Skip to main content
Endocrine Reviews logoLink to Endocrine Reviews
. 2008 Feb 27;29(3):303–316. doi: 10.1210/er.2007-0037

Islet Amyloid in Type 2 Diabetes, and the Toxic Oligomer Hypothesis

Leena Haataja 1, Tatyana Gurlo 1, Chang J Huang 1, Peter C Butler 1
PMCID: PMC2528855  PMID: 18314421

Abstract

Type 2 diabetes (T2DM) is characterized by insulin resistance, defective insulin secretion, loss of β-cell mass with increased β-cell apoptosis and islet amyloid. The islet amyloid is derived from islet amyloid polypeptide (IAPP, amylin), a protein coexpressed and cosecreted with insulin by pancreatic β-cells. In common with other amyloidogenic proteins, IAPP has the propensity to form membrane permeant toxic oligomers. Accumulating evidence suggests that these toxic oligomers, rather than the extracellular amyloid form of these proteins, are responsible for loss of neurons in neurodegenerative diseases. In this review we discuss emerging evidence to suggest that formation of intracellular IAPP oligomers may contribute to β-cell loss in T2DM. The accumulated evidence permits the amyloid hypothesis originally developed for neurodegenerative diseases to be reformulated as the toxic oligomer hypothesis. However, as in neurodegenerative diseases, it remains unclear exactly why amyloidogenic proteins form oligomers in vivo, what their exact structure is, and to what extent these oligomers play a primary or secondary role in the cytotoxicity in what are now often called unfolded protein diseases.


  • I. Introduction
    • A. β-Cell mass and type 2 diabetes
  • II. Islet Amyloid in T2DM and the Amyloid Hypothesis
    • A. IAPP physiological functions
    • B. IAPP species comparison
  • III. Lessons Learned from Transgenic Rodents; the Amyloid Hypothesis Challenged
    • A. hIAPP toxic oligomers, and not amyloid, induce β-cell apoptosis; the amyloid hypothesis modified
    • B. Where do hIAPP toxic oligomers form and act?
  • IV. Cellular Mechanisms to Prevent hIAPP Oligomer Formation. The Unfolded Protein Response (UPR), proIAPP Processing, and Vesicle Environment
    • A. Why do hIAPP oligomers (and amyloid) form?
    • B. hIAPP-induced β-cell apoptosis
  • V. Summary

I. Introduction

TYPE 2 DIABETES (T2DM) is a loosely defined clinical syndrome that likely has a number of different causes. Risk factors for T2DM include a positive family history, aging, and a variety of causes of insulin resistance, most commonly obesity (Refs. 1,2,3,4 and references therein). Most individuals respond to insulin resistance by adaptively increasing β-cell mass and insulin secretion to maintain normal blood glucose concentrations (5,6,7). This is consistent with the response of other endocrine organs when chronically stimulated. As an example, the low ionized calcium concentration in chronic renal failure provokes adaptive hyperplasia of parathyroid glands, not PTH failure (8).

Therefore, in those individuals who develop T2DM, the deficient adaptive response to increased insulin demand is an abnormal response. As such, by definition, T2DM is primarily due to insulin resistance and impaired insulin secretion.

A. β-Cell mass and type 2 diabetes

The underlying cause of impaired insulin secretion in T2DM is unknown and likely has multiple origins in different individuals. A relative deficit in the number of β-cells (often collectively referred to as β-cell mass) appears to be an important contributory factor (6,9,10). In obese individuals with impaired fasting glucose, β-cell mass is approximately 50% less than that of healthy individuals (6). The relationship between fasting blood glucose and β-cell mass is curvilinear, with a wide range of β-cell mass in nondiabetic individuals but with a steep increase in blood glucose with each decrement in β-cell mass beyond 50% (11) (Fig. 1).

Figure 1.

Figure 1

Relationship between percentage of pancreas volume occupied by β-cells and fasting plasma glucose in obese humans without insulin or oral antidiabetic treatment. The solid line is derived from nonlinear regression analysis (monoexponential fit, r = 0.50; P < 0.0001 by ANOVA). The dashed vertical lines indicate the mean β-cell area in obese nondiabetic subjects (OND) (right) and the computed inflection point of the curve (left). IFG, Impaired fasting glucose; OD, obese diabetic. Adapted from Ref. 11. [Copyright 2006 American Diabetes Association. From Diabetes Care 29:717–718. Reprinted with permission from The American Diabetes Association.]

In adult pigs, a 65% decrease in β-cell mass led to diabetes with most of the metabolic characteristics of T2DM (fasting and postprandial hyperglycemia, impaired insulin secretion, postprandial hyperglucagonemia) (12). Also, a 50% partial pancreatectomy in dogs or humans causes impaired fasting glucose initially and often diabetes subsequently (13,14,15,16). Collectively, these data imply that the deficit in β-cell mass present in T2DM can be sufficient to induce diabetes, particularly in the setting of associated insulin resistance.

II. Islet Amyloid in T2DM and the Amyloid Hypothesis

The islet in T2DM is characterized by what was originally referred to as hyaline deposits, later demonstrated to consist of amyloid (17). The term amyloid was developed to describe abnormal extracellular deposits of protein that appeared somewhat like amylopectin (starch; amyloid “like amylopectin”) (Fig. 2). Amyloid when present is always abnormal and consists of an insoluble protein precipitate, composed of protein monomers arranged in a β-pleated sheet structure (18). The resulting aggregated monomers appear as nonbranching fibrils by electron microscopy and are detected by congo red or thioflavine staining by light microscopy (Fig. 3). In 1987 two groups identified the constitutive protein in islet amyloid, naming it amylin and islet amyloid polypeptide (IAPP), respectively (19,20). Because amylin also became the name of a pharmaceutical company, we have preferred IAPP. IAPP is coexpressed and cosecreted with insulin by pancreatic β-cells (21,22,23).

Figure 2.

Figure 2

Human islets from T2DM subjects (right) have less β-cells than those from nondiabetic subjects (left) and contain deposits of amyloid (arrow) derived from IAPP. Human islets were stained for insulin. This figure originally appeared in an article by Matveyenko and Butler (39). It is reprinted with permission from the ILAR Journal, Institute for Laboratory Animal Research, The National Academies (www.nationalacademies.org/ilar).

Figure 3.

Figure 3

The islets from T2DM human (A), diabetic vervet monkey (B), and diabetic obese hemizygous hIAPP transgenic mouse (C) stained for amyloid using Congo red. (Unpublished images from the Butler laboratory.)

The islet in T2DM has parallels with the neuropathology in neurodegenerative diseases such as Alzheimer’s disease (24,25,26,27), Parkinson’s disease (28,29), prion encephalopathy (30), amyotrophic lateral sclerosis (31), and Huntington’s disease (32) (Table 1). In neurodegenerative diseases and T2DM, there is cell loss associated with abnormal aggregation of locally expressed protein. The proteins that form these aggregates share in common the propensity to form amyloid fibrils in an aqueous environment, prompting the question, is the formation of the amyloid fibrils in these diseases a cause or a consequence of the underlying cell attrition? Those in favor of a primary role of amyloid in neurodegenerative diseases coined the term “amyloid hypothesis” (24).

Table 1.

The common molecular basis of amyloid-related T2DM and neurodegenerative diseases

Diseases Protein that forms toxic oligomers Cells lost
Type 2 diabetes mellitus Islet amyloid polypeptide β-cells
Alzheimer’s disease β-Amyloid protein Cortical neurons
Parkinson’s disease Synuclein Dopaminergic neurons
Prion encephalopathy/transmissible spongiform encephalopathies Prion Cortical neurons
Amyotrophic lateral sclerosis Mutant superoxide dismutase Motor neurons
Polyglutamine/Huntington’s disease Huntingtin’s polyglutamine Pyramidal neurons

Evidence that was cited against the amyloid hypothesis is the observation that the severity of the disease state often correlates poorly with the extent of amyloid deposition (33,34,35). For example, the extent of dementia or cortical atrophy in Alzheimer’s disease does not correlate well with the extent of brain amyloid derived from amyloid β protein (AβP) hypothesis (36). Islet amyloid is found in nondiabetic individuals, particularly with aging (37), and is not present in all islets in people with T2DM (38).

Evidence that supports the amyloid hypothesis includes the recapitulation of disease states by some (but not all) transgenic models for amyloidogenic proteins and mutations in amyloidogenic proteins that lead to early onset disease (26,27,39). Mutations in AβP that increase its propensity to aggregate were identified in early onset familial forms of Alzheimer’s disease (25). A similar mutation in IAPP has been described in Japan and leads to an increased risk for T2DM (40).

This conflicting evidence for and against the amyloid hypothesis has been somewhat resolved by recent advances. The toxic form of amyloidogenic proteins appears not to be the extracellular amyloid fibrils detected by light microscopy, but rather smaller nonfibrillar oligomers (41,42,43,44). Before considering a revised model of the amyloid hypothesis further, we will briefly review the known properties of the protein that forms islet amyloid, IAPP.

A. IAPP physiological functions

IAPP is coexpressed with insulin by pancreatic β-cells (21,22,23). It is trafficked through the insulin secretory pathway and cosecreted with insulin, for example after meal ingestion. Although the physiological function of IAPP remains unknown, it is highly conserved between species, implying functional significance. Application of IAPP to rat soleus muscle strips was shown to inhibit insulin-mediated glucose uptake (45). The assumption was that the large amyloid deposits in the islet in T2DM would be associated with high circulating levels of IAPP, which in turn contributed to the insulin resistance of this disease. In reality, the circulating concentrations of IAPP (5–20 pm) were found to be far below that required to inhibit insulin action (in nanomoles), and furthermore plasma IAPP levels in T2DM were not increased compared with nondiabetic controls (21).

One well-characterized action of IAPP is a direct paracrine effect on β-cells to inhibit insulin secretion (46). It has also been suggested that IAPP delays gastric emptying and suppresses appetite (47), although it is unclear whether these actions are physiological actions of IAPP within normal circulating levels or are present only in response to pharmacological IAPP concentrations.

B. IAPP species comparison

IAPP is expressed as an 89-amino acid protein that undergoes processing to a 37-amino acid peptide (48). The primary sequence of IAPP is closely conserved between species, but there are some important differences in IAPP20–29, the region of the peptide believed to be important for conveying its propensity to form oligomers in an aqueous environment (49). As can be seen in Fig. 4, primates and humans share close homology in IAPP20–29, and synthetic forms of these peptides form amyloid. It is therefore intriguing that nonhuman primates and cats, like humans, are also prone to developing T2DM with a similar clinical course and islet pathology as that observed in humans (50).

Figure 4.

Figure 4

Alignment of IAPP ortholog proteins. Amino acid alignment of IAPP protein sequences identified in Homo sapiens (human, CAA39504), human mutant (S20G) (40), Macaca mulatto (monkey, XP_001098290), Felis catus (cat, NP_001036803), Canis lupus (dog, NP_001003233), Mus musculus (mouse, NP_034621), and Rattus norvegicus (rat, NP_036718). Dots correspond to conserved residues with human IAPP sequence. Red letters correspond to the amyloidic sequence.

In contrast, neither rats nor mice spontaneously develop T2DM. Rat and mouse IAPP20–29 is identical and not amyloidogenic, due to three proline residues that render rat and mouse IAPP water soluble (49). Moreover, the cytotoxicity of IAPP depends on its propensity to form oligomers (42,51,52). This distinction between human IAPP (hIAPP) and rodent IAPP (rIAPP) provides an opportunity to use the transgenic approach to examine the impact of hIAPP expression in mice or rats.

III. Lessons Learned from Transgenic Rodents; the Amyloid Hypothesis Challenged

Several different hIAPP transgenic mouse models have been reported (53,54,55,56,57,58,59,60). The phenotype and islet pathology of hIAPP transgenic rodents have been summarized in a recent review (39). Some but not all develop diabetes. In common with other mouse models of diabetes, there is a greater predisposition to diabetes in male mice compared with female mice, and also a background effect. hIAPP transgenic mice on a FVB background develop diabetes if IAPP expression is increased by induction of insulin resistance, whether through cross breeding onto an obese background (55,58) or pharmacologically (53). Alternatively, increasing the gene dosage of hIAPP by cross breeding hIAPP transgenic mice to homozygosity also leads to diabetes (54). The underlying mechanism for diabetes in hIAPP transgenic mice and rats is a deficit in β-cell mass due to increased β-cell apoptosis (59,61). The metabolic characteristics of T2DM, i.e., hyperglycemia, impaired insulin secretion, insulin resistance, and hyperglucagonemia, are all recapitulated in hIAPP (HIP) transgenic rats (62).

Because replicating β-cells are particularly vulnerable to hIAPP-induced apoptosis (63), the deficit in β-cell mass is due to loss of cells as well as an inability to adequately compensate through increased β-cell replication. Transgenic protein expression in β-cells has unexpectedly induced diabetes in mice (64). Therefore, it is important to note that comparably high β-cell-specific transgenic expression of soluble rIAPP in mice does not lead to increased β-cell apoptosis, loss of β-cell mass, or diabetes (65). Collectively, these studies imply that hIAPP can be expressed and trafficked successfully by mouse β-cells up to a threshold beyond which apoptosis may be induced, and this vulnerability to high expression rates of hIAPP depends on its propensity to oligomerize.

An important lesson that arose from studies of hIAPP transgenic mice is that the concept that extracellular IAPP amyloid causes β-cell apoptosis (the amyloid hypothesis) is implausible. Homozygous transgenic mice for hIAPP developed diabetes due to a high rate of β-cell apoptosis by 10 wk of age (54,65). However, extracellular islet amyloid was not yet present in these mice during the rapid loss of β-cells from age 5–10 wk. In obese hemizygous hIAPP transgenic mice that develop diabetes at approximately 20 wk of age, extensive islet amyloid develops, but there is no relationship between the extent of islet amyloid and the frequency of β-cell apoptosis (61). Moreover, the β-cells undergoing apoptosis are not adjacent to amyloid deposits, as would be expected if the amyloid deposits were the toxic form of amyloid (Fig. 5). This lack of proximity of β-cells undergoing apoptosis and islet amyloid is also evident in humans with T2DM (6). Further evidence against the toxicity of extracellular amyloid is provided by another hIAPP transgenic mouse model that develops extensive islet amyloid but not diabetes (66).

Figure 5.

Figure 5

The increased β-cell apoptosis in hemizygous hIAPP transgenic mice (OT) does not correspond to areas of amyloid. Islets from obese nontransgenic (ONT, left panels) and obese hemizygous hIAPP transgenic mice (OT, right panels) immunostained for insulin (upper panels) and corresponding islets stained for TUNEL (lower panels). Adapted from Ref. 61. [Copyright 2003 American Diabetes Association. From Diabetes 52:2304–2314. Reprinted with permission from The American Diabetes Association.]

Because cytotoxicity induced by hIAPP overexpression and amyloid formation were readily dissociated in hIAPP transgenic mice, the amyloid hypothesis in its literal form, that extracellular amyloid causes cytotoxicity, was challenged. The dissociation of hIAPP induced cytotoxicity and hIAPP amyloid formation gives rise to the question, what is the cytotoxic form of hIAPP?

A. hIAPP toxic oligomers, and not amyloid, induce β-cell apoptosis; the amyloid hypothesis modified

Cytoxicity by hIAPP was first reported when hIAPP or rIAPP was applied to human islet cells in culture (51). Application of hIAPP but not rIAPP caused β-cell apoptosis. Not surprisingly, the islet cells exposed to hIAPP (but nor rIAPP) were also subsequently observed to be decorated with amyloid. The authors concluded that the association between the appearance of extracellular amyloid fibrils and apoptosis implied causality, i.e., that the extracellular amyloid fibrils had induced apoptosis. However, in subsequent studies other investigators were unable to support this conclusion (42,52). When amyloid fibrils were added to islet cells in culture, apoptosis was not induced and electron microscopy revealed viable cells decorated with amyloid. In contrast, if a freshly prepared aqueous solution of hIAPP was added to islet cells in culture apoptosis was reproducibly induced, and under these circumstances electron microscopy revealed the presence of small nonfibrillar hIAPP oligomers, apparently disrupting the cell membrane, and indeed penetrating the cell (42). These data together with those from the hIAPP transgenic mice implied that it was not the amyloid fibrils that induce apoptosis, but much smaller oligomers that form rapidly after free hIAPP monomers interact in an aqueous environment.

This impression was reinforced by studies of membrane bilayers (42) (Fig. 6). Freshly prepared solutions of hIAPP induced nonselective ion channels and ultimately disrupted the membranes, whereas neither amyloid fibrils nor freshly prepared solutions of rIAPP had any discernible effect. Dissociation between formation and actions of hIAPP toxic oligomers vs. hIAPP-derived amyloid was further demonstrated by use of rifampicin (153). Rifampicin, as previously reported, inhibited hIAPP amyloid formation but failed to inhibit formation of either hIAPP toxic oligomers or hIAPP cytotoxicity. These data established the fact that hIAPP toxic oligomers are not simply “pre-amyloid” fibrils or protofibrils, but are an off-amyloid fibril pathway form of oligomer (Fig. 7). This finding was already implicit in transgenic rodent studies where β-cell toxicity was unrelated to the extent or location of amyloid formation (61). If hIAPP toxic oligomer formation is off the fibril pathway, then inhibition of amyloid formation may not only fail to prevent toxicity of amyloidogenic proteins, but may even promote formation of toxic oligomers and enhance toxicity.

Figure 6.

Figure 6

Stability of planar bilayer membranes is disrupted by addition of hIAPP. Control recording of bilayer capacitance and the same membrane 5 min after adding 10 μmol/liter freshly dissolved hIAPP to the cis chamber. Note membrane instability and increase in membrane electrical noise. Filtered at 0.3 kHz. Adapted from Ref. 42. [Copyright 1999 American Diabetes Association. From Diabetes 48:491–498. Reprinted with permission from The American Diabetes Association.]

Figure 7.

Figure 7

Proposed model for the balance between the different aggregation states of hIAPP in an aqueous solution. Once hIAPP oligomers are dissolved in an aqueous solution, IAPP intermediate structures (protofibrils) further assemble into amyloid fibrils. Alternatively, they may form toxic membrane-perforating toxic oligomers (top). In the presence of rifampicin, formation of amyloid fibrils is inhibited, but the formation of toxic oligomers is unaffected, consistent with continued cytotoxicity (bottom). This figure originally appeared in an article by J. J. Meier et al.: Am J Physiol Endocrinol Metab 291:E1317–E1324, 2006 (153). It is used with permission from the American Physiological Society.

Having established that small membrane permeant oligomers, but not amyloid fibrils, are the toxic form of hIAPP, the amyloid hypothesis can be modified rather than rejected. Restated, proteins with the propensity to form amyloid fibrils have the capacity to form membrane-permeant toxic oligomers. To date, the relationship between the propensity to form toxic oligomers and amyloid fibrils is not fully characterized. Moreover, the structure of toxic oligomers remains to be established. Our own interest with respect to hIAPP toxic oligomers reverted to the questions, do toxic hIAPP oligomers form intra- or extracellularly, and in vivo do they act intra- or extracellularly to induce apoptosis?

B. Where do hIAPP toxic oligomers form and act?

A recent breakthrough in this field came from the laboratory of Charles Glabe at the University of California, Irvine (68). Glabe’s group developed a method to reproducibly synthesize a molecular mimic of AβP toxic oligomers using a colloidal gold core (69). When this molecular mimic was injected into rabbits, antibodies were raised that bound to the toxic form of AβP but did not bind to monomers or amyloid fibrils of AβP (68). Unexpectedly, this antibody also binds to the toxic oligomeric form of hIAPP, prion, and synuclein but not to the monomers or the amyloid fibrils of these proteins. One implication of this intriguing landmark paper is that the tertiary structure of the toxic oligomers of hIAPP, AβP, synuclein, and prion must be remarkably similar even though they are composed of monomers of distinct proteins. Another implication is that it is now possible to use this antibody to identify the location (intra- vs. extracellular) of the toxic oligomers. A third implication is that the mechanisms subserving formation of these toxic oligomers and their mechanism of action to induce cytotoxicity are likely to be similar in these diseases, now often referred to as unfolded protein diseases (70).

By use of the toxic oligomer-specific antibody, we were able to establish that the toxic oligomers of hIAPP in hIAPP transgenic mice form intracellularly (Fig. 8) (71). Staining for toxic oligomers was not found in the extracellular islet amyloid. To establish whether the toxic oligomers act intra- or extracellularly, we used a vaccine approach. After vaccination of hIAPP transgenic mice with the same oligomer preparation used by the Glabe laboratory to raise the antitoxic oligomer antibody, high titers of antioligomer antibodies developed in hIAPP transgenic mice. However, β-cell apoptosis was not decreased. By implication, toxic hIAPP oligomers in hIAPP transgenic mice form and act intracellularly in vivo.

Figure 8.

Figure 8

hIAPP toxic oligomers in islets from obese hemizygous hIAPP transgenic mice. hIAPP toxic oligomer immunoreactivity does not coincide with extracellular amyloid; it is predominantly intracellular, confined to β-cells, and it is perinuclear or in vesicle-like structures. Immunofluorescent staining for toxic hIAPP oligomers (red), autofluorescence for amyloid (green), and nuclei 4′,6-diamidino-2-phenylindole (DAPI) (blue) (A, 20× magnification). Immunofluorescent staining for toxic hIAPP oligomers (red), insulin (green), and nuclei DAPI (blue) (B, 100× magnification). Adapted from Ref. 71. [Copyright 2007 American Diabetes Association. From Diabetes 56:1324–1332. Reprinted with permission from The American Diabetes Association.]

On a technical note, to date the antibodies for detection of toxic oligomers are ineffective in paraffin-embedded tissue, requiring frozen sections. This is not surprising because they detect the shared structure of toxic oligomers rather than a particular amino acid sequence in the monomers that compose them, a structure likely disturbed by tissue processing. Therefore, to date there are no data in human pancreas from patients with T2DM to confirm that the toxic hIAPP oligomers present in transgenic mice are also present in T2DM. Comparable intracellular IAPP oligomers were noted intracellularly in human β-cells from surgically resected insulinoma (72).

Although hIAPP oligomers form intracellularly in mice with high transgenic expression rates of hIAPP, not all β-cells even in these models have detectable hIAPP oligomers (71). These data imply that at any given time most β-cells in these mouse models have mechanisms in place that prevent hIAPP oligomer formation. Because hIAPP monomers rapidly form toxic oligomers in an aqueous solution, to better understand why these oligomers form in some β-cells, it seems prudent to consider the cellular mechanisms that prevent formation of hIAPP oligomers.

IV. Cellular Mechanisms to Prevent hIAPP Oligomer Formation. The Unfolded Protein Response (UPR), proIAPP Processing, and Vesicle Environment

The endoplasmic reticulum (ER) is responsible for the synthesis, folding, and appropriate targeting of all client secretory proteins before their export to the Golgi (most prominently insulin and IAPP in β-cells). The ER has several important properties to facilitate protein folding. These include a Ca2+ concentration of approximately 300 μm (vs. 0.1 μm in the cytosol) (73,74), a relatively oxidative state favoring disulfide bond formation by protein sulfide isomerases (75), and a protein quality control system (76,77). Unfolded proteins are exported from the ER by retrograde translocation to the cytosol and degradation by the proteosome (78,79). In addition, the ER contains abundant chaperone proteins that shield hydrophobic regions of unfolded proteins from surrounding proteins (80,81,82,83). Chaperone protein binding has been shown in vitro to inhibit AβP oligomerization (84). Given that the ER protein concentration is approximately 100 g/liter, these properties are remarkably successful at preventing ER protein aggregation. This is particularly the case for proteins such as hIAPP that are highly prone to form self-aggregates at much lower concentrations in an aqueous environment (85).

In addition to these properties of the ER, the unfolded protein response (UPR) balances ER protein delivery with the capacity of the ER to fold and traffic these proteins [Refs. 86,87,88,89,90; also see article by Scheuner and Kaufman in this issue of Endocrine Reviews (152)] (Fig. 9). By doing so the UPR defends the ER from being overwhelmed by misfolded and more importantly aggregated proteins that may lead to ER stress and apoptosis (91). Three independent proteins, PERK (protein kinase-like ER kinase), IRE1α (inositol requiring 1α), and ATF6 (activating transcription factor 6) detect increased abundance of unfolded proteins in ER and activate a sequence of events that globally decreases translation of major ER client proteins, increases transcription and translation of ER chaperone proteins [e.g., binding Ig protein (BiP)], and increases expression of proteins involved in clearance of unfolded ER proteins (88,92,93). The importance of PERK in the protection of β-cells was illustrated by the development of diabetes due to increased β-cell apoptosis in the PERK −/− mouse (94).

Figure 9.

Figure 9

The schematic illustration of the UPR in protein secretory cells. Increased demand for BiP leads to detachment of BiP from PERK, IRE1α, and ATF6, which get activated. Activated PERK phosphorylates α-subunit of eukaryotic translation initiation factor 2 (eIF2α), which subsequently suppresses ER protein translation and leads to more ATF4. Activated IRE1α has a RNA editing function and removes the hairpin structure on inactive X-box-binding protein-1 (XBP1) mRNA (unspliced, u-XBP1), which later becomes active transcription factor (spliced, s-XBP1). Activated ATF6 translocates into the Golgi and undergoes partial intramembrane proteolysis by site-1 protease (S1P) and site-2 protease (S2P), then migrates to the nucleus. These three activated transcriptional factors then induce a series of responses to increase chaperone proteins, limit new protein translation, and increase the degradation of unfolded proteins (see Fig. 10 and Refs. 65,83,86,123 and 131).

In summary, the UPR allows secretory cells such as the β-cell to balance ER delivery of major client proteins (insulin and IAPP in the β-cell) to the capacity of the ER to fold and traffic these proteins to the Golgi and secretory vesicles. Because expression of IAPP increases disproportionately to insulin under conditions of insulin demand (95), under these conditions, IAPP competes for ER resources and presumably constrains the maximal insulin synthesis rate. Theoretically, this can be overcome by increasing the number of β-cells. Although this adaptation appears readily available in mice (61), in adult humans there is a limited capacity for β-cell replication (6). Therefore, under conditions of sustained insulin resistance (e.g., obesity), the ER in human β-cells will be in a state of prolonged high demand. Any additional stress to the β-cell under these conditions that adversely influences ER function (e.g., oxidative damage, ER Ca2+ depletion, ER membrane leakage) will readily disturb this balance, potentially leading to ER stress-induced apoptosis (96,97).

In common with the ER, the insulin secretory vesicles presumably sustain hIAPP concentrations that far exceed the solubility of hIAPP in a typical aqueous environment. A property of insulin secretory vesicles that is likely important in preventing IAPP oligomer formation is the acid pH of the vesicle lumen (85,98,99). At the pH 5.5 present in insulin secretory vesicles, hIAPP is maintained in monomers (99). In addition, insulin interacts with hIAPP to reduce oligomer formation (85). It is likely that there are other factors (chaperone proteins, ions) that restrain oligomer formation in insulin vesicles.

In conclusion, the synthesis and trafficking of nascent client secretory proteins by the ER is closely regulated to prevent oligomer formation. Likewise, the properties of the lumen of the insulin secretory vesicles favor maintaining hIAPP in a monomeric form. Because hIAPP oligomers apparently form intracellularly, these protective mechanisms presumably fail under those circumstances.

A. Why do hIAPP oligomers (and amyloid) form?

Given the potent mechanisms in place to prevent intracellular oligomerization of amyloidogenic proteins such as hIAPP, why does this fail in T2DM? Mice transgenic for hIAPP provide some insights. First, the increased risk of hIAPP oligomerization with increasing hIAPP expression implies that the mechanisms that protect against hIAPP aggregation and toxicity are saturable. This is consistent with the observation that circumstances that increase expression of IAPP per β-cell in humans increase risk for developing T2DM.

Thus insulin resistance (which disproportionately increases IAPP compared with insulin expression) (100,101) is a major risk factor for T2DM. Also, a 50% decrease in β-cell mass (which doubles the secretory demand per β-cell) often leads to subsequent diabetes in dogs (15) and humans (14,16), both of which express an amyloidogenic form of IAPP, but not rats that secrete a soluble form of IAPP (102) (Fig. 4). Recent genome-wide linkage studies show linkage between risk for T2DM and several cell cycle transcriptional regulatory proteins (3). This together with the wide range of β-cell mass observed in nondiabetic humans (Fig. 1) raises the possibility that a relatively low adult β-cell mass might serve as a risk factor for T2DM. Under these circumstances, insulin resistance would place a substantial increased demand per β-cell in adult humans and, presumably, a greater risk for hIAPP expression rates that exceed the threshold for trafficking hIAPP in a soluble form.

Another potential cause of IAPP oligomer formation despite the protective mechanisms against it would be expression of a mutant hIAPP that increases the propensity for hIAPP to form oligomers (103). Although a rare cause of T2DM, the S20G mutation (Fig. 4) reported in Japan meets this criterion (40).

One other potential mechanism that would increase the risk for hIAPP toxic oligomer formation would be a decrease in the capacity of β-cells to neutralize toxic oligomers as they form. Insulin degrading enzyme (IDE) has been reported to have this property, and therefore it is of note that the IDE gene shows linkage to both T2DM and Alzheimer’s disease (3,104). IDE has been shown in vitro to inhibit hIAPP (and AβP) aggregate formation and cytotoxicity (105,106). Polymorphisms in chaperone proteins important in trafficking hIAPP are an obvious candidate for increased propensity to form hIAPP oligomers. In this regard it is of interest that hIAPP and AβP share close structural properties and that the prevalence of Alzheimer’s disease is increased in people with T2DM (107).

Also, any factors, inherited or acquired, that disturb the function of the ER might reasonably be expected to increase risk for hIAPP oligomer formation. Compromised ER function leads to mitochondrial dysfunction (108). Because ER function requires high energy, it is reasonable to expect that compromised mitochondrial function might lead to ER dysfunction. Therefore mitochondrial dysfunction in β-cells in T2DM (109) might be expected to lead to increased risk of hIAPP oligomer formation.

Factors in the secretory pathway and vesicle environment might also contribute to risk for oligomer formation. In cystic fibrosis, acidification of intracellular vesicles is impaired, and it is therefore of interest that a high proportion of patients with cystic fibrosis develop T2DM with islet amyloid (110). Also, impaired hIAPP processing, a function of the insulin secretory pathway, has been implicated as a predisposing factor to hIAPP amyloid formation (111). This might be a factor in the tendency for islet amyloid formation in human insulinoma (72).

In conclusion, there are several potential factors already well recognized as associated with T2DM that might lead to an increased risk for hIAPP oligomerization. From the above discussion, it can be appreciated that the formation of toxic hIAPP oligomers might occur as a consequence or as a cause of β-cell failure. Because hIAPP oligomers are cytotoxic, it would seem reasonable to predict that they contribute to the progression of β-cell failure in T2DM, if they form secondary to islet dysfunction. Given the probable multiple causes of the clinical syndrome of T2DM, hIAPP oligomers likely form and contribute early in some forms of T2DM and late in others, depending on the underlying mechanisms that initiated β-cell dysfunction.

B. hIAPP-induced β-cell apoptosis

Apoptosis was first used to describe cell death with specific morphological characteristics (cell shrinkage, nuclear condensation, chromatin margination, and clumping and blebbing of the cell surface) (112) and then on the presence of free 3-OH strand breaks in DNA by the TUNEL (terminal deoxynucleotidyl transferase-mediated dUTP nick-end labeling) method (113). Subsequently specific biochemical pathways were identified that initiate and execute this form of cell death (114). Furthermore, apoptosis was found to be an evolutionarily conserved means of tissue remodeling during normal development (115). In adult tissue, apoptosis continues to play a role in regulated cell turnover (116). Induction of high rates of apoptosis has been identified as the underlying mechanism leading to loss of cell mass in neurodegenerative diseases, e.g., Alzheimer’s disease (117,118) and Parkinson’s disease (28).

Two major pathways of apoptosis are the extrinsic pathway (119) and the intrinsic pathway, which includes the ER stress pathway (91,96,97,120,121,122,123). The extrinsic pathway is classically mediated by binding of death signals (Fas ligands) to death receptors (Fas) on the cell surface leading to aggregation of Fas receptors, caspase-8 and Fas-associated death domain protein into a death-inducing signaling complex in which caspase-8 is proteolytically activated and then released (119). Activated caspase-8 can then either directly activate the execution phase of apoptosis (via caspase-3) or amplify its signal by proteolytic activation of the proapoptotic member of the Bcl-2 family Bid leading to subsequent release of mitochondrial proapoptotic factors (e.g., cytochrome c). The extrinsic pathway of apoptosis is active in autoimmune-mediated β-cell death (124,125) and has also been invoked as a mediator of β-cell glucose toxicity (type-1 and type-2 diabetes) through the actions of high glucose concentration to induce expression of the Fas ligand and IL1-β in β-cells (126,127).

The intrinsic pathway of apoptosis is mediated through a number of cell stresses. In the context of degenerative diseases, documented inducers of the intrinsic pathway include ER stress, mitochondrial dysfunction, generation of oxygen free radicals, metabolic toxins, disruption of the actin cytoskeleton, and anoxia (97,128).

The ER stress pathway of apoptosis (Fig. 10) is a mechanism of apoptosis to which cells with a high secretory burden (such as β-cells) are particularly vulnerable (65,91,94,123,129,130,131). The primary defense mechanism against ER stress-induced apoptosis is the UPR (86,87,88,89,90). For clarity in this article, we define ER stress as the circumstances that provoke induction of apoptosis as a consequence of the accumulation of aggregated proteins. We distinguish this from the UPR that we define as the adaptive efforts by the cell described above to prevent ER stress. This is in distinction to some publications in this recently evolving field that use UPR and ER stress interchangeably. When the UPR is unable to clear the ER of unfolded, and especially aggregated proteins, ER stress by this definition may develop.

Figure 10.

Figure 10

Proposed molecular signaling pathways of hIAPP-induced ER stress and apoptosis in pancreatic β-cells. BiP shortage activates three transcriptional factors (ATF4, ATF6, and XBP1), which collectively launch the UPR. In hIAPP overexpressing models and under insulin resistance conditions, an increased number of proteins in the ER leads to molecular crowding, which promotes protein aggregation and misfolding, especially of the amyloidogenic protein like hIAPP. Aggregated or unfolded hIAPP can compromise ER membrane barriers for ionic calcium. Decreased calcium inside the ER lumen and increased calcium in the cytosol may lead to ER stress, which is represented by nuclear translocation of CHOP, induction of death receptor DR5, down-regulation of BCL-2, cleavage of caspase-12, and accumulation of ubiquitinated proteins. Decreased ER calcium will decrease the efficiency of protein folding machinery and result in more unfolded proteins. Increased calcium in the cytosol may open up the mitochondrial permeability transition pore (Mito PTP), leading to cytochrome c release and caspase-9 activation (64,84,87,123,131).

Two distinct circumstances can be anticipated that might lead to ER stress. In one, mutant proteins are synthesized with the property of an increased propensity to oligomerize. The common mutations of the cystic fibrosis gene (132), the Huntington’s gene (133), some known mutations of the AβP gene (27,29), and IAPP (40,103) leading to familial Alzheimer’s disease or T2DM have this property. The Akita mouse model of diabetes has a mutation in the insulin gene leading to ER aggregation of insulin, ER stress-induced β-cell apoptosis, and diabetes (123). In humans, a rare mutation in IAPP that increases the propensity of IAPP to oligomerize is linked to a familial form of T2DM (40,103). The other more common circumstance is ER overload, the expression of an oligomeric protein such as hIAPP at a rate that exceeds the ER capacity to fold and traffic the protein. This fits with the known risk factors for T2DM including insulin resistance and a deficit in β-cell mass (for example after partial pancreatectomy) (13).

Marchetti et al. (134) showed increased markers of ER stress in isolated islets from patients with T2DM. Interestingly, β-cells showed modest signs of ER stress when the islets were cultured at normal glucose, but increased when the islets were cultured at higher glucose. This finding implies a genetic predisposition in islets from individuals with T2DM to ER stress when β-cells are chronically stimulated, a predisposition absent in islets of nondiabetic individuals. These data are consistent with the notion of a lower capacity to traffic and fold major client secretory proteins, such as IAPP in individuals with T2DM. We have previously reported that ER stress is characteristic of β-cells in humans with type 2 diabetes but interestingly not in type 1 diabetes (Fig. 11B) (65).

Figure 11.

Figure 11

hIAPP induces ER stress responses in pancreatic β-cells. A, Immunoblotting of UPR and ER stress markers in islets from wild-type, rIAPP and hIAPP transgenic mice (r-TG and h-TG, respectively), and INS cells overexpressing GFP, rIAPP-EGFP, or hIAPP-EGFP. B, Nuclear translocation of CHOP (a–c); increased caspase-12 expression (d, e), and accumulation of polyubiquitinated proteins (Ubi-P, f) in islets from hIAPP transgenic mouse (a, d, f), hIAPP transgenic rat (HIP rat; b, e), and human islets from an obese T2DM subject (f). C, Nuclear CHOP is colocalized with the appearance of TUNEL staining in INS cells overexpressing hIAPP-EGFP. D, Knockdown of CHOP by small interfering RNA reduces the cleavage of caspase-3 in INS cells overexpressing hIAPP. This figure is adapted from the original figure that appeared in an article by C. J. Huang et al. (65,131). It is used with permission from The American Physiological Society. [Copyright 2007 American Diabetes Association. From Diabetes 56:2016–2027. Reprinted with permission from The American Diabetes Association.]

The exact mechanism linking protein oligomer formation and ER stress-induced apoptosis is unknown. One proposed mechanism is that toxic oligomers interact with the ER membrane leading to Ca2+ leakage (74,135). This might directly lead to mitochondrial membrane permeability, leakage of cytochrome c, and activation of executioner caspases (caspase-3), as well as indirectly contributing to apoptosis by increasing the number of unfolded proteins in the ER due to depletion of ER Ca2+. Ca2+ leakage from the ER can also activate ER-associated calpain, which can then directly induce apoptosis in a caspase-independent manner (136,137). In rodents, activation of the ER membrane resident caspase-12 is associated with induction of the ER stress pathway of apoptosis, although it is not clear whether caspase-12 activation is a consequence of, or a mechanism contributing to the ER stress-induced pathway of apoptosis (31,138,139). Caspase-12 expression was detected in hIAPP transgenic mice and rats, but not in rIAPP transgenic mice (Fig. 11B). Caspase-4 appears to have the same properties (activated by chronic ER stress) in humans (117,140).

ER stress has been identified as an important mechanism inducing apoptosis in Alzheimer’s disease (117,118,141), Parkinson’s disease (29,142), and T2DM (65,123,131), all three of which share the characteristic of increased apoptosis in relation to protein misfolding of amyloidogenic proteins (Table 1). Therefore, ER stress is obviously a strong candidate for mediating hIAPP oligomer-induced apoptosis.

ER stress has been observed in β-cell lines transduced with hIAPP as well as mice and rats transgenic for hIAPP (65,131) (Fig. 11A). Nuclear C/EBP homologous protein/GADD153 (CHOP) staining was found in hIAPP, but not rIAPP transgenic mice, and also in pancreatic section of a T2DM subject (Fig. 11B). The appearance of CHOP preceded execution of apoptosis as measured by TUNEL. We occasionally observed concordant nuclear and TUNEL in hIAPP-expressing INS cells (Fig. 11C). Furthermore, when we knocked down CHOP by small interfering RNA, apoptosis was decreased (Fig. 11D). Application of hIAPP oligomers extracellularly has also been shown to impair the ubiquitin proteasomal pathway (65). The accumulation of polyubiquitinated proteins was also identified in hIAPP, but not rIAPP transgenic mice (Fig. 11B). Because hIAPP oligomers induce membrane leakage and disruption, application of these oligomers extracellularly or formation of them within the secretory pathway intracellularly might reasonably be expected to permit Ca2+ influx into cytoplasm, a known signal to induce the intrinsic pathway of apoptosis. Moreover, ER stress has been shown to induce expression of death receptor (143), potentially invoking the extrinsic pathway of apoptosis so that in reality, both classical pathways of apoptosis will likely be active once cell membranes have been disrupted. Furthermore, a recent study showed that addition of hIAPP to the cells induces β-cell apoptosis through Fas-associated death receptor (144). Consistent with this, activation of p38 MAPK and JNK1 has also been noted after application of hIAPP oligomers to cells extracellularly (145,146,147).

Islet amyloid has also been noted in islets from hIAPP transgenic mice (56) and human islets either in culture at high glucose (148,149) or after transplantation (150). It is not known why transplanted islets, or islets in culture, have increased hIAPP amyloid formation, but loss of vasculature might be a predisposing factor. Also, isolated islets and transplanted islets are relatively anoxic (151), with mitochondrial dysfunction likely leading to decreased ER function. hIAPP oligomer formation might be a contributory factor to early β-cell loss after islet transplantation.

V. Summary

The islet in T2DM shares much in common with neuropathology in neurodegenerative diseases such as Alzheimer’s disease. Because measurement of β-cell dysfunction is more sensitive and less demanding than cognitive function, hIAPP transgenic models are an appealing means to advance an understanding of both fields. Most of the focus in neurodegenerative diseases is now focused on protein misfolding and aggregation, the diseases now often referred to as unfolded protein diseases. Important questions that remain to be answered include, why do toxic oligomers of amyloidogenic proteins form; what is the precise structure of these oligomers; and are there therapeutic approaches that can prevent their formation or toxicity? Once these questions are addressed, the importance of the toxic oligomer hypothesis will be better defined.

Supplementary Material

[RPHR Note]

Footnotes

P.C.B. is supported by grants from the National Institutes of Health (DK0595791), the Larry L. Hillblom Foundation (2003/1B), and the Juvenile Diabetes Research Foundation (7-2005-1152).

Disclosure Statement: The authors have nothing to disclose.

First Published Online February 27, 2008

Abbreviations: AβP, Amyloid β protein; ATF6, activating transcription factor 6; BiP, binding Ig protein; CHOP, C/EBP homologous protein/GADD153; ER, endoplasmic reticulum; hIAPP, human IAPP; IAPP, islet amyloid polypeptide; IDE, insulin degrading enzyme; IRE1α, inositol requiring 1α; PERK, protein kinase-like ER kinase; rIAPP, rodent IAPP; T2DM, type 2 diabetes; TUNEL, terminal deoxynucleotidyl transferase-mediated dUTP nick-end labeling; UPR, unfolded protein response.

References

  1. Zimmet P, Alberti KG, Shaw J 2001 Global and societal implications of the diabetes epidemic. Nature 414:782–787 [DOI] [PubMed] [Google Scholar]
  2. Zeggini E, Weedon MN, Lindgren CM, Frayling TM, Elliott KS, Lango H, Timpson NJ, Perry JR, Rayner NW, Freathy RM, Barrett JC, Shields B, Morris AP, Ellard S, Groves CJ, Harries LW, Marchini JL, Owen KR, Knight B, Cardon LR, Walker M, Hitman GA, Morris AD, Doney AS, McCarthy MI, Hattersley AT 2007 Replication of genome-wide association signals in UK samples reveals risk loci for type 2 diabetes. Science 316:1336–1341 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Sladek R, Rocheleau G, Rung J, Dina C, Shen L, Serre D, Boutin P, Vincent D, Belisle A, Hadjadj S, Balkau B, Heude B, Charpentier G, Hudson TJ, Montpetit A, Pshezhetsky AV, Prentki M, Posner BI, Balding DJ, Meyre D, Polychronakos C, Froguel P 2007 A genome-wide association study identifies novel risk loci for type 2 diabetes. Nature 445:881–885 [DOI] [PubMed] [Google Scholar]
  4. Jack Jr L, Boseman L, Vinicor F 2004 Aging Americans and diabetes. A public health and clinical response. Geriatrics 59:14–17 [PubMed] [Google Scholar]
  5. Polonsky KS 2000 Dynamics of insulin secretion in obesity and diabetes. Int J Obes Relat Metab Disord 24(Suppl 2):S29–S31 [DOI] [PubMed] [Google Scholar]
  6. Butler AE, Janson J, Bonner-Weir S, Ritzel R, Rizza RA, Butler PC 2003 β-Cell deficit and increased β-cell apoptosis in humans with type 2 diabetes. Diabetes 52:102–110 [DOI] [PubMed] [Google Scholar]
  7. Kloppel G, Lohr M, Habich K, Oberholzer M, Heitz PU 1985 Islet pathology and the pathogenesis of type 1 and type 2 diabetes mellitus revisited. Surv Synth Pathol Res 4:110–125 [DOI] [PubMed] [Google Scholar]
  8. Drueke TB 2000 Cell biology of parathyroid gland hyperplasia in chronic renal failure. J Am Soc Nephrol 11:1141–1152 [DOI] [PubMed] [Google Scholar]
  9. Gepts W, Lecompte PM 1981 The pancreatic islets in diabetes. Am J Med 70:105–115 [DOI] [PubMed] [Google Scholar]
  10. Yoon KH, Ko SH, Cho JH, Lee JM, Ahn YB, Song KH, Yoo SJ, Kang MI, Cha BY, Lee KW, Son HY, Kang SK, Kim HS, Lee IK, Bonner-Weir S 2003 Selective β-cell loss and α-cell expansion in patients with type 2 diabetes mellitus in Korea. J Clin Endocrinol Metab 88:2300–2308 [DOI] [PubMed] [Google Scholar]
  11. Ritzel RA, Butler AE, Rizza RA, Veldhuis JD, Butler PC 2006 Relationship between β-cell mass and fasting blood glucose concentration in humans. Diabetes Care 29:717–718 [DOI] [PubMed] [Google Scholar]
  12. Kjems LL, Kirby BM, Welsh EM, Veldhuis JD, Straume M, McIntyre SS, Yang D, Lefebvre P, Butler PC 2001 Decrease in β-cell mass leads to impaired pulsatile insulin secretion, reduced postprandial hepatic insulin clearance, and relative hyperglucagonemia in the minipig. Diabetes 50:2001–2012 [DOI] [PubMed] [Google Scholar]
  13. Matveyenko AV, Veldhuis JD, Butler PC 2006 Mechanisms of impaired fasting glucose and glucose intolerance induced by an approximate 50% pancreatectomy. Diabetes 55:2347–2356 [DOI] [PubMed] [Google Scholar]
  14. Robertson RP, Lanz KJ, Sutherland DE, Seaquist ER 2002 Relationship between diabetes and obesity 9 to 18 years after hemipancreatectomy and transplantation in donors and recipients. Transplantation 73:736–741 [DOI] [PubMed] [Google Scholar]
  15. Stagner JI, Samols E 1991 Deterioration of islet β-cell function after hemipancreatectomy in dogs. Diabetes 40:1472–1479 [DOI] [PubMed] [Google Scholar]
  16. Lee BW, Kang HW, Heo JS, Choi SH, Kim SY, Min YK, Chung JH, Lee MK, Lee MS, Kim KW 2006 Insulin secretory defect plays a major role in the development of diabetes in patients with distal pancreatectomy. Metabolism 55:135–141 [DOI] [PubMed] [Google Scholar]
  17. Ehrlich JC, Ratner IM 1961 Amyloidosis of the islets of Langerhans. A restudy of islet hyalin in diabetic and non-diabetic individuals. Am J Pathol 38:49–59 [PMC free article] [PubMed] [Google Scholar]
  18. Margittai M, Langen R 2006 Spin labeling analysis of amyloids and other protein aggregates. Methods Enzymol 413:122–139 [DOI] [PubMed] [Google Scholar]
  19. Cooper GJ, Willis AC, Clark A, Turner RC, Sim RB, Reid KB 1987 Purification and characterization of a peptide from amyloid-rich pancreases of type 2 diabetic patients. Proc Natl Acad Sci USA 84:8628–8632 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Westermark P, Wernstedt C, O’Brien TD, Hayden DW, Johnson KH 1987 Islet amyloid in type 2 human diabetes mellitus and adult diabetic cats contains a novel putative polypeptide hormone. Am J Pathol 127:414–417 [PMC free article] [PubMed] [Google Scholar]
  21. Butler PC, Chou J, Carter WB, Wang YN, Bu BH, Chang D, Chang JK, Rizza RA 1990 Effects of meal ingestion on plasma amylin concentration in NIDDM and nondiabetic humans. Diabetes 39:752–756 [DOI] [PubMed] [Google Scholar]
  22. Kahn SE, D’Alessio DA, Schwartz MW, Fujimoto WY, Ensinck JW, Taborsky Jr GJ, Porte Jr D 1990 Evidence of cosecretion of islet amyloid polypeptide and insulin by β-cells. Diabetes 39:634–638 [DOI] [PubMed] [Google Scholar]
  23. Leffert JD, Newgard CB, Okamoto H, Milburn JL, Luskey KL 1989 Rat amylin: cloning and tissue-specific expression in pancreatic islets. Proc Natl Acad Sci USA 86:3127–3130 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Hardy J, Selkoe DJ 2002 The amyloid hypothesis of Alzheimer’s disease: progress and problems on the road to therapeutics. Science 297:353–356 [DOI] [PubMed] [Google Scholar]
  25. Tanzi RE, Bertram L 2005 Twenty years of the Alzheimer’s disease amyloid hypothesis: a genetic perspective. Cell 120:545–555 [DOI] [PubMed] [Google Scholar]
  26. Gotz J, Streffer JR, David D, Schild A, Hoerndli F, Pennanen L, Kurosinski P, Chen F 2004 Transgenic animal models of Alzheimer’s disease and related disorders: histopathology, behavior and therapy. Mol Psychiatry 9:664–683 [DOI] [PubMed] [Google Scholar]
  27. Chartier-Harlin MC, Crawford F, Houlden H, Warren A, Hughes D, Fidani L, Goate A, Rossor M, Roques P, Hardy J, Mullan M 1991 Early-onset Alzheimer’s disease caused by mutations at codon 717 of the β-amyloid precursor protein gene. Nature 353:844–846 [DOI] [PubMed] [Google Scholar]
  28. Ryu EJ, Harding HP, Angelastro JM, Vitolo OV, Ron D, Greene LA 2002 Endoplasmic reticulum stress and the unfolded protein response in cellular models of Parkinson’s disease. J Neurosci 22:10690–10698 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Imai Y, Soda M, Inoue H, Hattori N, Mizuno Y, Takahashi R 2001 An unfolded putative transmembrane polypeptide, which can lead to endoplasmic reticulum stress, is a substrate of Parkin. Cell 105:891–902 [DOI] [PubMed] [Google Scholar]
  30. Simoneau S, Rezaei H, Sales N, Kaiser-Schulz G, Lefebvre-Roque M, Vidal C, Fournier JG, Comte J, Wopfner F, Grosclaude J, Schatzl H, Lasmezas CI 2007 In vitro and in vivo neurotoxicity of prion protein oligomers. PLoS Pathog 3:e125 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Wootz H, Hansson I, Korhonen L, Napankangas U, Lindholm D 2004 Caspase-12 cleavage and increased oxidative stress during motoneuron degeneration in transgenic mouse model of ALS. Biochem Biophys Res Commun 322:281–286 [DOI] [PubMed] [Google Scholar]
  32. Chen S, Berthelier V, Hamilton JB, O’Nuallain B, Wetzel R 2002 Amyloid-like features of polyglutamine aggregates and their assembly kinetics. Biochemistry 41:7391–7399 [DOI] [PubMed] [Google Scholar]
  33. Lue LF, Kuo YM, Roher AE, Brachova L, Shen Y, Sue L, Beach T, Kurth JH, Rydel RE, Rogers J 1999 Soluble amyloid β peptide concentration as a predictor of synaptic change in Alzheimer’s disease. Am J Pathol 155:853–862 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. McLean CA, Cherny RA, Fraser FW, Fuller SJ, Smith MJ, Beyreuther K, Bush AI, Masters CL 1999 Soluble pool of Aβ amyloid as a determinant of severity of neurodegeneration in Alzheimer’s disease. Ann Neurol 46:860–866 [DOI] [PubMed] [Google Scholar]
  35. Clark A, Nilsson MR 2004 Islet amyloid: a complication of islet dysfunction or an aetiological factor in type 2 diabetes? Diabetologia 47:157–169 [DOI] [PubMed] [Google Scholar]
  36. Dickson DW, Crystal HA, Bevona C, Honer W, Vincent I, Davies P 1995 Correlations of synaptic and pathological markers with cognition of the elderly. Neurobiol Aging 16:285–298; discussion, 298–304 [DOI] [PubMed] [Google Scholar]
  37. Westermark P 1976 The nature of amyloid in islets of Langerhans in old age. New York: Academic Press [Google Scholar]
  38. Westermark P 1973 Fine structure of islets of Langerhans in insular amyloidosis. Virchows Arch A Pathol Anat 359:1–18 [DOI] [PubMed] [Google Scholar]
  39. Matveyenko AV, Butler PC 2006 Islet amyloid polypeptide (IAPP) transgenic rodents as models for type 2 diabetes. ILAR J 47:225–233 [DOI] [PubMed] [Google Scholar]
  40. Seino S 2001 S20G mutation of the amylin gene is associated with type II diabetes in Japanese. Study Group of Comprehensive Analysis of Genetic Factors in Diabetes Mellitus. Diabetologia 44:906–909 [DOI] [PubMed] [Google Scholar]
  41. Jayasinghe SA, Langen R 2007 Membrane interaction of islet amyloid polypeptide. Biochim Biophys Acta 1768:2002–2009 [DOI] [PubMed] [Google Scholar]
  42. Janson J, Ashley RH, Harrison D, McIntyre S, Butler PC 1999 The mechanism of islet amyloid polypeptide toxicity is membrane disruption by intermediate-sized toxic amyloid particles. Diabetes 48:491–498 [DOI] [PubMed] [Google Scholar]
  43. Mirzabekov TA, Lin MC, Kagan BL 1996 Pore formation by the cytotoxic islet amyloid peptide amylin. J Biol Chem 271:1988–1992 [DOI] [PubMed] [Google Scholar]
  44. Kawahara M, Kuroda Y, Arispe N, Rojas E 2000 Alzheimer’s β-amyloid, human islet amylin, and prion protein fragment evoke intracellular free calcium elevations by a common mechanism in a hypothalamic GnRH neuronal cell line. J Biol Chem 275:14077–14083 [DOI] [PubMed] [Google Scholar]
  45. Leighton B, Cooper GJ 1988 Pancreatic amylin and calcitonin gene-related peptide cause resistance to insulin in skeletal muscle in vitro. Nature 335:632–635 [DOI] [PubMed] [Google Scholar]
  46. Ohsawa H, Kanatsuka A, Yamaguchi T, Makino H, Yoshida S 1989 Islet amyloid polypeptide inhibits glucose-stimulated insulin secretion from isolated rat pancreatic islets. Biochem Biophys Res Commun 160:961–967 [DOI] [PubMed] [Google Scholar]
  47. Arnelo U, Permert J, Larsson J, Reidelberger RD, Arnelo C, Adrian TE 1997 Chronic low dose islet amyloid polypeptide infusion reduces food intake, but does not influence glucose metabolism, in unrestrained conscious rats: studies using a novel aortic catheterization technique. Endocrinology 138:4081–4085 [DOI] [PubMed] [Google Scholar]
  48. Sanke T, Bell GI, Sample C, Rubenstein AH, Steiner DF 1988 An islet amyloid peptide is derived from an 89-amino acid precursor by proteolytic processing. J Biol Chem 263:17243–17246 [PubMed] [Google Scholar]
  49. Westermark P, Engstrom U, Johnson KH, Westermark GT, Betsholtz C 1990 Islet amyloid polypeptide: pinpointing amino acid residues linked to amyloid fibril formation. Proc Natl Acad Sci USA 87:5036–5040 [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. O’Brien TD, Butler PC, Westermark P, Johnson KH 1993 Islet amyloid polypeptide: a review of its biology and potential roles in the pathogenesis of diabetes mellitus. Vet Pathol 30:317–332 [DOI] [PubMed] [Google Scholar]
  51. Lorenzo A, Razzaboni B, Weir GC, Yankner BA 1994 Pancreatic islet cell toxicity of amylin associated with type-2 diabetes mellitus. Nature 368:756–760 [DOI] [PubMed] [Google Scholar]
  52. Konarkowska B, Aitken JF, Kistler J, Zhang S, Cooper GJ 2006 The aggregation potential of human amylin determines its cytotoxicity towards islet β-cells. FEBS J 273:3614–3624 [DOI] [PubMed] [Google Scholar]
  53. Couce M, Kane LA, O’Brien TD, Charlesworth J, Soeller W, McNeish J, Kreutter D, Roche P, Butler PC 1996 Treatment with growth hormone and dexamethasone in mice transgenic for human islet amyloid polypeptide causes islet amyloidosis and β-cell dysfunction. Diabetes 45:1094–1101 [DOI] [PubMed] [Google Scholar]
  54. Janson J, Soeller WC, Roche PC, Nelson RT, Torchia AJ, Kreutter DK, Butler PC 1996 Spontaneous diabetes mellitus in transgenic mice expressing human islet amyloid polypeptide. Proc Natl Acad Sci USA 93:7283–7288 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Soeller WC, Janson J, Hart SE, Parker JC, Carty MD, Stevenson RW, Kreutter DK, Butler PC 1998 Islet amyloid-associated diabetes in obese A(vy)/a mice expressing human islet amyloid polypeptide. Diabetes 47:743–750 [DOI] [PubMed] [Google Scholar]
  56. de Koning EJ, Morris ER, Hofhuis FM, Posthuma G, Hoppener JW, Morris JF, Capel PJ, Clark A, Verbeek JS 1994 Intra- and extracellular amyloid fibrils are formed in cultured pancreatic islets of transgenic mice expressing human islet amyloid polypeptide. Proc Natl Acad Sci USA 91:8467–8471 [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Yagui K, Yamaguchi T, Kanatsuka A, Shimada F, Huang CI, Tokuyama Y, Ohsawa H, Yamamura K, Miyazaki J, Mikata A, Yoshida S, Makino H 1995 Formation of islet amyloid fibrils in β-secretory granules of transgenic mice expressing human islet amyloid polypeptide/amylin. Eur J Endocrinol 132:487–496 [DOI] [PubMed] [Google Scholar]
  58. Hoppener JW, Oosterwijk C, Nieuwenhuis MG, Posthuma G, Thijssen JH, Vroom TM, Ahren B, Lips CJ 1999 Extensive islet amyloid formation is induced by development of type II diabetes mellitus and contributes to its progression: pathogenesis of diabetes in a mouse model. Diabetologia 42:427–434 [DOI] [PubMed] [Google Scholar]
  59. Butler AE, Jang J, Gurlo T, Carty MD, Soeller WC, Butler PC 2004 Diabetes due to a progressive defect in β-cell mass in rats transgenic for human islet amyloid polypeptide (HIP Rat): a new model for type 2 diabetes. Diabetes 53:1509–1516 [DOI] [PubMed] [Google Scholar]
  60. Wang F, Hull RL, Vidal J, Cnop M, Kahn SE 2001 Islet amyloid develops diffusely throughout the pancreas before becoming severe and replacing endocrine cells. Diabetes 50:2514–2520 [DOI] [PubMed] [Google Scholar]
  61. Butler AE, Janson J, Soeller WC, Butler PC 2003 Increased β-cell apoptosis prevents adaptive increase in β-cell mass in mouse model of type 2 diabetes: evidence for role of islet amyloid formation rather than direct action of amyloid. Diabetes 52:2304–2314 [DOI] [PubMed] [Google Scholar]
  62. Matveyenko AV, Butler PC 2006 β-cell deficit due to increased apoptosis in the human islet amyloid polypeptide transgenic (HIP) rat recapitulates the metabolic defects present in type 2 diabetes. Diabetes 55:2106–2114 [DOI] [PubMed] [Google Scholar]
  63. Ritzel RA, Butler PC 2003 Replication increases β-cell vulnerability to human islet amyloid polypeptide-induced apoptosis. Diabetes 52:1701–1708 [DOI] [PubMed] [Google Scholar]
  64. Lo D, Burkly LC, Widera G, Cowing C, Flavell RA, Palmiter RD, Brinster RL 1988 Diabetes and tolerance in transgenic mice expressing class II MHC molecules in pancreatic β cells. Cell 53:159–168 [DOI] [PubMed] [Google Scholar]
  65. Huang CJ, Haataja L, Gurlo T, Butler AE, Wu X, Soeller WC, Butler PC 2007 Induction of endoplasmic reticulum stress-induced cell apoptosis and accumulation of polyubiquitinated proteins by human islet amyloid polypeptide. Am J Physiol Endocrinol Metab 293:E1656–E1662 [DOI] [PubMed] [Google Scholar]
  66. Hull RL, Shen ZP, Watts MR, Kodama K, Carr DB, Utzschneider KM, Zraika S, Wang F, Kahn SE 2005 Long-term treatment with rosiglitazone and metformin reduces the extent of, but does not prevent, islet amyloid deposition in mice expressing the gene for human islet amyloid polypeptide. Diabetes 54:2235–2244 [DOI] [PubMed] [Google Scholar]
  67. Meier JJ, Butler AE, Galasso R, Rizza RA, Butler PC 2006 Increased islet β cell replication adjacent to intrapancreatic gastrinomas in humans. Diabetologia 49:2689–2696 [DOI] [PubMed] [Google Scholar]
  68. Kayed R, Head E, Thompson JL, McIntire TM, Milton SC, Cotman CW, Glabe CG 2003 Common structure of soluble amyloid oligomers implies common mechanism of pathogenesis. Science 300:486–489 [DOI] [PubMed] [Google Scholar]
  69. Kayed R, Glabe CG 2006 Conformation-dependent anti-amyloid oligomer antibodies. Methods Enzymol 413:326–344 [DOI] [PubMed] [Google Scholar]
  70. Glabe CG 2006 Common mechanisms of amyloid oligomer pathogenesis in degenerative disease. Neurobiol Aging 27:570–575 [DOI] [PubMed] [Google Scholar]
  71. Lin CY, Gurlo T, Kayed R, Butler AE, Haataja L, Glabe CG, Butler PC 2007 Toxic human islet amyloid polypeptide (h-IAPP) oligomers are intracellular, and vaccination to induce anti-toxic oligomer antibodies does not prevent h-IAPP-induced β-cell apoptosis in h-IAPP transgenic mice. Diabetes 56:1324–1332 [DOI] [PubMed] [Google Scholar]
  72. O’Brien TD, Butler AE, Roche PC, Johnson KH, Butler PC 1994 Islet amyloid polypeptide in human insulinomas. Evidence for intracellular amyloidogenesis. Diabetes 43:329–336 [DOI] [PubMed] [Google Scholar]
  73. Robert V, De Giorgi F, Massimino ML, Cantini M, Pozzan T 1998 Direct monitoring of the calcium concentration in the sarcoplasmic and endoplasmic reticulum of skeletal muscle myotubes. J Biol Chem 273:30372–30378 [DOI] [PubMed] [Google Scholar]
  74. Orrenius S, Zhivotovsky B, Nicotera P 2003 Regulation of cell death: the calcium-apoptosis link. Nat Rev Mol Cell Biol 4:552–565 [DOI] [PubMed] [Google Scholar]
  75. Hwang C, Sinskey AJ, Lodish HF 1992 Oxidized redox state of glutathione in the endoplasmic reticulum. Science 257:1496–1502 [DOI] [PubMed] [Google Scholar]
  76. Ellgaard L, Molinari M, Helenius A 1999 Setting the standards: quality control in the secretory pathway. Science 286:1882–1888 [DOI] [PubMed] [Google Scholar]
  77. Wormald MR, Dwek RA 1999 Glycoproteins: glycan presentation and protein-fold stability. Structure Fold Des 7:R155–R160 [DOI] [PubMed] [Google Scholar]
  78. Molinari M, Calanca V, Galli C, Lucca P, Paganetti P 2003 Role of EDEM in the release of misfolded glycoproteins from the calnexin cycle. Science 299:1397–1400 [DOI] [PubMed] [Google Scholar]
  79. Oda Y, Hosokawa N, Wada I, Nagata K 2003 EDEM as an acceptor of terminally misfolded glycoproteins released from calnexin. Science 299:1394–1397 [DOI] [PubMed] [Google Scholar]
  80. Baxter BK, James P, Evans T, Craig EA 1996 SSI1 encodes a novel Hsp70 of the Saccharomyces cerevisiae endoplasmic reticulum. Mol Cell Biol 16:6444–6456 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Hamilton TG, Norris TB, Tsuruda PR, Flynn GC 1999 Cer1p functions as a molecular chaperone in the endoplasmic reticulum of Saccharomyces cerevisiae. Mol Cell Biol 19:5298–5307 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Melnick J, Dul JL, Argon Y 1994 Sequential interaction of the chaperones BiP and GRP94 with immunoglobulin chains in the endoplasmic reticulum. Nature 370:373–375 [DOI] [PubMed] [Google Scholar]
  83. Argon Y, Simen BB 1999 GRP94, an ER chaperone with protein and peptide binding properties. Semin Cell Dev Biol 10:495–505 [DOI] [PubMed] [Google Scholar]
  84. Kudva YC, Hiddinga HJ, Butler PC, Mueske CS, Eberhardt NL 1997 Small heat shock proteins inhibit in vitro A β (1–42) amyloidogenesis. FEBS Lett 416:117–121 [DOI] [PubMed] [Google Scholar]
  85. Kudva YC, Mueske C, Butler PC, Eberhardt NL 1998 A novel assay in vitro of human islet amyloid polypeptide amyloidogenesis and effects of insulin secretory vesicle peptides on amyloid formation. Biochem J 331:809–813 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Kaufman RJ 2002 Orchestrating the unfolded protein response in health and disease. J Clin Invest 110:1389–1398 [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Lee AS 2005 The ER chaperone and signaling regulator GRP78/BiP as a monitor of endoplasmic reticulum stress. Methods 35:373–381 [DOI] [PubMed] [Google Scholar]
  88. Bertolotti A, Zhang Y, Hendershot LM, Harding HP, Ron D 2000 Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nat Cell Biol 2:326–332 [DOI] [PubMed] [Google Scholar]
  89. Allen JR, Nguyen LX, Sargent KE, Lipson KL, Hackett A, Urano F 2004 High ER stress in β-cells stimulates intracellular degradation of misfolded insulin. Biochem Biophys Res Commun 324:166–170 [DOI] [PubMed] [Google Scholar]
  90. Haynes CM, Titus EA, Cooper AA 2004 Degradation of misfolded proteins prevents ER-derived oxidative stress and cell death. Mol Cell 15:767–776 [DOI] [PubMed] [Google Scholar]
  91. Harding HP, Ron D 2002 Endoplasmic reticulum stress and the development of diabetes: a review. Diabetes 51(Suppl 3):S455–S461 [DOI] [PubMed] [Google Scholar]
  92. Harding HP, Zhang Y, Ron D 1999 Protein translation and folding are coupled by an endoplasmic-reticulum-resident kinase. Nature 397:271–274 [DOI] [PubMed] [Google Scholar]
  93. Liu CY, Schroder M, Kaufman RJ 2000 Ligand-independent dimerization activates the stress response kinases IRE1 and PERK in the lumen of the endoplasmic reticulum. J Biol Chem 275:24881–24885 [DOI] [PubMed] [Google Scholar]
  94. Harding HP, Zeng H, Zhang Y, Jungries R, Chung P, Plesken H, Sabatini DD, Ron D 2001 Diabetes mellitus and exocrine pancreatic dysfunction in perk−/− mice reveals a role for translational control in secretory cell survival. Mol Cell 7:1153–1163 [DOI] [PubMed] [Google Scholar]
  95. Novials A, Sarri Y, Casamitjana R, Rivera F, Gomis R 1993 Regulation of islet amyloid polypeptide in human pancreatic islets. Diabetes 42:1514–1519 [DOI] [PubMed] [Google Scholar]
  96. Fumarola C, Guidotti GG 2004 Stress-induced apoptosis: toward a symmetry with receptor-mediated cell death. Apoptosis 9:77–82 [DOI] [PubMed] [Google Scholar]
  97. Ferri KF, Kroemer G 2001 Organelle-specific initiation of cell death pathways. Nat Cell Biol 3:E255–E263 [DOI] [PubMed] [Google Scholar]
  98. Rhodes CJ, Brennan SO, Hutton JC 1989 Proalbumin to albumin conversion by a proinsulin processing endopeptidase of insulin secretory granules. J Biol Chem 264:14240–14245 [PubMed] [Google Scholar]
  99. Charge SB, de Koning EJ, Clark A 1995 Effect of pH and insulin on fibrillogenesis of islet amyloid polypeptide in vitro. Biochemistry 34:14588–14593 [DOI] [PubMed] [Google Scholar]
  100. Mulder H, Ahren B, Stridsberg M, Sundler F 1995 Non-parallelism of islet amyloid polypeptide (amylin) and insulin gene expression in rats islets following dexamethasone treatment. Diabetologia 38:395–402 [DOI] [PubMed] [Google Scholar]
  101. Bretherton-Watt D, Ghatei MA, Bloom SR, Jamal H, Ferrier GJ, Girgis SI, Legon S 1989 Altered islet amyloid polypeptide (amylin) gene expression in rat models of diabetes. Diabetologia 32:881–883 [DOI] [PubMed] [Google Scholar]
  102. Liu YQ, Nevin PW, Leahy JL 2000 β-Cell adaptation in 60% pancreatectomy rats that preserves normoinsulinemia and normoglycemia. Am J Physiol Endocrinol Metab 279:E68–E73 [DOI] [PubMed] [Google Scholar]
  103. Ma Z, Westermark GT, Sakagashira S, Sanke T, Gustavsson A, Sakamoto H, Engstrom U, Nanjo K, Westermark P 2001 Enhanced in vitro production of amyloid-like fibrils from mutant (S20G) islet amyloid polypeptide. Amyloid 8:242–249 [DOI] [PubMed] [Google Scholar]
  104. Qiu WQ, Folstein MF 2006 Insulin, insulin-degrading enzyme and amyloid-β peptide in Alzheimer’s disease: review and hypothesis. Neurobiol Aging 27:190–198 [DOI] [PubMed] [Google Scholar]
  105. Bennett RG, Hamel FG, Duckworth WC 2003 An insulin-degrading enzyme inhibitor decreases amylin degradation, increases amylin-induced cytotoxicity, and increases amyloid formation in insulinoma cell cultures. Diabetes 52:2315–2320 [DOI] [PubMed] [Google Scholar]
  106. Leissring MA, Farris W, Chang AY, Walsh DM, Wu X, Sun X, Frosch MP, Selkoe DJ 2003 Enhanced proteolysis of β-amyloid in APP transgenic mice prevents plaque formation, secondary pathology, and premature death. Neuron 40:1087–1093 [DOI] [PubMed] [Google Scholar]
  107. Janson J, Laedtke T, Parisi JE, O’Brien P, Petersen RC, Butler PC 2004 Increased risk of type 2 diabetes in Alzheimer disease. Diabetes 53:474–481 [DOI] [PubMed] [Google Scholar]
  108. Hori O, Ichinoda F, Tamatani T, Yamaguchi A, Sato N, Ozawa K, Kitao Y, Miyazaki M, Harding HP, Ron D, Tohyama M, D MS, Ogawa S 2002 Transmission of cell stress from endoplasmic reticulum to mitochondria: enhanced expression of Lon protease. J Cell Biol 157:1151–1160 [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Anello M, Lupi R, Spampinato D, Piro S, Masini M, Boggi U, Del Prato S, Rabuazzo AM, Purrello F, Marchetti P 2005 Functional and morphological alterations of mitochondria in pancreatic β cells from type 2 diabetic patients. Diabetologia 48:282–289 [DOI] [PubMed] [Google Scholar]
  110. Couce M, O’Brien TD, Moran A, Roche PC, Butler PC 1996 Diabetes mellitus in cystic fibrosis is characterized by islet amyloidosis. J Clin Endocrinol Metab 81:1267–1272 [DOI] [PubMed] [Google Scholar]
  111. Marzban L, Rhodes CJ, Steiner DF, Haataja L, Halban PA, Verchere CB 2006 Impaired NH2-terminal processing of human proislet amyloid polypeptide by the prohormone convertase PC2 leads to amyloid formation and cell death. Diabetes 55:2192–2201 [DOI] [PubMed] [Google Scholar]
  112. Bellamy CO, Malcomson RD, Harrison DJ, Wyllie AH 1995 Cell death in health and disease: the biology and regulation of apoptosis. Semin Cancer Biol 6:3–16 [DOI] [PubMed] [Google Scholar]
  113. Gavrieli Y, Sherman Y, Ben-Sasson SA 1992 Identification of programmed cell death in situ via specific labeling of nuclear DNA fragmentation. J Cell Biol 119:493–501 [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Riedl SJ, Shi Y 2004 Molecular mechanisms of caspase regulation during apoptosis. Nat Rev Mol Cell Biol 5:897–907 [DOI] [PubMed] [Google Scholar]
  115. Meier P, Finch A, Evan G 2000 Apoptosis in development. Nature 407:796–801 [DOI] [PubMed] [Google Scholar]
  116. Boyce M, Degterev A, Yuan J 2004 Caspases: an ancient cellular sword of Damocles. Cell Death Differ 11:29–37 [DOI] [PubMed] [Google Scholar]
  117. Katayama T, Imaizumi K, Manabe T, Hitomi J, Kudo T, Tohyama M 2004 Induction of neuronal death by ER stress in Alzheimer’s disease. J Chem Neuroanat 28:67–78 [DOI] [PubMed] [Google Scholar]
  118. Eckert A, Marques CA, Keil U, Schussel K, Muller WE 2003 Increased apoptotic cell death in sporadic and genetic Alzheimer’s disease. Ann NY Acad Sci 1010:604–609 [DOI] [PubMed] [Google Scholar]
  119. Wajant H 2002 The Fas signaling pathway: more than a paradigm. Science 296:1635–1636 [DOI] [PubMed] [Google Scholar]
  120. Masud A, Mohapatra A, Lakhani SA, Ferrandino A, Hakem R, Flavell RA 2007 Endoplasmic reticulum stress-induced death of mouse embryonic fibroblasts requires the intrinsic pathway of apoptosis. J Biol Chem 282:14132–14139 [DOI] [PubMed] [Google Scholar]
  121. Schroder M, Kaufman RJ 2005 ER stress and the unfolded protein response. Mutat Res 569:29–63 [DOI] [PubMed] [Google Scholar]
  122. Rutkowski DT, Kaufman RJ 2004 A trip to the ER: coping with stress. Trends Cell Biol 14:20–28 [DOI] [PubMed] [Google Scholar]
  123. Araki E, Oyadomari S, Mori M 2003 Impact of endoplasmic reticulum stress pathway on pancreatic β-cells and diabetes mellitus. Exp Biol Med (Maywood) 228:1213–1217 [DOI] [PubMed] [Google Scholar]
  124. Meier JJ, Bhushan A, Butler AE, Rizza RA, Butler PC 2005 Sustained β cell apoptosis in patients with long-standing type 1 diabetes: indirect evidence for islet regeneration? Diabetologia 48:2221–2228 [DOI] [PubMed] [Google Scholar]
  125. Augstein P, Elefanty AG, Allison J, Harrison LC 1998 Apoptosis and β-cell destruction in pancreatic islets of NOD mice with spontaneous and cyclophosphamide-accelerated diabetes. Diabetologia 41:1381–1388 [DOI] [PubMed] [Google Scholar]
  126. Maedler K, Sergeev P, Ris F, Oberholzer J, Joller-Jemelka HI, Spinas GA, Kaiser N, Halban PA, Donath MY 2002 Glucose-induced β cell production of IL-1β contributes to glucotoxicity in human pancreatic islets. J Clin Invest 110:851–860 [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Maedler K, Spinas GA, Lehmann R, Sergeev P, Weber M, Fontana A, Kaiser N, Donath MY 2001 Glucose induces β-cell apoptosis via upregulation of the Fas receptor in human islets. Diabetes 50:1683–1690 [DOI] [PubMed] [Google Scholar]
  128. Finkel E 2001 The mitochondrion: is it central to apoptosis? Science 292:624–626 [DOI] [PubMed] [Google Scholar]
  129. Oyadomari S, Takeda K, Takiguchi M, Gotoh T, Matsumoto M, Wada I, Akira S, Araki E, Mori M 2001 Nitric oxide-induced apoptosis in pancreatic β cells is mediated by the endoplasmic reticulum stress pathway. Proc Natl Acad Sci USA 98:10845–10850 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Yoshioka M, Kayo T, Ikeda T, Koizumi A 1997 A novel locus, Mody4, distal to D7Mit189 on chromosome 7 determines early-onset NIDDM in nonobese C57BL/6 (Akita) mutant mice. Diabetes 46:887–894 [DOI] [PubMed] [Google Scholar]
  131. Huang CJ, Lin CY, Haataja L, Gurlo T, Butler AE, Rizza RA, Butler PC 2007 High expression rates of human islet amyloid polypeptide induce endoplasmic reticulum stress mediated β-cell apoptosis, a characteristic of humans with type 2 but not type 1 diabetes. Diabetes 56:2016–2027 [DOI] [PubMed] [Google Scholar]
  132. Knorre A, Wagner M, Schaefer HE, Colledge WH, Pahl HL 2002 δF508-CFTR causes constitutive NF-κB activation through an ER-overload response in cystic fibrosis lungs. Biol Chem 383:271–282 [DOI] [PubMed] [Google Scholar]
  133. Kouroku Y, Fujita E, Jimbo A, Kikuchi T, Yamagata T, Momoi MY, Kominami E, Kuida K, Sakamaki K, Yonehara S, Momoi T 2002 Polyglutamine aggregates stimulate ER stress signals and caspase-12 activation. Hum Mol Genet 11:1505–1515 [DOI] [PubMed] [Google Scholar]
  134. Marchetti P, Bugliani M, Lupi R, Marselli L, Masini M, Boggi U, Filipponi F, Weir GC, Eizirik DL, Cnop M 2007 The endoplasmic reticulum in pancreatic β cells of type 2 diabetes patients. Diabetologia 50:2486–2494 [DOI] [PubMed] [Google Scholar]
  135. Rizzuto R, Brini M, Murgia M, Pozzan T 1993 Microdomains with high Ca2+ close to IP3-sensitive channels that are sensed by neighboring mitochondria. Science 262:744–747 [DOI] [PubMed] [Google Scholar]
  136. Takano J, Tomioka M, Tsubuki S, Higuchi M, Iwata N, Itohara S, Maki M, Saido TC 2005 Calpain mediates excitotoxic DNA fragmentation via mitochondrial pathways in adult brains: evidence from calpastatin mutant mice. J Biol Chem 280:16175–16184 [DOI] [PubMed] [Google Scholar]
  137. Higuchi M, Tomioka M, Takano J, Shirotani K, Iwata N, Masumoto H, Maki M, Itohara S, Saido TC 2005 Distinct mechanistic roles of calpain and caspase activation in neurodegeneration as revealed in mice overexpressing their specific inhibitors. J Biol Chem 280:15229–15237 [DOI] [PubMed] [Google Scholar]
  138. Nakagawa T, Zhu H, Morishima N, Li E, Xu J, Yankner BA, Yuan J 2000 Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-β. Nature 403:98–103 [DOI] [PubMed] [Google Scholar]
  139. Lamkanfi M, Kalai M, Vandenabeele P 2004 Caspase-12: an overview. Cell Death Differ 11:365–368 [DOI] [PubMed] [Google Scholar]
  140. Hitomi J, Katayama T, Eguchi Y, Kudo T, Taniguchi M, Koyama Y, Manabe T, Yamagishi S, Bando Y, Imaizumi K, Tsujimoto Y, Tohyama M 2004 Involvement of caspase-4 in endoplasmic reticulum stress-induced apoptosis and Aβ-induced cell death. J Cell Biol 165:347–356 [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Katayama T, Imaizumi K, Honda A, Yoneda T, Kudo T, Takeda M, Mori K, Rozmahel R, Fraser P, George-Hyslop PS, Tohyama M 2001 Disturbed activation of endoplasmic reticulum stress transducers by familial Alzheimer’s disease-linked presenilin-1 mutations. J Biol Chem 276:43446–43454 [DOI] [PubMed] [Google Scholar]
  142. Wang C, Tan JM, Ho MW, Zaiden N, Wong SH, Chew CL, Eng PW, Lim TM, Dawson TM, Lim KL 2005 Alterations in the solubility and intracellular localization of parkin by several familial Parkinson’s disease-linked point mutations. J Neurochem 93:422–431 [DOI] [PubMed] [Google Scholar]
  143. Yamaguchi H, Wang HG 2004 CHOP is involved in endoplasmic reticulum stress-induced apoptosis by enhancing DR5 expression in human carcinoma cells. J Biol Chem 279:45495–45502 [DOI] [PubMed] [Google Scholar]
  144. Zhang S, Liu H, Yu H, Cooper GJ 2008 Fas-associated death receptor signaling evoked by human amylin in islet β-cells. Diabetes 57:348–356 [DOI] [PubMed] [Google Scholar]
  145. Zhang S, Liu J, Dragunow M, Cooper GJ 2003 Fibrillogenic amylin evokes islet β-cell apoptosis through linked activation of a caspase cascade and JNK1. J Biol Chem 278:52810–52819 [DOI] [PubMed] [Google Scholar]
  146. Zhang S, Liu H, Liu J, Tse CA, Dragunow M, Cooper GJ 2006 Activation of activating transcription factor 2 by p38 MAP kinase during apoptosis induced by human amylin in cultured pancreatic β-cells. FEBS J 273:3779–3791 [DOI] [PubMed] [Google Scholar]
  147. Zhang S, Liu J, MacGibbon G, Dragunow M, Cooper GJ 2002 Increased expression and activation of c-Jun contributes to human amylin-induced apoptosis in pancreatic islet β-cells. J Mol Biol 324:271–285 [DOI] [PubMed] [Google Scholar]
  148. Clark A, Edwards CA, Ostle LR, Sutton R, Rothbard JB, Morris JF, Turner RC 1989 Localization of islet amyloid peptide in lipofuscin bodies and secretory granules of human B-cells and in islets of type-2 diabetic subjects. Cell Tissue Res 257:179–185 [DOI] [PubMed] [Google Scholar]
  149. Marzban L, Scrocchi LA, Warnock GL, Rosenberg L, Fraser PE, Verchere CB 2005 Hexapeptide inhibitors of islet amyloid polypeptide aggregation prevent islet amyloid formation and enhance survival of cultured human islets. Diabetes 54:A391 [Google Scholar]
  150. Paulsson JF, Andersson A, Westermark P, Westermark GT 2006 Intracellular amyloid-like deposits contain unprocessed pro-islet amyloid polypeptide (proIAPP) in β cells of transgenic mice overexpressing the gene for human IAPP and transplanted human islets. Diabetologia 49:1237–1246 [DOI] [PubMed] [Google Scholar]
  151. Linn T, Schmitz J, Hauck-Schmalenberger I, Lai Y, Bretzel RG, Brandhorst H, Brandhorst D 2006 Ischaemia is linked to inflammation and induction of angiogenesis in pancreatic islets. Clin Exp Immunol 144:179–187 [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Scheuner D, Kaufman RJ 2008 The unfolded protein response: a pathway that links insulin demand with β-cell failure and diabetes. Endocr Rev 29:317–333 [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Meier JJ, Kayed R, Lin CY, Gurlo T, Haataja L, Jayasinghe S, Langen R, Glabe CG, Butler PC 2006 Inhibition of human IAPP fibril formation does not prevent β-cell death: evidence for distinct actions of oligomers and fibrils of human IAPP. Am J Physiol Endocrinol Metab 291:E1317–E1324. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

[RPHR Note]
er.2007-0037_1.pdf (10KB, pdf)

Articles from Endocrine Reviews are provided here courtesy of The Endocrine Society

RESOURCES