunction as a consequence of unsettled hyperglycemia[4,5]. In response to nutrient spill over in the setting of insulin resistance and eventual β-cell dysfunction, the general fuel homoeostasis of body is altered[2]. Insulin resistance in target tissues and a relative deficiency of insulin secretion from pancreatic β-cells are the major features of T2D. β-cell hyperplasia and hyperinsulinaemia in response to insulin resistance occur in the preclinical period of disease. Relative insulin deficiency as a consequence of failure of β-cells to compensate for insulin resistance, progresses into overt T2D[6].

Metabolic alterations associated with T2D are well characterised by epidemiological and research based studies. The pathogenesis and progression of T2D is ascribed to four mechanisms; increased advanced glycation end product (AGE) formation, increased polyol pathway flux, activation of protein kinase C (PKC) isoforms, and increased hexosamine pathway flux[7]. Till recently no common linking element was apparent for these mechanisms: however, recently production of superoxide emerged as a unifying mechanism for these four pathways. Downstream to oxidative stress, activation of inflammatory pathways has emerged as an imperative link between T2D and inflammation. Since, abundant data have elucidated the role of oxidative stress in T2D pathogenesis. In this review, we will evaluate the inflammatory component of T2D and underscore the link between metabolic alterations in T2D and inflammation.

T2D AS AN INFLAMMATORY CONDITION

Studies investigating the relation between inflammation and T2D have coalesced sufficient data implicating the role of inflammation towards the development of insulin resistance and pathogenesis of T2D[8,9]. Metabolism and immune system were conventionally regarded as two distinctive mechanisms governing nutrient disposal and body defense, respectively. Typically, little was known about the coordination and interplay between these two systems. However, present research has led to combining these distinct entities as studies perceive pathological activation of the immune system as a regulatory mechanism associated with multiple disorders underlying the metabolic syndrome[10]. Potency of steroid hormones as immune suppressors and hyperglycemic inductors, metabolic alterations associated with pyrexia, wasting syndrome initiated by chronic infections and of late, markers of acute-phase response have been associated with insulin resistance, insulin secretion defects, T2D and vascular complications of T2D[8,11-15].

T2D encompasses colossal cellular factors characteristic of triggering inflammatory signalling cascades. A detailed analysis of these molecules cannot be underscored in this review, however their particular roles in T2D has been outlined in Table 1. Consequently, T2D at the moment is considered an inflammatory disorder triggered by disordered metabolism[16]. The probable history of diabetes involves a more or less latent prodromal period followed by progressive deterioration of glucose tolerance culminating into explicit disease. Progression of islet β-cell failure results in hypertrophy of pancreatic islets and proliferation of β-cells. This phase is associated with an inflammatory response precipitating into reduction of cells by apoptosis and fibrosis of islets. In fact, an analogy of sequence of events involving an incipient inflammatory phase is associated with other T2D complications also[17]. Hyperglycemia is regarded as the major upstream mechanism, and micro-inflammation is regarded as the subsequent downstream driving force of diabetes related complications[17]. Epidemiological data advocate that markers of inflammation are predictive of T2D[18]. The role of inflammation in insulin resistance is traced by the integration of metabolism and innate immunity via nutrient-sensing pathways mutual to pathogen-sensing pathways. Components of nutrition (free fatty acids, glucose, and amino acids) signal through collective receptors and pathways in a similar way as pathogens and/or cytokines. Cells of the immune system (macrophages) and metabolism (adipocytes) also share many functions like secretion of cytokines, and trans-differentiation into macrophages. Nutrients can activate macrophages and adipocytes through common receptors, such as toll-like receptors (TLRs) that sense broad classes of molecular structures common to pathogen groups, and are central to innate immunity and inflammation.

Table 1 Role of various inflammatory molecules in type 2 diabetes.
CategoryMoleculeRole
Pro-inflammatory cytokines and signaling moleculesTNF-αIncreased levels related to IR and T2D Reduces insulin sensitivity by influencing the phosphorylation state of the insulin receptor
IL-6Major pro-inflammatory cytokine that induces inflammation and IR leading to T2D
CRPElevated serum CRP associated with the incidence of T2D
IL-1Associated with obesity and IR Affects insulin signaling directly through the induction of SOCS-3
IL-8Leads to IR via the inhibition of insulin-induced Akt phosphorylation in adipocytes
IL-1βMediates auto-inflammatory process resulting in β-cell death
Transcription factorsNF-κBIncrease the expression of genes encoding cytokines, chemokines, transcription factors and various receptors involved in IR and pathogenesis of T2D
JNKPromotes IR through phosphorylation of serine residues in IRS-1
IKKβLeads to IR through transcriptional activation of NF-κB
AdipokinesLeptinHigh leptin levels, reflecting leptin resistance predict increased risk of T2D
AdiponectinLow levels of this protective adipokine correlate with T2D. Adiponectin is downregulated by TNF-α
ResistinPromotes IR and decreases insulin-stimulated glucose transporters in adipose tissue
AdipsinRole in maintaining β cell function Lower levels of adipsin found in T2D patient
VisfatinVisfatin binds to the insulin receptor at a site distinct from that of insulin and causes hypoglycaemia by reducing glucose release from liver cells and stimulating glucose utilization in adipocytes and myocytes
ChemokinesMCP-1MCP-1 expression in adipose tissue contributes to the macrophage infiltration into this tissue, IR and T2D
IP-10/CXCL10Downstream effector of pro-inflammatory cytokines involved in T2D-related complications
CCR2Imitates tissue inflammation and IR
Toll like receptorTLR2 and TLR4TLR2 and TLR4 play a critical role in the pathogenesis of IR and T2D
Adhesion moleculesE-slectin/P-slectinLead to leukocyte recruitment in local tissue and contributes to inflammation, IR and T2D
ICAM-1/VCAM-1Alters endothelial and sub-endothelial structure leading to reduced vascular permeability, reduced insulin delivery to peripheral insulin sensitive tissues and ultimately T2D
Nuclear receptorsPPARα, PPARγ, and PPARβ/δMutations in PPAR genes associated with IR and T2D
VDRRegulates expression of insulin receptor preferentially by binding as a heterodimer with the RXR to VDREs in the promoter regions of insulin receptor gene
ADIPOSE TISSUE AS A SITE OF INFLAMMATION

Clinical and experimental studies show that adipose tissue acts as a site of inflammation. The first insight came from the study on adipose tissue of obese mice exhibiting elevated production of TNF-α[11]. Consequently, increase in adiposity is associated with upregulation of genes encoding pro-inflammatory molecules and associated with accumulation of immune cells[19-21]. Adipocytes hoard excessive nutrient load and become hypertrophic gradually. Events initiating a pro-inflammatory response involve synergistic contributions of various mechanisms like an increase in nuclear factor κB (NF-κB) and c-Jun NH2-terminal kinase (JNK) activity by hypertrophied adipocytes, endoplasmic reticulum (ER) stress causing altered unfolded protein response (UPR), hypoxic stress in adipose tissue, activation of TLR by excess free fatty acids (FFAs), or increased chylomicron-mediated transport from the gut lumen into the circulation in a lipid-rich diet[16,22,23]. Stressed adipocytes produce various cytokines and chemokines promoting immune-cell activation and accumulation in adipose tissue[24]. A pro-inflammatory loop is formed by several macrophages by clustering around adipocytes, particularly with dead adipocytes forming crown-like structures[19,21,25]. Sustained accumulation of lipids in adipose tissues results in switching of macrophages from an anti-inflammatory “M2” (alternatively activated) to a pro-inflammatory “M1” (classically activated) phenotype[19,21,26,27]. The skew in balance results in an increased secretion of inflammatory molecules that subsequently stimulate the hypertrophied adipocytes resulting into a pro-inflammatory response[28]. The inflammatory response in macrophages is induced by adipocyte-derived FFAs via TLR or NOD-like receptor family, the pyrin domain containing 3 (NLRP3) dependent pathways[29,30]. Local hypoxia as a result of vasculature insufficiency in hypertrophied adipocytes has been proposed to stimulate expression of inflammatory genes in adipocytes as well as immune cells[31]. However, the hypothesis lacks confirmation in the situation of human obesity[32]. Instead, mechanisms like ER stress and autophagocytosis have been proposed as origin of local inflammatory signalling pathways in adipose tissue[22,33]. Recently, the role of the incretin hormone glucose-dependent insulinotropic peptide has also been implicated[34,35]. In addition to adipose tissue, a pro-inflammatory state in liver and skeletal muscle result in disruption of systemic insulin sensitivity and glucose homeostasis that are characteristic of T2D[36-38].

Metabolic inflammation is regulated by critical orchestration of innate and adaptive immune cell interactions[39,40]. Studies investigating immuno-metabolism have recognised that the inflammatory status of immune cells is dictated by their metabolic programming, mitigating the progression of T2D. T2D is preceded by an extensive period of disease development, and inflammation has been shown to be a precipitating factor underpinning insulin resistance, preceding T2D[41,42]. The progression of T2D involves an intricate interplay between metabolism and immunity. The progression of T2D has been causally linked to various types of immune cells but the primary sources of inflammatory effectors contributing to insulin resistance are macrophages[43-45]. Among various cell types, pre-adipocytes, adipocytes, T cells, dendritic cells and macrophages are major cell types involved in obesity-induced inflammation and insulin resistance[46]. Their prime functions are shown in Figure 1. The key inducers of cytokine release in metabolic organs leading to impaired insulin action are tissue-resident macrophages[47].

Figure 1
Figure 1 Functions of various immune cell types in pathogenesis of type 2 diabetes. IL: Interleukin; MCP-1: Monocyte chemoattractant protein-1; TNF-α: Tumor necrosis factor α.

Nutrient overload corresponds to increased infiltration of macrophages in metabolic tissues promoting a pro-inflammatory environment characterised by augmented TNF-α, IL-1β and inducible nitric oxide synthase (iNOS) levels. The accrual of these pro-inflammatory macrophages in metabolic organs like liver, adipose tissue and muscle directly supresses insulin action, thereby promoting hyperglycemia[48].

ROLE OF INFLAMMATION IN INSULIN RESISTANCE

Insulin is a key endocrine hormone produced by β-cells of pancreatic islets. Insulin is regarded as “hormone of abundance” owing to the array of functions it performs, the effects of which extend from metabolic to mitogenic activity (Figure 2). It is likely that disruption of insulin-mediated pathways will have pleiotropic effects that are not confined to carbohydrate metabolism only. Various mechanism working separately or in synergy have been linked to the development of insulin resistance among which chronic inflammation represents as a triggering point[8].

Figure 2
Figure 2 Various hormone functions of insulin.

Inflammation is an important component linking insulin resistance with nutrient overload and increased visceral adipocyte mass[42]. During an insulin-sensitive state, the signalling cascade of insulin upon binding to its receptor results in phosphorylation of tyrosine residues of the insulin receptor substrate 1 (IRS-1) ensuing in downstream insulin signalling[49]. However, in an insulin-resistance state, pro-inflammatory molecules activate various other serine kinases like JNK, inhibitor of NFκB kinase subunit β (IKK-β), extracellular-signal regulated kinase (ERK), ribosomal protein S6 kinase (S6K), mammalian target of rapamycin (mTOR), PKC and glycogen synthase kinase 3β[50]. The activation of these kinases inhibits insulin action by phosphorylating serine residues instead of tyrosine residues in the insulin signalling pathway[49].

The development of insulin resistance is linked to two prime transcription factor-sinalling pathways: JNK and IKKβ/NF-κB[51]. Activation of these two pathways involves a series of proinflammatory stimuli, many of which comprise of both activators and upregulators of NF-κB. In addition, these pathways are also activated by pattern recognition receptors like TLRs and receptors for advanced glycation end products (RAGE). Elevated levels of FFAs result in an increase in diacylglycerol (DAG) that activates PKC isoforms leading to concomitant activation of JNK and NF-κB pathways[52]. Further stimuli involve production of reactive oxygen species (ROS), ER stress and changes in adiposity[53-55].

The mechanisms in development of inflammation-induced insulin resistance are different for JNK and IKKβ. Unlike JNK that phosphorylates the serine residues of IRS-1, IKKB induces insulin resistance by transcriptional activation of NF-κB[56-59]. The physiological substrates of IKKβ are IκB protein inhibitors of NF-κB. IKKβ phosphorylation promotes proteosomal degradation of IκBα liberating NF-κB for nuclear translocation where it stimulates the expression of several target genes (Figure 3)[9]. The products of these target genes of NF-κB induce insulin resistance. The production of inflammatory molecules further activates JNK and NF-κB pathways promoting a vicious loop of insulin resistance by feed-forward mechanism.

Figure 3
Figure 3 Target genes activated by NF-κB. TNF-α: Tumor necrosis factor-alpha; IFN-γ: Interferon-gamma; IL: Interleukin; TGF-β: Tumor growth factor-beta; MCP-1: Monocyte chemoattractant protein-1; MIP: Major intrinsic protein; TNFR: Tumor necrosis factor receptor; INFR: Interferon receptor; IL-R: Interleukin receptor; CD: Cluster of differentiation; ICAM: Intracellular cell adhesion molecule; VCAM: Vascular cell adhesion molecule; CCR: Chemokine CC receptor; TLR: Toll-like receptor; Lox: Lysyl oxidase; RAGE: Receptor advanced glycation end product; PAI: Plasminogen inhibitor activator; SAA: Serum amyloid; CRP: C-reactive protein; COX: Cyclo-oxygenase; iNOS: Inducible nitric oxide synthase; VEGF: Vascular endothelial growth factor; IGFBPs: Insulin-like growth factor binding protein; MnSOD: Manganese superoxide dismutase; RelA: Reticuloendotheliosis viral oncogene homolog A; NF-κB: Nuclear factor-kappa B; IKK: Inhibitor Kappa B kinase; IκBα: Inhibitor of NF-κB; TNFAIP3: TNF-α induced protein 3.
PANCREATIC ISLET INFLAMMATION IN T2D

Increasing evidence suggests the presence of an inflammatory milieu in pancreatic islets in T2D, such as increased cytokine levels, chemokine levels and immune cell infiltration. Evidence of islet inflammation was initially observed in hyperglycemia induced β-cell apoptosis[60]. Recent studies on human islets and monocytes have shown that the combination of hyperglycemia and elevated FFAs induces a more efficient pro-inflammatory phenotype[61,62]. Various T2D experimental animal models like db/db mice and Goto-Kakisaki rats showed increased infiltration by immune cells in the pancreatic islets[63]. Studies on experimental animal models elucidated islet inflammation and macrophage infiltration as an event occurring as early as eight weeks before the onset of frank diabetes[63]. Recruitment of macrophages is a consequence of phagocytic clearance owing to the death of islet β-cells[64]. Alternately, in a diabetic milieu endocrine cell-derived inflammatory molecules like IL-6 and IL-8 produced in islets are also attributed to increased macrophage infiltration[63]. Production of pro-inflammatory cytokines and secretion of chemokines by β-cells results in a vicious cycle speeding up islet inflammation. In humans, IL-1β secreted by infiltrating immune cells is related to the pathogenic process of T2D, as blockade of IL-1 has been associated with reduced hyperglycemia, improved β-cell function and reduced expression of inflammatory markers[65]. However, recent studies involving human islets have shown that induction of IL-1β plays a role in precipitating the clinical features of diabetes and is unlikely involved in initial pathogenesis[66-68]. The first study demonstrating the hyperglycemia-induced IL-1β secretion documented a pro-inflammatory response induced by a non-autoimmune mechanism in β-cells[12]. Ex vivo experiments on isolated human islets exposed to high glucose levels showed increased IL-1β production preceding activation of NF-κB, upregulation of Fas, fragmentation of DNA, and reduction of insulin secretion[69]. Upregulation of IL-1β plays a predominant role as a major cytokine regulating other chemokines and cytokines in islets of T2D patients[12,66,70,71]. This master cytokine elicits a broader response by recruitment of various immune cells and also by induction of IL-1β in β-cells, provoking a vicious inflammatory cycle[66]. The critical role of IL-1β in islet inflammation was recently confirmed by analysing global gene expression in pancreatic islets of humans that showed an association of a group of co-expressed modules enriched for IL-1 related genes with T2D and insulin resistance[72]. SFRP4 gene encoding the secreted frizzled-related protein 4 was one of the interesting genes that were overexpressed, likely mediating the effect of IL-1β on islets[72]. In islets of both T2D subjects as well as in animal models, an eminent number of immune cells along with cytokines and chemokines has been observed[63,66,73]. In fact, T2D animal models invariably exhibit islet immune cell infiltration[63,71].

ISLET INFLAMMATION AND β-CELL DEATH

Islet tissue sections of T2D subjects show well-defined fibrosis which is a hallmark of the late stage of a chronic inflammatory process. In clinically overt T2D subjects a decreased β-cell mass has been reported indicating a probable role in its pathogenesis[4,74]. Decreased β-cell mass in T2D has been attributed to pancreatic β-cell apoptosis and to β-cell dedifferentiation[75]. In slowly progressing T2D, the probability of detecting β-cell damage in pancreatic sections is low, thus very few studies on this aspect have been reported[4,76]. Several mechanisms like amyloid deposition in islets, presence of long-chain FFAs[77], and chronic hyperglycemia[60] has been implicated in β-cell apoptosis. Sustained gluco-lipotoxic conditions amplifies the β-cell stress responses by potentiating effects of elevated levels of FFAs, glucose causing ER stress and mTORC1 activation[78,79,54]. The underlying mechanism for hyperglycemia-induced β-cell apoptosis is attributed to the glucose-induced IL-1β production that upregulates the Fas receptor[80,81,12]. FFAs act as important effector molecules causing β-cell dysfunction by lipoapoptosis (a metabolic cause of programmed cell death). The most abundant saturated FFA in blood is palmitate that has direct lipotoxic effects on β-cells by inducing ER stress and ROS[82-85]. Ceramide, an effector molecule responsible for inducing lipoapoptosis of β-cells, is a metabolic product of FFAs that activates JNK[86-88]. Likewise, incomplete β-cell oxidation of fatty acids resulting in metabolites like DAG and triglycerides (TGs) also elicits final effector molecules contributing to FFA-induced lipotoxicity as well as insulin resistance[89-91]. In addition to this, FFA-induced activation of JNK by Src has also been reported in a recent study[92]. These studies show that islet inflammation contributes to β-cell dysfunction.

TRIGGERING OF THE INNATE IMMUNE SYSTEM IN T2D

Nutrient excess in metabolic tissues resulting in metabolic inflammation, i.e., a low-level pro-inflammatory milieu, has emerged as an important factor underlying the development of T2D[11-15,93,94]. Activation of innate immunity in T2D is linked to the activation of TLRs. These receptors have been implicated in diabetes-induced inflammation and vascular complications[95]. TLRs comprise the pattern-recognition receptors characteristic of the innate immune system. Various pathogen-associated molecular patterns (PAMPs) encompassing carbohydrates, proteins, nucleic acids and lipids, are recognised by TLRs followed by initiation of an immune response. TLR2, a receptor for pathogen lipoproteins and TLR4, a receptor of lipopolysaccharides, are activated by FFAs[96,29]. Binding of FFAs to TLRs has been postulated to directly induce a pro-inflammatory response[97,98]. Also, various indirect ways of TLR activation by FFAs has been postulated recently[99]. In vitro studies have demonstrated that, unlike the short chain FFAs, the long chain palmitate and oleate that comprise 80% of circulating FFAs are pro-inflammatory in various cell types[29,96,98,100,101]. Contemporary studies report the activation of TLR signalling by FFA-induced formation of lipid rafts that favour TLR dimerization in cell membranes[92,102]. Recently, fatty acid transporter CD36 binding to TLR2 and liver-derived glycoprotein fetuin-A binding to TLR4 were identified as endogenous ligands linking FFAs to TLRs, eliciting inflammation and prompting insulin resistance[103,104]. In addition, damage-associated molecular patterns (DAMPs) like high-mobility group box 1 (HMGB1) and AGEs also act as endogenous ligands which are recognised by TLRs, thereby activating pro-inflammatory pathways[105]. TLR2 is responsible for upregulation of inflammatory molecules like NF-κB, myeloid differentiating factor 88 (MyD88) and chemokine (C-C motif) ligand 2 (CCL2)[106]. TLR4 knockout mice have been shown to be protected from insulin resistance as well as from fat-induced inflammation[106]. TLR4 silencing by siRNA technology has been shown to attenuate the hyperglycemia-induced activation of IκB/NF-κB[107]. TLR5 is a receptor for bacterial flagellin that controls metabolic pathways through sensing gut microbiota. TLR5 knockout mice have been reported to exhibit increased adiposity along with hyperphagia, hypertension, hyperlipidemia and insulin resistance[108]. Activation of inflammatory pathways in a TLR-independent mechanism by metabolic stress involves generation of ROS that induce stress kinases and NLRP3 inflammasome (multiprotein complexes responsible for production of bioactive IL-1β) formation[109].

Both TLR-dependent and TLR-independent mechanisms function in concert. This finding is demonstrated by animal models of diabetes in which there is partly protection of pro-inflammatory cytokine production in case of deficiency of TLR2 or TLR4, whereas deficiency of a universal intracellular docking protein MyD88 required for TLR signalling, exerted total protection[61]. Apart from FFAs, systemic inflammatory responses are also elicited by elevated glucose levels[110]. Sustained hyperglycemia results in non-enzymatic glycation of lipids and proteins resulting in the formation of AGEs. AGEs stimulate the pattern recognition receptor RAGE. Numerous cell types, like macrophages, T cells, smooth muscle cells, neuronal cells, podocytes and cardiomyocytes, express RAGE[111]. RAGE activates the pleiotropic pro-inflammatory transcription factor NF-κB along with stress kinases ERK1 and ERK2[112]. Excessive glucose metabolized by oxidative phosphorylation to ATP results in ROS generation that tends to activate the NLRP3 inflammasome concomitantly with FFAs[67]. This results in release of active IL-1β along with IL-1-dependent cytokine and chemokine production[61].

FROM INNATE TO ADAPTIVE IMMUNITY IN T2DM

The role of specific or adaptive immunity comes from the recent clinical overlap between type 1 diabetes (T1D) and T2D such as younger age of onset in T2D and increasing body mass index (BMI) coinciding with increased incidence in T1D. Moreover, progressive decrease in β-cell mass observed in T2D and evidence of insulin resistance in T1D has blurred the etiology[113]. The argument supporting the involvement of autoimmunity in islets of T2D patients is evident from the presence of β-cell specific antibodies in nearly 10% of T2D patients and presence T cells reactive to β cell antigens in some patients[114]. The number of autoantigen-responsive T lymphocytes in islets from T2D patients has been reported to correlate with disease progression[114], however the exact role of islet autoimmunity in T2D requires further studies. A monogenic form of diabetes characterised by typical features of T1D like lean body mass, young age of onset, autoantibodies to β-cells, rapid disappearance of C-peptide and insulin requirement concomitantly with T2D-associated insulin resistance provides genetic support for the overlap between T1D and T2D[115]. The genetic alteration is attributed to an autosomal-dominant mutation in the SIRT1 gene, and the pathogenesis involves β-cell impairment and death, paralleling a state of activation of immune system[115]. As a consequence of insulin resistance, stress induced β-cell death results in the release of autoantigens along with alarmins (endogenous molecules released by necrotic cells causing activation of immune system). Alarmins have potentiating effects of promoting pathologic self-antigen presentation, resulting in enhanced adaptive immune response[116]. In light of these observations, sirtuins are recognised as novel regulators of immuno-metabolism in humans. Apart from SIRT1, SIRT2 has been recently linked to cytoskeleton remodeling and activation of NLRP3 in intracellular pathways[117]. Apart from the activation of innate immunity, the contribution of adaptive immune cells in inducing inflammation is now established in T2D at the cellular level.

Experimental animal models of insulin resistance have demonstrated a Th2/Th1 shift in favour of Th1, shifting the Treg/Th17 shift towards Th17 and shifting the CD8/CD4 ratio in favour of CD8 and finally reduction of T-cell receptor (TCR) diversity[118-121]. These studies have recently been extrapolated to human subjects[122] and confirm the observation that an increase pro-inflammatory stimuli (IFN-γ) causing M1 phenotype switching of adipose tissue macrophages result in the activation of a Th1 type response[121]. IFN-γ and IL-17 produced by these T cell populations interact directly with adipocytes in addition to contributing to a pro-inflammatory loop in cells of innate immunity. IFN-γ inhibits the JAK-STAT pathway, and IL-17 induces the secretion of IL-6 from adipocytes[123]. β-cells isolated from T2D patients exhibit increased IL-8 and decreased IL-10 secretion[124]. Recent studies regard the contribution of B-cell humoral immunity in adipose tissue inflammation. A study on experimental mice models involving B-cell knockout mice and anti-CD20 therapy showed a significantly improved metabolic phenotype and adipose tissue inflammation[120].

LINK BETWEEN ER STRESS AND INFLAMMATION IN T2D

Activation of ER stress and the UPR forms a convincing hypothesis for the induction of inflammatory pathways in T2D. ER stress in T2D occurs by virtue of nutrient overload, hypoxia and accumulation of unfolded proteins in metabolic organs[22]. Under normal conditions, the flux of proteins through ER is high, and in the setting of insulin resistance or glucotoxicity, a prolonged state of insulin need generates ER stress[125].

Three ER localized sensors control the activation of ER stress and UPR (Figure 4): (1) the double-stranded RNA-activated protein kinase (PKR)-like ER kinase (PERK); (2) inositol-requiring kinase 1 (IRE1); and (3) activating transcription factor 6 (ATF6). ER stress by protein overload or accumulation of unfolded proteins causes dissociation of GRP78, and the subsequent binding to unfolded proteins in ER prevents their transport to cis Golgi.

Figure 4
Figure 4 Mechanism of endoplasmic reticulum stress. ER: Endoplasmic-reticulum; ATF6: Activating transcription factor 6; PERK: Double-stranded RNA-activated protein kinase (PKR)-like ER kinase; IRE1: Inositol-requiring kinase 1; IL: Interleukin; MCP: Monocyte chemoattractant protein; TNF-α: Tumor necrosis factor α; JNK: c-Jun NH2-terminal kinase; NF-κB: Nuclear factor κB.

Prominently, UPR activation stimulates inflammatory stress kinases like JNK and IKK and their critical downstream transcriptional targets; activator protein 1 (AP-1) and NF-κB, respectively[126,127].

These transcription factors control the induction of inflammatory cytokines and chemoattractants that are known to have a direct link with the development of insulin resistance[128,129]. ER stress can also impair insulin signalling by activation of stress kinases (JNK, IKK) that can inhibit insulin receptor substrates by direct phosphorylation. Recently, death protein 5 (DP5) and p53-upregulated modulator of apoptosis (PUMA) have been reported as inducers of β-cell apoptosis by mediating ER stress[130]. ER stress can also cause induction of lipogenic genes that promote lipid accumulation and thereby contributes to the development of lipid-induced insulin resistance[131].

ER stress and UPR pathways

Triggering of inflammatory signals by three pathways of UPR is initiated by activation of JNKs and NF-κB in B cells. This activation acts as the linkage point between metabolic and immune pathways since the activation of these very kinases is analogous to that elicited by an immune response[94,132]. JNKs play an important role in T2D, as increased activity has been shown to promote insulin resistance [56,133].

The first responses for opposing ER stress involve decreasing the translation of proteins. This involves phosphorylation of α subunit of eIF2 by PERK. In humans and mice, loss of PERK expression is linked to dysregulation of the UPR response which is fundamental to ER stress, resulting in increased cell death and T2D[134]. A permanent form of neonatal diabetes in humans is related to elevated ER stress markers as a result of a mutation in PERK, confirming the pivotal role of PERK in regulating ER stress during fetal development[135-137].

A factor in the second pathway of UPR, IRE1, is a prime regulator of ER stress and is highly expressed in the pancreas. An in vitro knockdown study on IRE1 signalling showed a decreased synthesis of insulin[138,139]. Upon activation, IRE1 initiates activation of X-box binding protein 1 (XBP1) that leads to upregulation of ER expansion and biogenesis[140]. The critical role of XBP1 in achieving an optimal insulin secretion and glucose control was demonstrated in β-cell-specific XBP1-deficient mice that exhibited impaired pro-insulin processing and secretion, reduced β-cell proliferation and hyperactivation of IRE1[141].

The third pathway of UPR involves the activation of ATF6, the basic leucine zipper domain protein, that upregulates PERK1 and IRE1 pathways by suppressing the apoptotic UPR signalling cascade under chronic ER stress. The role of ATF6 activation in β-cell dysfunction has been concluded in studies that showed decreased expression of insulin gene by ER stress-induced ATF6 activation and a decrease in ER chaperones along with induction of apoptosis in ATF6 knockdown insulinoma cells[142,143].

ACTIVATION OF INFLAMMASOME IN T2D

Inflammasomes are multiprotein complexes in the intracellular machinery responsible for production of bioactive IL-1β in response to multiple stimuli[144]. NLRP is a subfamily of Nod-like receptors containing a central nucleotide binding and oligomerization (NACHT) domain with flanking C-terminal leucine-rich repeats (LRRs) and N-terminal caspase recruitment (CARD) or pyrin (PYD) domains[145]. The NOD-like receptor family, the pyrin domain containing 3 (NLRP3) inflammasome is in a pathway that controls the production of IL-1β and IL-18[146-148]. Unlike TLR, a potential role of NLR in metabolic abnormalities has not been extensively investigated. NLRP forms a constituent of the inflammasomes responsible for maturation and release of IL-1β, and thus is a relevant candidate for metabolic disorders and T2D[149]. NLRP3-dependent activation of inflammaosomes in diabetes was proposed by studies implicating the release of IL-1β as a consequence of elevated levels of glucose, FFAs and human islet amylopeptide (hIAPP)[16,150,151]. However, the effective metabolites involved in activation of inflammasomes are not clearly elucidated yet (Figure 5).

Figure 5
Figure 5 Activation of inflammasomes in type 2 diabetes (metabolic stress activates multiprotein complex, inflammasome in β-cells that induce caspase-1 to cleave pro-interleukin-1β (pro-IL-1β) into active IL-1β. β-cell-derived IL-1β promote the release of chemokines and recruitment of macrophages that are activated by human islet amyloid polypeptide, leading to deleterious concentrations of IL-1β. FFA: Free fatty acid; IL: Interleukin.

The NLRP3 inflammasome is a general metabolic alarmin stimulated by different endogenous and exogenous stimuli[152]. NLRP3 inflammasome activation is augmented in T2D patients[153]. Dysregulation of lipid metabolism, paving the way to aberrant lipid accumulation, as well as formation of oxidized LDL and cholesterol, triggers NLRP3 activation[30,153,154] similar to ER stress that acts as one of the important factors triggering NLRP3 activation[155,156]. In T2D subjects, increased oxidative stress also contributes to NLRP3 inflammasome activation[157,158].

Studies on obesity-induced inflammation and insulin resistance are also indicative of the role of NLRP3. In experimental models of calorie-restricted mice, a positive correlation has been observed between IL-1β/ NLRP3 mRNA and body weight[30] whereas disruption of NLRP3 gene in obese mice has revealed changes in metabolic profiles. Insulin resistance as a consequence of inflammasome activation is directly related to FFAs and LPS[109]. Apart from Insulin resistance, activation of inflammasomes is related to β-cell dysfunction, as NLRP3-knockout mice exhibit improved glycemic profiles after consumption of a high-fat diet, likely due to attenuation of IL-1β[67]. In response to hyperglycemia-induced increased production of ROS, NLRP3 activation occurs as a result of dissociation of thioredoxin interacting protein (TXNIP) from thioredoxin and its subsequent binding to NLRP3[67]. Nevertheless, shortage of TXNIP has shown effects on glucose metabolism in addition to the NLRP3 activation[159]. A substantial role of inflammasome activation in β-cell dysfunction was recently reported by ablation of NLRP3 that conferred protection to β-cell function and structure from injury inflicted by metabolic stress[160].

Secretion of IL-1β requires two induction stimuli; the first stimulus induces pro-IL-1β expression and the second inflammasome activation. Inflammasome activation triggers caspase-1 resulting in cleavage of pro-IL-1β and release of mature IL-1β. In T2D, the first stimulus comes from minimally-modified LDL in islets which prime the macrophages for processing of IL-1β by activation of TLR4 signalling. Recently, the second stimulus was recognized to regard islet hIAPP, secreted by β-cells in response to high glucose levels[151]. hIAPP was shown to direct NLRP3 activation by inducing β-cell injury. In islets, interaction of macrophages and β-cells is essential for the activation of inflammasomes. hIAPP, a soluble oligomer induces activation of NLRP3 and subsequent release of IL-1β from macrophages and dendritic cells which are primed with TLR4 agonists like LPS or modified LDL molecules[151]. The macrophages are attracted to islets by hIAPP-induced synthesis of chemokines (CCL2 and CXCL1). It has been reported that overexpression of hIAPP in islet grafts increases the recruitment of macrophages by 50%[161]. Recently the activation of inflammasomes in myeloid cells in T2D patients was elucidated. A study on untreated T2D subjects showed upregulation of IL-1β production and maturation in macrophages[153]. Treatment of macrophages with various alarmins like FFA, hIAPP, HMGB1and ATP resulted in release of inflammasome products. Studies have shown that T2D subjects exhibit elevated levels of circulating alarmin molecules thereby advocating a possible role of these molecules in NLRP3 inflammasome activation in myeloid cells[162].

PERSPECTIVES

The concept of chronic low-level inflammation in T2D has given an impetus to the field of immune-metabolism. Elucidation of various cellular mechanisms linking inflammation to insulin resistance and β-cell dysfunction has revolutionized insights in the molecular pathogenesis of diabetes. Insights into intricate pathways provide a platform to tackle the distinct pathway without compromising immuno-surveillance. Nutritional and therapeutic interventions aimed at controlling/inhibiting the escalating pro-inflammatory response can help in attenuating the pathogenesis and progression of T2D. Well-designed studies should offer the development of novel targeted therapeutics to deal with the disease burden of T2D and its associated complications.

Footnotes

P- Reviewer: Bener A S- Editor: Tian YL L- Editor: A E- Editor: Wu HL

References
1.  International Diabetes Federation IDF Diabetes Atlas. 6th ed. Brussels, Belgium: International Diabetes Federation 2013;  Available from: http://www.idf.org/diabetesatlas.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Nolan CJ, Damm P, Prentki M. Type 2 diabetes across generations: from pathophysiology to prevention and management. Lancet. 2011;378:169-181.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 617]  [Cited by in F6Publishing: 589]  [Article Influence: 45.3]  [Reference Citation Analysis (0)]
3.  Shaw JE, Sicree RA, Zimmet PZ. Global estimates of the prevalence of diabetes for 2010 and 2030. Diabetes Res Clin Pract. 2010;87:4-14.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4438]  [Cited by in F6Publishing: 4217]  [Article Influence: 301.2]  [Reference Citation Analysis (3)]
4.  Butler AE, Janson J, Bonner-Weir S, Ritzel R, Rizza RA, Butler PC. Beta-cell deficit and increased beta-cell apoptosis in humans with type 2 diabetes. Diabetes. 2003;52:102-110.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3031]  [Cited by in F6Publishing: 2912]  [Article Influence: 138.7]  [Reference Citation Analysis (0)]
5.  Ashcroft FM, Rorsman P. Diabetes mellitus and the β cell: the last ten years. Cell. 2012;148:1160-1171.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 624]  [Cited by in F6Publishing: 649]  [Article Influence: 54.1]  [Reference Citation Analysis (0)]
6.  Quan W, Jo EK, Lee MS. Role of pancreatic β-cell death and inflammation in diabetes. Diabetes Obes Metab. 2013;15 Suppl 3:141-151.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 61]  [Cited by in F6Publishing: 64]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
7.  Brownlee M. Biochemistry and molecular cell biology of diabetic complications. Nature. 2001;414:813-820.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6145]  [Cited by in F6Publishing: 5933]  [Article Influence: 258.0]  [Reference Citation Analysis (0)]
8.  Pickup JC, Crook MA. Is type II diabetes mellitus a disease of the innate immune system? Diabetologia. 1998;41:1241-1248.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 695]  [Cited by in F6Publishing: 727]  [Article Influence: 28.0]  [Reference Citation Analysis (0)]
9.  Shoelson SE, Lee J, Goldfine AB. Inflammation and insulin resistance. J Clin Invest. 2006;116:1793-1801.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2726]  [Cited by in F6Publishing: 2865]  [Article Influence: 159.2]  [Reference Citation Analysis (0)]
10.  Donath MY, Dalmas É, Sauter NS, Böni-Schnetzler M. Inflammation in obesity and diabetes: islet dysfunction and therapeutic opportunity. Cell Metab. 2013;17:860-872.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 247]  [Cited by in F6Publishing: 240]  [Article Influence: 21.8]  [Reference Citation Analysis (0)]
11.  Hotamisligil GS, Shargill NS, Spiegelman BM. Adipose expression of tumor necrosis factor-alpha: direct role in obesity-linked insulin resistance. Science. 1993;259:87-91.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5334]  [Cited by in F6Publishing: 5219]  [Article Influence: 168.4]  [Reference Citation Analysis (0)]
12.  Maedler K, Sergeev P, Ris F, Oberholzer J, Joller-Jemelka HI, Spinas GA, Kaiser N, Halban PA, Donath MY. Glucose-induced beta cell production of IL-1beta contributes to glucotoxicity in human pancreatic islets. J Clin Invest. 2002;110:851-860.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 863]  [Cited by in F6Publishing: 844]  [Article Influence: 38.4]  [Reference Citation Analysis (0)]
13.  Berk BC, Weintraub WS, Alexander RW. Elevation of C-reactive protein in “active” coronary artery disease. Am J Cardiol. 1990;65:168-172.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Ridker PM, Cushman M, Stampfer MJ, Tracy RP, Hennekens CH. Inflammation, aspirin, and the risk of cardiovascular disease in apparently healthy men. N Engl J Med. 1997;336:973-979.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3798]  [Cited by in F6Publishing: 3698]  [Article Influence: 137.0]  [Reference Citation Analysis (1)]
15.  Navarro-González JF, Mora-Fernández C. The role of inflammatory cytokines in diabetic nephropathy. J Am Soc Nephrol. 2008;19:433-442.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 592]  [Cited by in F6Publishing: 632]  [Article Influence: 39.5]  [Reference Citation Analysis (0)]
16.  Donath MY, Shoelson SE. Type 2 diabetes as an inflammatory disease. Nat Rev Immunol. 2011;11:98-107.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2178]  [Cited by in F6Publishing: 2347]  [Article Influence: 180.5]  [Reference Citation Analysis (0)]
17.  Wada J, Makino H. Inflammation and the pathogenesis of diabetic nephropathy. Clin Sci (Lond). 2013;124:139-152.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 447]  [Cited by in F6Publishing: 519]  [Article Influence: 47.2]  [Reference Citation Analysis (0)]
18.  Schmidt MI, Duncan BB, Sharrett AR, Lindberg G, Savage PJ, Offenbacher S, Azambuja MI, Tracy RP, Heiss G. Markers of inflammation and prediction of diabetes mellitus in adults (Atherosclerosis Risk in Communities study): a cohort study. Lancet. 1999;353:1649-1652.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 739]  [Cited by in F6Publishing: 723]  [Article Influence: 28.9]  [Reference Citation Analysis (0)]
19.  Weisberg SP, McCann D, Desai M, Rosenbaum M, Leibel RL, Ferrante AW. Obesity is associated with macrophage accumulation in adipose tissue. J Clin Invest. 2003;112:1796-1808.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6517]  [Cited by in F6Publishing: 6528]  [Article Influence: 310.9]  [Reference Citation Analysis (0)]
20.  Wu H, Ghosh S, Perrard XD, Feng L, Garcia GE, Perrard JL, Sweeney JF, Peterson LE, Chan L, Smith CW. T-cell accumulation and regulated on activation, normal T cell expressed and secreted upregulation in adipose tissue in obesity. Circulation. 2007;115:1029-1038.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 479]  [Cited by in F6Publishing: 491]  [Article Influence: 28.9]  [Reference Citation Analysis (0)]
21.  Xu H, Barnes GT, Yang Q, Tan G, Yang D, Chou CJ, Sole J, Nichols A, Ross JS, Tartaglia LA. Chronic inflammation in fat plays a crucial role in the development of obesity-related insulin resistance. J Clin Invest. 2003;112:1821-1830.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4595]  [Cited by in F6Publishing: 4426]  [Article Influence: 221.3]  [Reference Citation Analysis (0)]
22.  Hotamisligil GS. Endoplasmic reticulum stress and the inflammatory basis of metabolic disease. Cell. 2010;140:900-917.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1987]  [Cited by in F6Publishing: 2078]  [Article Influence: 148.4]  [Reference Citation Analysis (0)]
23.  Maury E, Brichard SM. Adipokine dysregulation, adipose tissue inflammation and metabolic syndrome. Mol Cell Endocrinol. 2010;314:1-16.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 713]  [Cited by in F6Publishing: 716]  [Article Influence: 51.1]  [Reference Citation Analysis (0)]
24.  Skurk T, Alberti-Huber C, Herder C, Hauner H. Relationship between adipocyte size and adipokine expression and secretion. J Clin Endocrinol Metab. 2007;92:1023-1033.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 863]  [Cited by in F6Publishing: 839]  [Article Influence: 49.4]  [Reference Citation Analysis (0)]
25.  Cinti S, Mitchell G, Barbatelli G, Murano I, Ceresi E, Faloia E, Wang S, Fortier M, Greenberg AS, Obin MS. Adipocyte death defines macrophage localization and function in adipose tissue of obese mice and humans. J Lipid Res. 2005;46:2347-2355.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1679]  [Cited by in F6Publishing: 1680]  [Article Influence: 88.4]  [Reference Citation Analysis (0)]
26.  Lumeng CN, DelProposto JB, Westcott DJ, Saltiel AR. Phenotypic switching of adipose tissue macrophages with obesity is generated by spatiotemporal differences in macrophage subtypes. Diabetes. 2008;57:3239-3246.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 674]  [Cited by in F6Publishing: 642]  [Article Influence: 40.1]  [Reference Citation Analysis (0)]
27.  Prieur X, Mok CY, Velagapudi VR, Núñez V, Fuentes L, Montaner D, Ishikawa K, Camacho A, Barbarroja N, O’Rahilly S. Differential lipid partitioning between adipocytes and tissue macrophages modulates macrophage lipotoxicity and M2/M1 polarization in obese mice. Diabetes. 2011;60:797-809.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 268]  [Cited by in F6Publishing: 260]  [Article Influence: 20.0]  [Reference Citation Analysis (0)]
28.  Maury E, Noël L, Detry R, Brichard SM. In vitro hyperresponsiveness to tumor necrosis factor-alpha contributes to adipokine dysregulation in omental adipocytes of obese subjects. J Clin Endocrinol Metab. 2009;94:1393-1400.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 57]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
29.  Nguyen MT, Favelyukis S, Nguyen AK, Reichart D, Scott PA, Jenn A, Liu-Bryan R, Glass CK, Neels JG, Olefsky JM. A subpopulation of macrophages infiltrates hypertrophic adipose tissue and is activated by free fatty acids via Toll-like receptors 2 and 4 and JNK-dependent pathways. J Biol Chem. 2007;282:35279-35292.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 730]  [Cited by in F6Publishing: 749]  [Article Influence: 44.1]  [Reference Citation Analysis (0)]
30.  Vandanmagsar B, Youm YH, Ravussin A, Galgani JE, Stadler K, Mynatt RL, Ravussin E, Stephens JM, Dixit VD. The NLRP3 inflammasome instigates obesity-induced inflammation and insulin resistance. Nat Med. 2011;17:179-188.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1698]  [Cited by in F6Publishing: 1841]  [Article Influence: 141.6]  [Reference Citation Analysis (0)]
31.  Ye J. Emerging role of adipose tissue hypoxia in obesity and insulin resistance. Int J Obes (Lond). 2009;33:54-66.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 366]  [Cited by in F6Publishing: 390]  [Article Influence: 24.4]  [Reference Citation Analysis (0)]
32.  Goossens GH, Bizzarri A, Venteclef N, Essers Y, Cleutjens JP, Konings E, Jocken JW, Cajlakovic M, Ribitsch V, Clément K. Increased adipose tissue oxygen tension in obese compared with lean men is accompanied by insulin resistance, impaired adipose tissue capillarization, and inflammation. Circulation. 2011;124:67-76.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 115]  [Reference Citation Analysis (0)]
33.  Martinez J, Verbist K, Wang R, Green DR. The relationship between metabolism and the autophagy machinery during the innate immune response. Cell Metab. 2013;17:895-900.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 52]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
34.  Nie Y, Ma RC, Chan JC, Xu H, Xu G. Glucose-dependent insulinotropic peptide impairs insulin signaling via inducing adipocyte inflammation in glucose-dependent insulinotropic peptide receptor-overexpressing adipocytes. FASEB J. 2012;26:2383-2393.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 39]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
35.  Timper K, Grisouard J, Sauter NS, Herzog-Radimerski T, Dembinski K, Peterli R, Frey DM, Zulewski H, Keller U, Müller B. Glucose-dependent insulinotropic polypeptide induces cytokine expression, lipolysis, and insulin resistance in human adipocytes. Am J Physiol Endocrinol Metab. 2013;304:E1-13.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 55]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
36.  Brun P, Castagliuolo I, Di Leo V, Buda A, Pinzani M, Palù G, Martines D. Increased intestinal permeability in obese mice: new evidence in the pathogenesis of nonalcoholic steatohepatitis. Am J Physiol Gastrointest Liver Physiol. 2007;292:G518-G525.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 592]  [Cited by in F6Publishing: 600]  [Article Influence: 35.3]  [Reference Citation Analysis (0)]
37.  Hijona E, Hijona L, Arenas JI, Bujanda L. Inflammatory mediators of hepatic steatosis. Mediators Inflamm. 2010;2010:837419.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 76]  [Cited by in F6Publishing: 63]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
38.  Varma V, Yao-Borengasser A, Rasouli N, Nolen GT, Phanavanh B, Starks T, Gurley C, Simpson P, McGehee RE, Kern PA. Muscle inflammatory response and insulin resistance: synergistic interaction between macrophages and fatty acids leads to impaired insulin action. Am J Physiol Endocrinol Metab. 2009;296:E1300-E1310.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 158]  [Cited by in F6Publishing: 166]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
39.  Winer S, Winer DA. The adaptive immune system as a fundamental regulator of adipose tissue inflammation and insulin resistance. Immunol Cell Biol. 2012;90:755-762.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 91]  [Cited by in F6Publishing: 96]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
40.  Lumeng CN. Innate immune activation in obesity. Mol Aspects Med. 2013;34:12-29.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 99]  [Cited by in F6Publishing: 110]  [Article Influence: 9.2]  [Reference Citation Analysis (0)]
41.  Hotamisligil GS. Inflammation and metabolic disorders. Nature. 2006;444:860-867.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5683]  [Cited by in F6Publishing: 5905]  [Article Influence: 347.4]  [Reference Citation Analysis (1)]
42.  Hotamisligil GS, Erbay E. Nutrient sensing and inflammation in metabolic diseases. Nat Rev Immunol. 2008;8:923-934.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 710]  [Cited by in F6Publishing: 714]  [Article Influence: 44.6]  [Reference Citation Analysis (0)]
43.  Galic S, Oakhill JS, Steinberg GR. Adipose tissue as an endocrine organ. Mol Cell Endocrinol. 2010;316:129-139.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1129]  [Cited by in F6Publishing: 1096]  [Article Influence: 78.3]  [Reference Citation Analysis (0)]
44.  Bhargava P, Lee CH. Role and function of macrophages in the metabolic syndrome. Biochem J. 2012;442:253-262.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 84]  [Article Influence: 7.0]  [Reference Citation Analysis (0)]
45.  Odegaard JI, Ricardo-Gonzalez RR, Goforth MH, Morel CR, Subramanian V, Mukundan L, Red Eagle A, Vats D, Brombacher F, Ferrante AW. Macrophage-specific PPARgamma controls alternative activation and improves insulin resistance. Nature. 2007;447:1116-1120.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1522]  [Cited by in F6Publishing: 1614]  [Article Influence: 94.9]  [Reference Citation Analysis (0)]
46.  McArdle MA, Finucane OM, Connaughton RM, McMorrow AM, Roche HM. Mechanisms of obesity-induced inflammation and insulin resistance: insights into the emerging role of nutritional strategies. Front Endocrinol (Lausanne). 2013;4:52.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 326]  [Cited by in F6Publishing: 327]  [Article Influence: 29.7]  [Reference Citation Analysis (0)]
47.  Olefsky JM, Glass CK. Macrophages, inflammation, and insulin resistance. Annu Rev Physiol. 2010;72:219-246.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1867]  [Cited by in F6Publishing: 1964]  [Article Influence: 140.3]  [Reference Citation Analysis (1)]
48.  Steinberg GR. Inflammation in obesity is the common link between defects in fatty acid metabolism and insulin resistance. Cell Cycle. 2007;6:888-894.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 100]  [Cited by in F6Publishing: 102]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
49.  Schenk S, Saberi M, Olefsky JM. Insulin sensitivity: modulation by nutrients and inflammation. J Clin Invest. 2008;118:2992-3002.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 843]  [Cited by in F6Publishing: 833]  [Article Influence: 52.1]  [Reference Citation Analysis (0)]
50.  Gual P, Le Marchand-Brustel Y, Tanti JF. Positive and negative regulation of insulin signaling through IRS-1 phosphorylation. Biochimie. 2005;87:99-109.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 605]  [Cited by in F6Publishing: 615]  [Article Influence: 32.4]  [Reference Citation Analysis (0)]
51.  Shoelson SE, Herrero L, Naaz A. Obesity, inflammation, and insulin resistance. Gastroenterology. 2007;132:2169-2180.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1219]  [Cited by in F6Publishing: 1219]  [Article Influence: 71.7]  [Reference Citation Analysis (0)]
52.  Gao Z, Zhang X, Zuberi A, Hwang D, Quon MJ, Lefevre M, Ye J. Inhibition of insulin sensitivity by free fatty acids requires activation of multiple serine kinases in 3T3-L1 adipocytes. Mol Endocrinol. 2004;18:2024-2034.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 224]  [Cited by in F6Publishing: 241]  [Article Influence: 12.1]  [Reference Citation Analysis (0)]
53.  Furukawa S, Fujita T, Shimabukuro M, Iwaki M, Yamada Y, Nakajima Y, Nakayama O, Makishima M, Matsuda M, Shimomura I. Increased oxidative stress in obesity and its impact on metabolic syndrome. J Clin Invest. 2004;114:1752-1761.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3370]  [Cited by in F6Publishing: 3614]  [Article Influence: 190.2]  [Reference Citation Analysis (0)]
54.  Ozcan U, Cao Q, Yilmaz E, Lee AH, Iwakoshi NN, Ozdelen E, Tuncman G, Görgün C, Glimcher LH, Hotamisligil GS. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes. Science. 2004;306:457-461.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2756]  [Cited by in F6Publishing: 2754]  [Article Influence: 137.7]  [Reference Citation Analysis (0)]
55.  Summers SA. Ceramides in insulin resistance and lipotoxicity. Prog Lipid Res. 2006;45:42-72.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 591]  [Cited by in F6Publishing: 584]  [Article Influence: 30.7]  [Reference Citation Analysis (0)]
56.  Hirosumi J, Tuncman G, Chang L, Görgün CZ, Uysal KT, Maeda K, Karin M, Hotamisligil GS. A central role for JNK in obesity and insulin resistance. Nature. 2002;420:333-336.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2445]  [Cited by in F6Publishing: 2381]  [Article Influence: 108.2]  [Reference Citation Analysis (0)]
57.  Aguirre V, Uchida T, Yenush L, Davis R, White MF. The c-Jun NH(2)-terminal kinase promotes insulin resistance during association with insulin receptor substrate-1 and phosphorylation of Ser(307). J Biol Chem. 2000;275:9047-9054.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1069]  [Cited by in F6Publishing: 1064]  [Article Influence: 44.3]  [Reference Citation Analysis (0)]
58.  Aguirre V, Werner ED, Giraud J, Lee YH, Shoelson SE, White MF. Phosphorylation of Ser307 in insulin receptor substrate-1 blocks interactions with the insulin receptor and inhibits insulin action. J Biol Chem. 2002;277:1531-1537.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 706]  [Cited by in F6Publishing: 704]  [Article Influence: 32.0]  [Reference Citation Analysis (0)]
59.  Werner ED, Lee J, Hansen L, Yuan M, Shoelson SE. Insulin resistance due to phosphorylation of insulin receptor substrate-1 at serine 302. J Biol Chem. 2004;279:35298-35305.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 169]  [Cited by in F6Publishing: 174]  [Article Influence: 8.7]  [Reference Citation Analysis (0)]
60.  Donath MY, Gross DJ, Cerasi E, Kaiser N. Hyperglycemia-induced beta-cell apoptosis in pancreatic islets of Psammomys obesus during development of diabetes. Diabetes. 1999;48:738-744.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 289]  [Cited by in F6Publishing: 304]  [Article Influence: 12.2]  [Reference Citation Analysis (0)]
61.  Böni-Schnetzler M, Boller S, Debray S, Bouzakri K, Meier DT, Prazak R, Kerr-Conte J, Pattou F, Ehses JA, Schuit FC. Free fatty acids induce a proinflammatory response in islets via the abundantly expressed interleukin-1 receptor I. Endocrinology. 2009;150:5218-5229.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 245]  [Cited by in F6Publishing: 251]  [Article Influence: 16.7]  [Reference Citation Analysis (0)]
62.  Dasu MR, Jialal I. Free fatty acids in the presence of high glucose amplify monocyte inflammation via Toll-like receptors. Am J Physiol Endocrinol Metab. 2011;300:E145-E154.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 155]  [Cited by in F6Publishing: 157]  [Article Influence: 12.1]  [Reference Citation Analysis (0)]
63.  Ehses JA, Perren A, Eppler E, Ribaux P, Pospisilik JA, Maor-Cahn R, Gueripel X, Ellingsgaard H, Schneider MK, Biollaz G. Increased number of islet-associated macrophages in type 2 diabetes. Diabetes. 2007;56:2356-2370.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 552]  [Cited by in F6Publishing: 548]  [Article Influence: 32.2]  [Reference Citation Analysis (0)]
64.  Cnop M, Welsh N, Jonas JC, Jörns A, Lenzen S, Eizirik DL. Mechanisms of pancreatic beta-cell death in type 1 and type 2 diabetes: many differences, few similarities. Diabetes. 2005;54 Suppl 2:S97-107.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1089]  [Cited by in F6Publishing: 1080]  [Article Influence: 56.8]  [Reference Citation Analysis (0)]
65.  Larsen CM, Faulenbach M, Vaag A, Vølund A, Ehses J, Seifert B, Mandrup-Poulsen T, Donath M. Interleukin-1 receptor antagonist-treatment of patients with type 2 diabetes. Ugeskr Laeger. 2007;169:3868-3871.  [PubMed]  [DOI]  [Cited in This Article: ]
66.  Böni-Schnetzler M, Thorne J, Parnaud G, Marselli L, Ehses JA, Kerr-Conte J, Pattou F, Halban PA, Weir GC, Donath MY. Increased interleukin (IL)-1beta messenger ribonucleic acid expression in beta -cells of individuals with type 2 diabetes and regulation of IL-1beta in human islets by glucose and autostimulation. J Clin Endocrinol Metab. 2008;93:4065-4074.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 246]  [Cited by in F6Publishing: 242]  [Article Influence: 15.1]  [Reference Citation Analysis (0)]
67.  Zhou R, Tardivel A, Thorens B, Choi I, Tschopp J. Thioredoxin-interacting protein links oxidative stress to inflammasome activation. Nat Immunol. 2010;11:136-140.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1681]  [Cited by in F6Publishing: 1921]  [Article Influence: 128.1]  [Reference Citation Analysis (0)]
68.  Welsh N, Cnop M, Kharroubi I, Bugliani M, Lupi R, Marchetti P, Eizirik DL. Is there a role for locally produced interleukin-1 in the deleterious effects of high glucose or the type 2 diabetes milieu to human pancreatic islets? Diabetes. 2005;54:3238-3244.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 104]  [Cited by in F6Publishing: 109]  [Article Influence: 5.7]  [Reference Citation Analysis (0)]
69.  Eguchi K, Manabe I. Macrophages and islet inflammation in type 2 diabetes. Diabetes Obes Metab. 2013;15 Suppl 3:152-158.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 87]  [Article Influence: 7.9]  [Reference Citation Analysis (0)]
70.  Dinarello CA. Immunological and inflammatory functions of the interleukin-1 family. Annu Rev Immunol. 2009;27:519-550.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2289]  [Cited by in F6Publishing: 2438]  [Article Influence: 162.5]  [Reference Citation Analysis (0)]
71.  Ehses JA, Lacraz G, Giroix MH, Schmidlin F, Coulaud J, Kassis N, Irminger JC, Kergoat M, Portha B, Homo-Delarche F. IL-1 antagonism reduces hyperglycemia and tissue inflammation in the type 2 diabetic GK rat. Proc Natl Acad Sci USA. 2009;106:13998-14003.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 256]  [Cited by in F6Publishing: 265]  [Article Influence: 17.7]  [Reference Citation Analysis (0)]
72.  Mahdi T, Hänzelmann S, Salehi A, Muhammed SJ, Reinbothe TM, Tang Y, Axelsson AS, Zhou Y, Jing X, Almgren P. Secreted frizzled-related protein 4 reduces insulin secretion and is overexpressed in type 2 diabetes. Cell Metab. 2012;16:625-633.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 143]  [Cited by in F6Publishing: 133]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
73.  Richardson SJ, Willcox A, Bone AJ, Foulis AK, Morgan NG. Islet-associated macrophages in type 2 diabetes. Diabetologia. 2009;52:1686-1688.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 172]  [Cited by in F6Publishing: 166]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
74.  Kahn SE. The relative contributions of insulin resistance and beta-cell dysfunction to the pathophysiology of Type 2 diabetes. Diabetologia. 2003;46:3-19.  [PubMed]  [DOI]  [Cited in This Article: ]
75.  Talchai C, Xuan S, Lin HV, Sussel L, Accili D. Pancreatic β cell dedifferentiation as a mechanism of diabetic β cell failure. Cell. 2012;150:1223-1234.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 939]  [Cited by in F6Publishing: 1023]  [Article Influence: 85.3]  [Reference Citation Analysis (0)]
76.  Kaneto H, Kajimoto Y, Miyagawa J, Matsuoka T, Fujitani Y, Umayahara Y, Hanafusa T, Matsuzawa Y, Yamasaki Y, Hori M. Beneficial effects of antioxidants in diabetes: possible protection of pancreatic beta-cells against glucose toxicity. Diabetes. 1999;48:2398-2406.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 502]  [Cited by in F6Publishing: 472]  [Article Influence: 18.9]  [Reference Citation Analysis (0)]
77.  Chakir H, Lefebvre DE, Wang H, Caraher E, Scott FW. Wheat protein-induced proinflammatory T helper 1 bias in mesenteric lymph nodes of young diabetes-prone rats. Diabetologia. 2005;48:1576-1584.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 27]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
78.  Poitout V, Robertson RP. Glucolipotoxicity: fuel excess and beta-cell dysfunction. Endocr Rev. 2008;29:351-366.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 781]  [Cited by in F6Publishing: 758]  [Article Influence: 47.4]  [Reference Citation Analysis (0)]
79.  Bachar E, Ariav Y, Ketzinel-Gilad M, Cerasi E, Kaiser N, Leibowitz G. Glucose amplifies fatty acid-induced endoplasmic reticulum stress in pancreatic beta-cells via activation of mTORC1. PLoS One. 2009;4:e4954.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 102]  [Cited by in F6Publishing: 105]  [Article Influence: 7.0]  [Reference Citation Analysis (0)]
80.  Maedler K, Spinas GA, Lehmann R, Sergeev P, Weber M, Fontana A, Kaiser N, Donath MY. Glucose induces beta-cell apoptosis via upregulation of the Fas receptor in human islets. Diabetes. 2001;50:1683-1690.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 273]  [Cited by in F6Publishing: 280]  [Article Influence: 12.2]  [Reference Citation Analysis (0)]
81.  Maedler K, Fontana A, Ris F, Sergeev P, Toso C, Oberholzer J, Lehmann R, Bachmann F, Tasinato A, Spinas GA. FLIP switches Fas-mediated glucose signaling in human pancreatic beta cells from apoptosis to cell replication. Proc Natl Acad Sci USA. 2002;99:8236-8241.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 109]  [Cited by in F6Publishing: 116]  [Article Influence: 5.3]  [Reference Citation Analysis (0)]
82.  Weinberg JM. Lipotoxicity. Kidney Int. 2006;70:1560-1566.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 268]  [Cited by in F6Publishing: 275]  [Article Influence: 15.3]  [Reference Citation Analysis (0)]
83.  Cnop M. Fatty acids and glucolipotoxicity in the pathogenesis of Type 2 diabetes. Biochem Soc Trans. 2008;36:348-352.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 141]  [Cited by in F6Publishing: 144]  [Article Influence: 9.0]  [Reference Citation Analysis (0)]
84.  Chang-Chen KJ, Mullur R, Bernal-Mizrachi E. Beta-cell failure as a complication of diabetes. Rev Endocr Metab Disord. 2008;9:329-343.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 156]  [Cited by in F6Publishing: 161]  [Article Influence: 10.1]  [Reference Citation Analysis (0)]
85.  Fonseca SG, Gromada J, Urano F. Endoplasmic reticulum stress and pancreatic β-cell death. Trends Endocrinol Metab. 2011;22:266-274.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 189]  [Article Influence: 14.5]  [Reference Citation Analysis (0)]
86.  Shimabukuro M, Wang MY, Zhou YT, Newgard CB, Unger RH. Protection against lipoapoptosis of beta cells through leptin-dependent maintenance of Bcl-2 expression. Proc Natl Acad Sci USA. 1998;95:9558-9561.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 169]  [Cited by in F6Publishing: 173]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
87.  Paumen MB, Ishida Y, Muramatsu M, Yamamoto M, Honjo T. Inhibition of carnitine palmitoyltransferase I augments sphingolipid synthesis and palmitate-induced apoptosis. J Biol Chem. 1997;272:3324-3329.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 272]  [Cited by in F6Publishing: 288]  [Article Influence: 10.7]  [Reference Citation Analysis (0)]
88.  Couri CE, Oliveira MC, Stracieri AB, Moraes DA, Pieroni F, Barros GM, Madeira MI, Malmegrim KC, Foss-Freitas MC, Simões BP. C-peptide levels and insulin independence following autologous nonmyeloablative hematopoietic stem cell transplantation in newly diagnosed type 1 diabetes mellitus. JAMA. 2009;301:1573-1579.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 306]  [Cited by in F6Publishing: 274]  [Article Influence: 18.3]  [Reference Citation Analysis (0)]
89.  Koves TR, Ussher JR, Noland RC, Slentz D, Mosedale M, Ilkayeva O, Bain J, Stevens R, Dyck JR, Newgard CB. Mitochondrial overload and incomplete fatty acid oxidation contribute to skeletal muscle insulin resistance. Cell Metab. 2008;7:45-56.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1394]  [Cited by in F6Publishing: 1414]  [Article Influence: 88.4]  [Reference Citation Analysis (0)]
90.  Montell E, Turini M, Marotta M, Roberts M, Noé V, Ciudad CJ, Macé K, Gómez-Foix AM. DAG accumulation from saturated fatty acids desensitizes insulin stimulation of glucose uptake in muscle cells. Am J Physiol Endocrinol Metab. 2001;280:E229-E237.  [PubMed]  [DOI]  [Cited in This Article: ]
91.  Perseghin G, Scifo P, De Cobelli F, Pagliato E, Battezzati A, Arcelloni C, Vanzulli A, Testolin G, Pozza G, Del Maschio A. Intramyocellular triglyceride content is a determinant of in vivo insulin resistance in humans: a 1H-13C nuclear magnetic resonance spectroscopy assessment in offspring of type 2 diabetic parents. Diabetes. 1999;48:1600-1606.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 666]  [Cited by in F6Publishing: 622]  [Article Influence: 24.9]  [Reference Citation Analysis (0)]
92.  Holzer RG, Park EJ, Li N, Tran H, Chen M, Choi C, Solinas G, Karin M. Saturated fatty acids induce c-Src clustering within membrane subdomains, leading to JNK activation. Cell. 2011;147:173-184.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 201]  [Cited by in F6Publishing: 212]  [Article Influence: 16.3]  [Reference Citation Analysis (0)]
93.  Steinberg GR, Schertzer JD. AMPK promotes macrophage fatty acid oxidative metabolism to mitigate inflammation: implications for diabetes and cardiovascular disease. Immunol Cell Biol. 2014;92:340-345.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 100]  [Article Influence: 10.0]  [Reference Citation Analysis (0)]
94.  Wellen KE, Hotamisligil GS. Inflammation, stress, and diabetes. J Clin Invest. 2005;115:1111-1119.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1182]  [Cited by in F6Publishing: 1075]  [Article Influence: 56.6]  [Reference Citation Analysis (0)]
95.  Jialal I, Huet BA, Kaur H, Chien A, Devaraj S. Increased toll-like receptor activity in patients with metabolic syndrome. Diabetes Care. 2012;35:900-904.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 124]  [Cited by in F6Publishing: 124]  [Article Influence: 10.3]  [Reference Citation Analysis (0)]
96.  Shi H, Kokoeva MV, Inouye K, Tzameli I, Yin H, Flier JS. TLR4 links innate immunity and fatty acid-induced insulin resistance. J Clin Invest. 2006;116:3015-3025.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2565]  [Cited by in F6Publishing: 2597]  [Article Influence: 144.3]  [Reference Citation Analysis (0)]
97.  Lee JY, Sohn KH, Rhee SH, Hwang D. Saturated fatty acids, but not unsaturated fatty acids, induce the expression of cyclooxygenase-2 mediated through Toll-like receptor 4. J Biol Chem. 2001;276:16683-16689.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 920]  [Cited by in F6Publishing: 907]  [Article Influence: 39.4]  [Reference Citation Analysis (0)]
98.  Lee JY, Zhao L, Youn HS, Weatherill AR, Tapping R, Feng L, Lee WH, Fitzgerald KA, Hwang DH. Saturated fatty acid activates but polyunsaturated fatty acid inhibits Toll-like receptor 2 dimerized with Toll-like receptor 6 or 1. J Biol Chem. 2004;279:16971-16979.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 344]  [Cited by in F6Publishing: 349]  [Article Influence: 17.5]  [Reference Citation Analysis (0)]
99.  Schaeffler A, Gross P, Buettner R, Bollheimer C, Buechler C, Neumeier M, Kopp A, Schoelmerich J, Falk W. Fatty acid-induced induction of Toll-like receptor-4/nuclear factor-kappaB pathway in adipocytes links nutritional signalling with innate immunity. Immunology. 2009;126:233-245.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 254]  [Cited by in F6Publishing: 266]  [Article Influence: 16.6]  [Reference Citation Analysis (0)]
100.  Senn JJ. Toll-like receptor-2 is essential for the development of palmitate-induced insulin resistance in myotubes. J Biol Chem. 2006;281:26865-26875.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 199]  [Cited by in F6Publishing: 199]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
101.  Song MJ, Kim KH, Yoon JM, Kim JB. Activation of Toll-like receptor 4 is associated with insulin resistance in adipocytes. Biochem Biophys Res Commun. 2006;346:739-745.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 319]  [Cited by in F6Publishing: 288]  [Article Influence: 16.0]  [Reference Citation Analysis (0)]
102.  Wong SW, Kwon MJ, Choi AM, Kim HP, Nakahira K, Hwang DH. Fatty acids modulate Toll-like receptor 4 activation through regulation of receptor dimerization and recruitment into lipid rafts in a reactive oxygen species-dependent manner. J Biol Chem. 2009;284:27384-27392.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 389]  [Cited by in F6Publishing: 389]  [Article Influence: 25.9]  [Reference Citation Analysis (0)]
103.  Seimon TA, Nadolski MJ, Liao X, Magallon J, Nguyen M, Feric NT, Koschinsky ML, Harkewicz R, Witztum JL, Tsimikas S. Atherogenic lipids and lipoproteins trigger CD36-TLR2-dependent apoptosis in macrophages undergoing endoplasmic reticulum stress. Cell Metab. 2010;12:467-482.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 350]  [Cited by in F6Publishing: 339]  [Article Influence: 24.2]  [Reference Citation Analysis (0)]
104.  Pal D, Dasgupta S, Kundu R, Maitra S, Das G, Mukhopadhyay S, Ray S, Majumdar SS, Bhattacharya S. Fetuin-A acts as an endogenous ligand of TLR4 to promote lipid-induced insulin resistance. Nat Med. 2012;18:1279-1285.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 615]  [Cited by in F6Publishing: 641]  [Article Influence: 53.4]  [Reference Citation Analysis (0)]
105.  Park SW, Zhou Y, Lee J, Lu A, Sun C, Chung J, Ueki K, Ozcan U. The regulatory subunits of PI3K, p85alpha and p85beta, interact with XBP-1 and increase its nuclear translocation. Nat Med. 2010;16:429-437.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 224]  [Cited by in F6Publishing: 227]  [Article Influence: 16.2]  [Reference Citation Analysis (0)]
106.  Li J, Jin C, Cleveland JC, Ao L, Xu D, Fullerton DA, Meng X. Enhanced inflammatory responses to toll-like receptor 2/4 stimulation in type 1 diabetic coronary artery endothelial cells: the effect of insulin. Cardiovasc Diabetol. 2010;9:90.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 21]  [Cited by in F6Publishing: 25]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
107.  Lin M, Yiu WH, Wu HJ, Chan LY, Leung JC, Au WS, Chan KW, Lai KN, Tang SC. Toll-like receptor 4 promotes tubular inflammation in diabetic nephropathy. J Am Soc Nephrol. 2012;23:86-102.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 240]  [Cited by in F6Publishing: 281]  [Article Influence: 21.6]  [Reference Citation Analysis (0)]
108.  Vijay-Kumar M, Aitken JD, Carvalho FA, Cullender TC, Mwangi S, Srinivasan S, Sitaraman SV, Knight R, Ley RE, Gewirtz AT. Metabolic syndrome and altered gut microbiota in mice lacking Toll-like receptor 5. Science. 2010;328:228-231.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1496]  [Cited by in F6Publishing: 1501]  [Article Influence: 107.2]  [Reference Citation Analysis (0)]
109.  Wen H, Gris D, Lei Y, Jha S, Zhang L, Huang MT, Brickey WJ, Ting JP. Fatty acid-induced NLRP3-ASC inflammasome activation interferes with insulin signaling. Nat Immunol. 2011;12:408-415.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1214]  [Cited by in F6Publishing: 1313]  [Article Influence: 101.0]  [Reference Citation Analysis (0)]
110.  Deopurkar R, Ghanim H, Friedman J, Abuaysheh S, Sia CL, Mohanty P, Viswanathan P, Chaudhuri A, Dandona P. Differential effects of cream, glucose, and orange juice on inflammation, endotoxin, and the expression of Toll-like receptor-4 and suppressor of cytokine signaling-3. Diabetes Care. 2010;33:991-997.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 186]  [Cited by in F6Publishing: 192]  [Article Influence: 13.7]  [Reference Citation Analysis (0)]
111.  Yan SF, Ramasamy R, Schmidt AM. Receptor for AGE (RAGE) and its ligands-cast into leading roles in diabetes and the inflammatory response. J Mol Med (Berl). 2009;87:235-247.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 167]  [Cited by in F6Publishing: 169]  [Article Influence: 11.3]  [Reference Citation Analysis (0)]
112.  Bierhaus A, Nawroth PP. Multiple levels of regulation determine the role of the receptor for AGE (RAGE) as common soil in inflammation, immune responses and diabetes mellitus and its complications. Diabetologia. 2009;52:2251-2263.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 219]  [Cited by in F6Publishing: 216]  [Article Influence: 14.4]  [Reference Citation Analysis (0)]
113.  Wentworth JM, Fourlanos S, Harrison LC. Reappraising the stereotypes of diabetes in the modern diabetogenic environment. Nat Rev Endocrinol. 2009;5:483-489.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 38]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
114.  Brooks-Worrell B, Tree T, Mannering SI, Durinovic-Bello I, James E, Gottlieb P, Wong S, Zhou Z, Yang L, Cilio CM. Comparison of cryopreservation methods on T-cell responses to islet and control antigens from type 1 diabetic patients and controls. Diabetes Metab Res Rev. 2011;27:737-745.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 19]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
115.  Biason-Lauber A, Böni-Schnetzler M, Hubbard BP, Bouzakri K, Brunner A, Cavelti-Weder C, Keller C, Meyer-Böni M, Meier DT, Brorsson C. Identification of a SIRT1 mutation in a family with type 1 diabetes. Cell Metab. 2013;17:448-455.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 89]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
116.  Chan JK, Roth J, Oppenheim JJ, Tracey KJ, Vogl T, Feldmann M, Horwood N, Nanchahal J. Alarmins: awaiting a clinical response. J Clin Invest. 2012;122:2711-2719.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 341]  [Cited by in F6Publishing: 348]  [Article Influence: 29.0]  [Reference Citation Analysis (0)]
117.  Misawa T, Takahama M, Kozaki T, Lee H, Zou J, Saitoh T, Akira S. Microtubule-driven spatial arrangement of mitochondria promotes activation of the NLRP3 inflammasome. Nat Immunol. 2013;14:454-460.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 546]  [Cited by in F6Publishing: 569]  [Article Influence: 51.7]  [Reference Citation Analysis (0)]
118.  Nishimura S, Manabe I, Nagasaki M, Eto K, Yamashita H, Ohsugi M, Otsu M, Hara K, Ueki K, Sugiura S. CD8+ effector T cells contribute to macrophage recruitment and adipose tissue inflammation in obesity. Nat Med. 2009;15:914-920.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1571]  [Cited by in F6Publishing: 1632]  [Article Influence: 108.8]  [Reference Citation Analysis (0)]
119.  Feuerer M, Herrero L, Cipolletta D, Naaz A, Wong J, Nayer A, Lee J, Goldfine AB, Benoist C, Shoelson S. Lean, but not obese, fat is enriched for a unique population of regulatory T cells that affect metabolic parameters. Nat Med. 2009;15:930-939.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1622]  [Cited by in F6Publishing: 1548]  [Article Influence: 103.2]  [Reference Citation Analysis (0)]
120.  Winer DA, Winer S, Shen L, Wadia PP, Yantha J, Paltser G, Tsui H, Wu P, Davidson MG, Alonso MN. B cells promote insulin resistance through modulation of T cells and production of pathogenic IgG antibodies. Nat Med. 2011;17:610-617.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 702]  [Cited by in F6Publishing: 732]  [Article Influence: 56.3]  [Reference Citation Analysis (0)]
121.  Yang H, Youm YH, Vandanmagsar B, Ravussin A, Gimble JM, Greenway F, Stephens JM, Mynatt RL, Dixit VD. Obesity increases the production of proinflammatory mediators from adipose tissue T cells and compromises TCR repertoire diversity: implications for systemic inflammation and insulin resistance. J Immunol. 2010;185:1836-1845.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 309]  [Cited by in F6Publishing: 329]  [Article Influence: 23.5]  [Reference Citation Analysis (0)]
122.  Jagannathan-Bogdan M, McDonnell ME, Shin H, Rehman Q, Hasturk H, Apovian CM, Nikolajczyk BS. Elevated proinflammatory cytokine production by a skewed T cell compartment requires monocytes and promotes inflammation in type 2 diabetes. J Immunol. 2011;186:1162-1172.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 283]  [Cited by in F6Publishing: 285]  [Article Influence: 20.4]  [Reference Citation Analysis (0)]
123.  McGillicuddy FC, Chiquoine EH, Hinkle CC, Kim RJ, Shah R, Roche HM, Smyth EM, Reilly MP. Interferon gamma attenuates insulin signaling, lipid storage, and differentiation in human adipocytes via activation of the JAK/STAT pathway. J Biol Chem. 2009;284:31936-31944.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 164]  [Cited by in F6Publishing: 189]  [Article Influence: 12.6]  [Reference Citation Analysis (0)]
124.  Jagannathan M, McDonnell M, Liang Y, Hasturk H, Hetzel J, Rubin D, Kantarci A, Van Dyke TE, Ganley-Leal LM, Nikolajczyk BS. Toll-like receptors regulate B cell cytokine production in patients with diabetes. Diabetologia. 2010;53:1461-1471.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 119]  [Cited by in F6Publishing: 120]  [Article Influence: 8.6]  [Reference Citation Analysis (0)]
125.  Kharroubi I, Ladrière L, Cardozo AK, Dogusan Z, Cnop M, Eizirik DL. Free fatty acids and cytokines induce pancreatic beta-cell apoptosis by different mechanisms: role of nuclear factor-kappaB and endoplasmic reticulum stress. Endocrinology. 2004;145:5087-5096.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 452]  [Cited by in F6Publishing: 439]  [Article Influence: 22.0]  [Reference Citation Analysis (0)]
126.  Hung JH, Su IJ, Lei HY, Wang HC, Lin WC, Chang WT, Huang W, Chang WC, Chang YS, Chen CC. Endoplasmic reticulum stress stimulates the expression of cyclooxygenase-2 through activation of NF-kappaB and pp38 mitogen-activated protein kinase. J Biol Chem. 2004;279:46384-46392.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 191]  [Cited by in F6Publishing: 207]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
127.  Urano F, Wang X, Bertolotti A, Zhang Y, Chung P, Harding HP, Ron D. Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science. 2000;287:664-666.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2199]  [Cited by in F6Publishing: 2229]  [Article Influence: 92.9]  [Reference Citation Analysis (0)]
128.  O’Neill LA, Bryant CE, Doyle SL. Therapeutic targeting of Toll-like receptors for infectious and inflammatory diseases and cancer. Pharmacol Rev. 2009;61:177-197.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 318]  [Cited by in F6Publishing: 335]  [Article Influence: 22.3]  [Reference Citation Analysis (0)]
129.  O’Neill LA. The interleukin-1 receptor/Toll-like receptor superfamily: 10 years of progress. Immunol Rev. 2008;226:10-18.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 452]  [Cited by in F6Publishing: 458]  [Article Influence: 30.5]  [Reference Citation Analysis (0)]
130.  Cunha DA, Igoillo-Esteve M, Gurzov EN, Germano CM, Naamane N, Marhfour I, Fukaya M, Vanderwinden JM, Gysemans C, Mathieu C. Death protein 5 and p53-upregulated modulator of apoptosis mediate the endoplasmic reticulum stress-mitochondrial dialog triggering lipotoxic rodent and human β-cell apoptosis. Diabetes. 2012;61:2763-2775.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 101]  [Cited by in F6Publishing: 109]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
131.  Flamment M, Hajduch E, Ferré P, Foufelle F. New insights into ER stress-induced insulin resistance. Trends Endocrinol Metab. 2012;23:381-390.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 210]  [Cited by in F6Publishing: 231]  [Article Influence: 19.3]  [Reference Citation Analysis (0)]
132.  Medzhitov R. Toll-like receptors and innate immunity. Nat Rev Immunol. 2001;1:135-145.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2861]  [Cited by in F6Publishing: 2788]  [Article Influence: 121.2]  [Reference Citation Analysis (0)]
133.  Lanuza-Masdeu J, Arévalo MI, Vila C, Barberà A, Gomis R, Caelles C. In vivo JNK activation in pancreatic β-cells leads to glucose intolerance caused by insulin resistance in pancreas. Diabetes. 2013;62:2308-2317.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 42]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
134.  Harding HP, Zeng H, Zhang Y, Jungries R, Chung P, Plesken H, Sabatini DD, Ron D. Diabetes mellitus and exocrine pancreatic dysfunction in perk-/- mice reveals a role for translational control in secretory cell survival. Mol Cell. 2001;7:1153-1163.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 933]  [Cited by in F6Publishing: 927]  [Article Influence: 40.3]  [Reference Citation Analysis (0)]
135.  Zhang W, Feng D, Li Y, Iida K, McGrath B, Cavener DR. PERK EIF2AK3 control of pancreatic beta cell differentiation and proliferation is required for postnatal glucose homeostasis. Cell Metab. 2006;4:491-497.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 205]  [Cited by in F6Publishing: 210]  [Article Influence: 11.7]  [Reference Citation Analysis (0)]
136.  Boyce M, Bryant KF, Jousse C, Long K, Harding HP, Scheuner D, Kaufman RJ, Ma D, Coen DM, Ron D. A selective inhibitor of eIF2alpha dephosphorylation protects cells from ER stress. Science. 2005;307:935-939.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1094]  [Cited by in F6Publishing: 1148]  [Article Influence: 60.4]  [Reference Citation Analysis (0)]
137.  Wang R, McGrath BC, Kopp RF, Roe MW, Tang X, Chen G, Cavener DR. Insulin secretion and Ca2+ dynamics in β-cells are regulated by PERK (EIF2AK3) in concert with calcineurin. J Biol Chem. 2013;288:33824-33836.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 66]  [Cited by in F6Publishing: 71]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
138.  Lipson KL, Ghosh R, Urano F. The role of IRE1alpha in the degradation of insulin mRNA in pancreatic beta-cells. PLoS One. 2008;3:e1648.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 140]  [Cited by in F6Publishing: 144]  [Article Influence: 9.0]  [Reference Citation Analysis (0)]
139.  Haze K, Okada T, Yoshida H, Yanagi H, Yura T, Negishi M, Mori K. Identification of the G13 (cAMP-response-element-binding protein-related protein) gene product related to activating transcription factor 6 as a transcriptional activator of the mammalian unfolded protein response. Biochem J. 2001;355:19-28.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 17]  [Reference Citation Analysis (0)]
140.  Lee AH, Chu GC, Iwakoshi NN, Glimcher LH. XBP-1 is required for biogenesis of cellular secretory machinery of exocrine glands. EMBO J. 2005;24:4368-4380.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 332]  [Cited by in F6Publishing: 347]  [Article Influence: 18.3]  [Reference Citation Analysis (0)]
141.  Lee AH, Heidtman K, Hotamisligil GS, Glimcher LH. Dual and opposing roles of the unfolded protein response regulated by IRE1alpha and XBP1 in proinsulin processing and insulin secretion. Proc Natl Acad Sci USA. 2011;108:8885-8890.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 187]  [Cited by in F6Publishing: 196]  [Article Influence: 15.1]  [Reference Citation Analysis (0)]
142.  Seo HY, Kim YD, Lee KM, Min AK, Kim MK, Kim HS, Won KC, Park JY, Lee KU, Choi HS. Endoplasmic reticulum stress-induced activation of activating transcription factor 6 decreases insulin gene expression via up-regulation of orphan nuclear receptor small heterodimer partner. Endocrinology. 2008;149:3832-3841.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 84]  [Cited by in F6Publishing: 91]  [Article Influence: 5.7]  [Reference Citation Analysis (0)]
143.  Teodoro T, Odisho T, Sidorova E, Volchuk A. Pancreatic β-cells depend on basal expression of active ATF6α-p50 for cell survival even under nonstress conditions. Am J Physiol Cell Physiol. 2012;302:C992-1003.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 48]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
144.  Gallucci S, Lolkema M, Matzinger P. Natural adjuvants: endogenous activators of dendritic cells. Nat Med. 1999;5:1249-1255.  [PubMed]  [DOI]  [Cited in This Article: ]
145.  Schroder K, Tschopp J. The inflammasomes. Cell. 2010;140:821-832.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3880]  [Cited by in F6Publishing: 4225]  [Article Influence: 301.8]  [Reference Citation Analysis (0)]
146.  Tannahill GM, O’Neill LA. The emerging role of metabolic regulation in the functioning of Toll-like receptors and the NOD-like receptor Nlrp3. FEBS Lett. 2011;585:1568-1572.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 55]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
147.  Wen H, Ting JP, O’Neill LA. A role for the NLRP3 inflammasome in metabolic diseases--did Warburg miss inflammation? Nat Immunol. 2012;13:352-357.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 323]  [Cited by in F6Publishing: 355]  [Article Influence: 29.6]  [Reference Citation Analysis (0)]
148.  Haneklaus M, O’Neill LA, Coll RC. Modulatory mechanisms controlling the NLRP3 inflammasome in inflammation: recent developments. Curr Opin Immunol. 2013;25:40-45.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 151]  [Cited by in F6Publishing: 141]  [Article Influence: 12.8]  [Reference Citation Analysis (0)]
149.  Kufer TA, Sansonetti PJ. NLR functions beyond pathogen recognition. Nat Immunol. 2011;12:121-128.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 145]  [Cited by in F6Publishing: 156]  [Article Influence: 12.0]  [Reference Citation Analysis (0)]
150.  Park JS, Svetkauskaite D, He Q, Kim JY, Strassheim D, Ishizaka A, Abraham E. Involvement of toll-like receptors 2 and 4 in cellular activation by high mobility group box 1 protein. J Biol Chem. 2004;279:7370-7377.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1174]  [Cited by in F6Publishing: 1218]  [Article Influence: 60.9]  [Reference Citation Analysis (0)]
151.  Masters SL, Dunne A, Subramanian SL, Hull RL, Tannahill GM, Sharp FA, Becker C, Franchi L, Yoshihara E, Chen Z. Activation of the NLRP3 inflammasome by islet amyloid polypeptide provides a mechanism for enhanced IL-1β in type 2 diabetes. Nat Immunol. 2010;11:897-904.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 929]  [Cited by in F6Publishing: 978]  [Article Influence: 69.9]  [Reference Citation Analysis (0)]
152.  Schroder K, Zhou R, Tschopp J. The NLRP3 inflammasome: a sensor for metabolic danger? Science. 2010;327:296-300.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 778]  [Cited by in F6Publishing: 818]  [Article Influence: 58.4]  [Reference Citation Analysis (0)]
153.  Lee HM, Kim JJ, Kim HJ, Shong M, Ku BJ, Jo EK. Upregulated NLRP3 inflammasome activation in patients with type 2 diabetes. Diabetes. 2013;62:194-204.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 489]  [Cited by in F6Publishing: 525]  [Article Influence: 47.7]  [Reference Citation Analysis (0)]
154.  Duewell P, Kono H, Rayner KJ, Sirois CM, Vladimer G, Bauernfeind FG, Abela GS, Franchi L, Nuñez G, Schnurr M. NLRP3 inflammasomes are required for atherogenesis and activated by cholesterol crystals. Nature. 2010;464:1357-1361.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2432]  [Cited by in F6Publishing: 2730]  [Article Influence: 195.0]  [Reference Citation Analysis (0)]
155.  Lerner AG, Upton JP, Praveen PV, Ghosh R, Nakagawa Y, Igbaria A, Shen S, Nguyen V, Backes BJ, Heiman M. IRE1α induces thioredoxin-interacting protein to activate the NLRP3 inflammasome and promote programmed cell death under irremediable ER stress. Cell Metab. 2012;16:250-264.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 601]  [Cited by in F6Publishing: 636]  [Article Influence: 53.0]  [Reference Citation Analysis (0)]
156.  Oslowski CM, Hara T, O’Sullivan-Murphy B, Kanekura K, Lu S, Hara M, Ishigaki S, Zhu LJ, Hayashi E, Hui ST. Thioredoxin-interacting protein mediates ER stress-induced β cell death through initiation of the inflammasome. Cell Metab. 2012;16:265-273.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 488]  [Cited by in F6Publishing: 513]  [Article Influence: 42.8]  [Reference Citation Analysis (0)]
157.  Lawlor KE, Vince JE. Ambiguities in NLRP3 inflammasome regulation: is there a role for mitochondria? Biochim Biophys Acta. 2014;1840:1433-1440.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 81]  [Article Influence: 7.4]  [Reference Citation Analysis (0)]
158.  Zhou R, Yazdi AS, Menu P, Tschopp J. A role for mitochondria in NLRP3 inflammasome activation. Nature. 2011;469:221-225.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3253]  [Cited by in F6Publishing: 3774]  [Article Influence: 269.6]  [Reference Citation Analysis (0)]
159.  Oka S, Yoshihara E, Bizen-Abe A, Liu W, Watanabe M, Yodoi J, Masutani H. Thioredoxin binding protein-2/thioredoxin-interacting protein is a critical regulator of insulin secretion and peroxisome proliferator-activated receptor function. Endocrinology. 2009;150:1225-1234.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 67]  [Cited by in F6Publishing: 71]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
160.  Youm YH, Adijiang A, Vandanmagsar B, Burk D, Ravussin A, Dixit VD. Elimination of the NLRP3-ASC inflammasome protects against chronic obesity-induced pancreatic damage. Endocrinology. 2011;152:4039-4045.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 123]  [Cited by in F6Publishing: 126]  [Article Influence: 9.7]  [Reference Citation Analysis (0)]
161.  Westwell-Roper C, Dai DL, Soukhatcheva G, Potter KJ, van Rooijen N, Ehses JA, Verchere CB. IL-1 blockade attenuates islet amyloid polypeptide-induced proinflammatory cytokine release and pancreatic islet graft dysfunction. J Immunol. 2011;187:2755-2765.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 141]  [Cited by in F6Publishing: 144]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
162.  Dasu MR, Devaraj S, Park S, Jialal I. Increased toll-like receptor (TLR) activation and TLR ligands in recently diagnosed type 2 diabetic subjects. Diabetes Care. 2010;33:861-868.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 390]  [Cited by in F6Publishing: 411]  [Article Influence: 29.4]  [Reference Citation Analysis (0)]