18
Endoplasmic Reticulum Stress Signaling in Disease STEFAN J. MARCINIAK AND DAVID RON Cambridge Institute for Medical Research, University of Cambridge, Cambridge, United Kingdom; and Skirball Institute of Biomolecular Medicine, New York University School of Medicine, New York, New York I. Introduction 1133 A. Synthesis, folding, and misfolding of ER protein 1134 B. ER-associated degradation 1134 C. The UPR 1135 II. Endoplasmic Reticulum Homeostasis 1135 A. Short-term perturbations 1135 B. Long-term change 1140 III. Concluding Remarks 1143 Marciniak, Stefan J., and David Ron. Endoplasmic Reticulum Stress Signaling in Disease. Physiol Rev 86: 1133–1149, 2006; doi:10.1152/physrev.00015.2006.—The extracellular space is an environment hostile to unmodified polypeptides. For this reason, many eukaryotic proteins destined for exposure to this environment through secretion or display at the cell surface require maturation steps within a specialized organelle, the endoplasmic reticulum (ER). A complex homeostatic mechanism, known as the unfolded protein response (UPR), has evolved to link the load of newly synthesized proteins with the capacity of the ER to mature them. It has become apparent that dysfunction of the UPR plays an important role in some human diseases, especially those involving tissues dedicated to extracellular protein synthesis. Diabetes mellitus is an example of such a disease, since the demands for constantly varying levels of insulin synthesis make pancreatic -cells dependent on efficient UPR signaling. Furthermore, recent discoveries in this field indicate that the importance of the UPR in diabetes is not restricted to the -cell but is also involved in peripheral insulin resistance. This review addresses aspects of the UPR currently understood to be involved in human disease, including their role in diabetes mellitus, atherosclerosis, and neoplasia. I. INTRODUCTION The success of multicellular organisms owes much to the efficiency gained from the cooperation between spe- cialized cell types. This cooperation requires communica- tion between distant components frequently achieved by the binding of molecules from one cell to receptors on another. Such information transfer requires proteins to be synthesized that can withstand the harsh extracellular environment, both as soluble ligands (hormones and transmitters) and cell surface molecules (receptors and adhesion molecules). The extracellular compartment differs sufficiently from the cytosol that proteins destined for secretion or insertion into the plasma membrane require modifications inappropriate to the cytosol, such as glycosylation and disulfide bond formation. Generation of these modifica- tions necessitates a compartment topologically distinct from the cytosol, which is provided by a membranous network called the endoplasmic reticulum (ER). With evolution of this compartment, eukaryotes internalized a portion of the extracellular space in which they modify, fold, and assemble secreted and membrane proteins. The evolution of eukaryotes also created the challenge of regulating a protein maturation machinery outside the confines of the cytosol. Failure of this machinery to fold newly synthesized endoplasmic reticulum “client” pro- teins presents unique dangers to the cell and is termed “ER stress.” Early in evolution, a homeostatic mechanism developed to maintain the balance between the demand for ER function and ER synthetic capacity; furthermore, as organisms became more complex, especially with the appearance of long-lived professional secretory cells, the importance and complexity of this machinery increased greatly. This review addresses the mechanisms that enable higher metazoans to produce extracellular proteins in the face of varying demand and the consequences should these mechanisms fail. In the first section, a brief over- view of the ER protein maturation machinery is pre- Physiol Rev 86: 1133–1149, 2006; doi:10.1152/physrev.00015.2006. www.prv.org 1133 0031-9333/06 $18.00 Copyright © 2006 the American Physiological Society on February 11, 2015 Downloaded from

Endoplasmic Reticulum Stress Signaling in Disease

Embed Size (px)

DESCRIPTION

science

Citation preview

Page 1: Endoplasmic Reticulum Stress Signaling in Disease

Endoplasmic Reticulum Stress Signaling in Disease

STEFAN J. MARCINIAK AND DAVID RON

Cambridge Institute for Medical Research, University of Cambridge, Cambridge, United Kingdom; and

Skirball Institute of Biomolecular Medicine, New York University School of Medicine, New York, New York

I. Introduction 1133A. Synthesis, folding, and misfolding of ER protein 1134B. ER-associated degradation 1134C. The UPR 1135

II. Endoplasmic Reticulum Homeostasis 1135A. Short-term perturbations 1135B. Long-term change 1140

III. Concluding Remarks 1143

Marciniak, Stefan J., and David Ron. Endoplasmic Reticulum Stress Signaling in Disease. Physiol Rev 86:1133–1149, 2006; doi:10.1152/physrev.00015.2006.—The extracellular space is an environment hostile to unmodifiedpolypeptides. For this reason, many eukaryotic proteins destined for exposure to this environment through secretionor display at the cell surface require maturation steps within a specialized organelle, the endoplasmic reticulum(ER). A complex homeostatic mechanism, known as the unfolded protein response (UPR), has evolved to link theload of newly synthesized proteins with the capacity of the ER to mature them. It has become apparent thatdysfunction of the UPR plays an important role in some human diseases, especially those involving tissues dedicatedto extracellular protein synthesis. Diabetes mellitus is an example of such a disease, since the demands forconstantly varying levels of insulin synthesis make pancreatic �-cells dependent on efficient UPR signaling.Furthermore, recent discoveries in this field indicate that the importance of the UPR in diabetes is not restricted tothe �-cell but is also involved in peripheral insulin resistance. This review addresses aspects of the UPR currentlyunderstood to be involved in human disease, including their role in diabetes mellitus, atherosclerosis, and neoplasia.

I. INTRODUCTION

The success of multicellular organisms owes much tothe efficiency gained from the cooperation between spe-cialized cell types. This cooperation requires communica-tion between distant components frequently achieved bythe binding of molecules from one cell to receptors onanother. Such information transfer requires proteins to besynthesized that can withstand the harsh extracellularenvironment, both as soluble ligands (hormones andtransmitters) and cell surface molecules (receptors andadhesion molecules).

The extracellular compartment differs sufficientlyfrom the cytosol that proteins destined for secretion orinsertion into the plasma membrane require modificationsinappropriate to the cytosol, such as glycosylation anddisulfide bond formation. Generation of these modifica-tions necessitates a compartment topologically distinctfrom the cytosol, which is provided by a membranousnetwork called the endoplasmic reticulum (ER). With

evolution of this compartment, eukaryotes internalized aportion of the extracellular space in which they modify,fold, and assemble secreted and membrane proteins. Theevolution of eukaryotes also created the challenge ofregulating a protein maturation machinery outside theconfines of the cytosol. Failure of this machinery to foldnewly synthesized endoplasmic reticulum “client” pro-teins presents unique dangers to the cell and is termed“ER stress.” Early in evolution, a homeostatic mechanismdeveloped to maintain the balance between the demandfor ER function and ER synthetic capacity; furthermore,as organisms became more complex, especially with theappearance of long-lived professional secretory cells, theimportance and complexity of this machinery increasedgreatly.

This review addresses the mechanisms that enablehigher metazoans to produce extracellular proteins in theface of varying demand and the consequences shouldthese mechanisms fail. In the first section, a brief over-view of the ER protein maturation machinery is pre-

Physiol Rev 86: 1133–1149, 2006;doi:10.1152/physrev.00015.2006.

www.prv.org 11330031-9333/06 $18.00 Copyright © 2006 the American Physiological Society

on February 11, 2015

Dow

nloaded from

Page 2: Endoplasmic Reticulum Stress Signaling in Disease

sented, along with some of the consequences of proteinmisfolding. Thereafter, the response to short-term pertur-bations in ER function is discussed, with emphasis givento human diseases caused by dysregulation of this re-sponse, followed finally by an examination of the impor-tance of ER stress signaling over the longer term tosecretory cell differentiation and lipid metabolism.

A. Synthesis, Folding, and Misfolding

of ER Protein

Translation of membrane and extracellular proteinsis performed by ribosomes on the cytosolic surface of theER (3, 94). A signal recognition particle within the cytosolcotranslationally recognizes a signal sequence within thenascent polypeptide chain and directs it to a protein-aceous pore in the ER membrane, the Sec61 complex (92,127, 149).

Proteins achieve a specific folded conformation firstby acquisition of secondary structure, e.g., helices, due tothe intrinsic properties of the amino acid sequence (39).This is followed by a further search for native structure,which involves diffusion of these preformed segmentsand assembly into larger modules (84). These processesentail the burial of amino acid side chains into a close-packed structure excluding water from the protein’s core(25). “Misfolding,” in this context, indicates that a proteinpersistently maintains the potential for nonnative interac-tions that interfere with its structure and function. Thesemay involve normally buried residues being exposed tothe solvent promoting illegitimate interactions with othercellular components. Another consequence of aberrantfolding is aggregation of proteins into insoluble higherorder structures. These may be of three varieties, all ofwhich may cause disease: disordered aggregates (e.g.,rhodopsin in autosomal retinitis pigmentosa, Ref. 163),amyloid fibrils (e.g., amyloid �-peptide/tau in Alzheimer’sdisease, Refs. 89, 148), and nonamyloid fibrils (e.g., �1-antitryspin in �1-antitrypsin deficiency, Ref. 110).

Disordered aggregates are complex insoluble accu-mulations of protein resistant to normal degradation,while both amyloid and nonamyloid fibrils are simplenoncovalent polymers, one-dimensional crystals formedfrom repeating subunits. Nonamyloid fibrils are charac-teristic of the serpinopathies, human diseases caused bymutations within the serpin family of proteins (111); thesemutations allow a specific intermolecular interaction re-sulting in polymerization. In contrast, amyloid can beformed through polymerization of a number of otherwiseunrelated proteins, each causing a distinct disease (79,148, 207). Despite intensive study of these aggregates,surprisingly little is certain about the mechanism of theirtoxicity; however, it is becoming apparent that even in thecase of amyloidoses, in which extracellular deposits are

characteristic, some of the deleterious effect comes fromtoxic gain of function due to protein oligomers within thecell (227).

The inappropriate interactions of misfolded proteinsthreaten to disrupt normal cellular function, so the bio-synthetic machinery has evolved sophisticated mecha-nisms to minimize their occurrence. Within the ER, chap-erones such as BiP bind to incompletely folded proteinsshielding them from other molecules (13, 64, 65, 126, 200).In addition, other catalysts directly promote correct fold-ing by the addition of carbohydrates (63), cis-trans

isomerization about peptide bonds (166), and the creationand rearrangement of disulfide bonds (189).

B. ER-Associated Degradation

Some proteins never attain their correct conforma-tion, perhaps due to a mutation impeding correct foldingor because the cell lacks the energy to drive sufficientcycles of chaperone interaction. After a lag period of30–90 min, misfolded ER client proteins are disposed ofby ER-associated degradation (ERAD) (107). The timerthat dictates this delay, at least for misfolded glycopro-teins, is likely to be ER mannosidase I, which operates bytrimming mannose residues from N-linked glycans (38,76). Evidence that this regulates degradation comes fromits inhibition, which retards ERAD (115, 186), and from itsoverexpression, which enhances ERAD (72). Recently,EDEM, a stress-inducible catalytically inactive mannosi-dase homolog, has been shown to interact with misfoldedglycoproteins and effect their extraction from the foldingcycle (43, 73, 122).

The path by which unfolded proteins are retrotrans-located into the cytosol is not clear. Some evidence sup-ports a role for the original Sec61 translocon complex(150, 228), although a new complex containing derlin-1and p97, a cytosolic ATPase, has been shown to be in-volved in retrotranslocation of major histocompatibilitycomplex (MHC) class I molecules (106, 219). Once withinthe cytosol, ERAD substrates are degraded by the ubiq-uitin/proteosome pathway (49, 51, 151).

For these synthetic and quality assurance processesto proceed efficiently, there needs to be coordinationbetween the input load of unfolded client proteins and thematuration machinery of the ER. If the client protein loadis excessive compared with the reserve of ER chaperones,the cell is said to be experiencing “ER stress.” If un-checked, ER stress threatens to overwhelm the process-ing capacity of the ER, leading to the accumulation ofunfolded proteins and collapse of the secretory pathway.Consequently, signaling pathways have evolved that re-spond to ER stress by regulating processes on both sidesof the ER membrane through an adaptive mechanismtermed the unfolded protein response (UPR).

1134 STEFAN J. MARCINIAK AND DAVID RON

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 3: Endoplasmic Reticulum Stress Signaling in Disease

C. The UPR

The nature of ER stress encountered by a cell dic-tates the nature of its UPR. During normal function asecretory cell will experience dramatic variations in theflux of new proteins through its ER in response tochanges in demand. During a transition from low proteinsynthesis to high, there is a need to increase ER protein-folding capacity to avoid overloading of chaperones. Con-sequently, the UPR regulates transcription factors whosetargets include genes for components of the ER proteinmaturation machinery. However, inherent in this re-sponse is a temporal delay, since new chaperones cannotbe made instantly. To avoid the accumulation of mis-folded client protein during this hiatus, the rate of secre-tory protein synthesis must also be under UPR-directedmodulation through a more rapid transcription-indepen-dent mechanism. Furthermore, when protein-folding effi-ciency falls, for instance, in ischemic tissue when lack ofenergy impedes the proper function of ER chaperonesand enzymes, the primary adaptation may be throughreduced client synthesis rather than by increasing thelevels of this machinery, which itself would be costly inenergy. Therefore, both transcriptional and translationalsignals emanate from the stressed ER to allow a coordi-nated response.

Not all changes in ER protein flux are transitory norcan all perturbations be survived. In some cases thechange in ER protein synthesis accompanies a change incellular phenotype, such as during differentiation into asecretory cell. Here the prolonged activation of UPR sig-naling itself appears to play a role in the differentiationprocess, leading to dramatic alterations in ER structure.When the level of ER stress is too great to allow adapta-tion, the cell may die. It remains unclear to what extentthe UPR plays a role directly in cell death, but studies ofER stress lethality have revealed novel potential thera-peutic targets for intervention in human disease.

These concepts will be elaborated upon in more de-tail below with emphasis on the human diseases to whichthey are most relevant.

II. ENDOPLASMIC RETICULUM HOMEOSTASIS

A. Short-Term Perturbations

Tissues whose primary function is the secretion ofprotein should depend most strongly on the UPR, espe-cially those liable to make large changes in ER clientprotein load. A good example is the pancreatic �-cell,which produces insulin in response to changes in circu-lating glucose. It constantly executes changes in ER syn-thetic capacity to track current glucose levels. Althoughthe precise mechanisms remain unclear, it has been es-

tablished that upswings in glucose, at least in the shortterm, promote proinsulin synthesis through enhancedprotein translation (75, 203). This stimulation of proinsu-lin translation by glucose is specific to the �-cell anddependent on cis-acting elements in untranslated regionsof proinsulin mRNA (206). The resultant large fluctuationsin ER client protein load are compensated for by the UPR,but render the �-cell highly vulnerable to defects in ERstress signaling in animal models (54) and in human dis-ease (34).

1. Regulating client protein load

Clues to the homeostatic importance of regulatingnew protein synthesis in response to changes in ER loadhave come from a rare human disease. Wolcott-Rallisonsyndrome is an autosomal recessive condition character-ized by early development of diabetes mellitus with asso-ciated bone, liver, renal, and neuronal defects (209). It hasbeen mapped to 2p12 where candidate gene analysisyielded mutations in EIF2AK3 (34), better known as PKR-like eukaryotic initiation factor 2 (eIF2�) kinase (PERK)(56) or pancreatic eIF2� kinase (PEK) (171).

PERK is an ER-resident transmembrane protein ubiq-uitously expressed, but highly enriched in professionalsecretory cells (56, 170, 171, 175). Its cytosolic portion ishighly homologous to a yeast stress-responsive kinase,Gcn2p (56, 170). Unlike PERK, Gcn2p is a soluble proteinthat enables yeast to adapt their rate of new proteinsynthesis to the levels of available amino acids (125, 190,191, 208). When amino acids are limiting, Gcn2p phos-phorylates the �-subunit of eIF2�.

The eIF2 complex is essential in all eukaryotes fornew protein synthesis, since it recruits the initiator me-thionyl tRNA to ribosomes about to begin translation (69).Phosphorylation of eIF2� inhibits this activity and thusglobally reduces protein translation. In metazoans, theGCN2 gene has expanded into a family of related eIF2�kinases, all of which inhibit protein translation in re-sponse to stress. These kinases all possess homologouscatalytic domains but have different stress-sensitive reg-ulatory domains. GCN2 persists to respond to amino acidstarvation, while new members sense disparate stressesincluding viral infection (PKR and PKZ) and iron defi-ciency (HRI) (9, 10, 23, 44, 159). The existence of PERK,an ER transmembrane eIF2� kinase, led to the apprecia-tion that cells are able to adjust their level of new proteinsynthesis to match their ER folding capacity and preventoverload of the secretory pathway (55, 56).

To respond to ER stress, PERK must transduce aluminal signal across the ER membrane to its cytosolickinase domain. The nature of this stimulus is intriguing.The ER is a site of secretory protein maturation and sowill, by necessity, harbor incompletely folded and activelyfolding protein intermediates. The level of these is a reg-

ENDOPLASMIC RETICULUM AND DISEASE 1135

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 4: Endoplasmic Reticulum Stress Signaling in Disease

ister of client protein throughput, rather than an indicatorof the ER’s capacity to fold them. A homeostatic mecha-nism that responded solely to unfolded protein could not,therefore, measure the efficiency of ER protein biosyn-thesis, but, instead, only respond to the absolute clientload. Evolution appears to have found a more subtlesolution: PERK responds not to the level of misfoldedprotein within the ER, but instead to changes in the ERchaperone reserve, that is, to the unfolded/misfolded pro-tein-to-chaperone ratio. BiP is an abundant chaperonewithin the ER that binds to folding proteins through in-teraction with exposed hydrophobic residues. BiP over-expression has long been known to suppress the UPR (41,197). Two potentially complementary models have beenproposed for the mechanism(s) involved. During un-stressed conditions, BiP also binds to the luminal domainof PERK (12, 108). This interaction correlates with theinactive, monomeric state of PERK. When an increase inER client load is experienced, BiP dissociates from PERK,perhaps through sequestration to unfolded clients or bymore direct mechanisms. The dissociation of the PERK-BiP complex is hypothesized to allow PERK to cluster inthe plane of the membrane, leading to activation of thecytosolic kinase domain through a process of trans-auto-phosphorylation and a dramatic increase in affinity to-wards eIF2� (12, 116).

Recently, crystal structure data on IRE1, whose lu-minal domain is homologous to that of PERK, raise thepossibility that direct binding of sensing molecules tounfolded proteins might also take place (32). Dimeriza-tion of the luminal portion of IRE1 appears to generate agroove similar to that found in the peptide-binding pocketof MHC molecules. If the groove formed by dimerizationof IRE1 can be shown to bind unfolded proteins, thiswould suggest an alternative model whereby unfoldeddomains that are unbuffered by chaperones signal di-rectly in the unfolded protein response, perhaps combin-ing direct elements (unfolded protein) and indirect ele-ments (chaperone reserve).

Humans with this Wolcott-Rallison syndrome sharesome clinical features with PERK�/� mice, which areborn with essentially normal islets that are capable ofnormal insulin synthesis and secretion (54). Indeed, whenisolated islets from prediabetic PERK�/� mice are chal-lenged with glucose, their secretion of insulin is greaterthan wild-type controls. During the first few weeks of lifethese pups develop overt diabetes due to progressive�-cell destruction. These findings indicate that PERK isessential for �-cells to cope with normal physiological ERstress arising from day-to-day insulin synthesis. This in-volves restraining the insulin synthetic response to glu-cose by attenuating new protein translation. In otherwords, protein synthesis is limited in normal animals byeIF2� phosphorylation to a level with which the existingER machinery can cope; in the PERK�/� animals and

Wolcott-Rallison patients, new proteins continue to enterthe ER regardless of its ability to fold them. This mani-fests as the accumulation of misfolded products (109) andexcessive stress signaling and culminates in cell death(54). Interestingly, individual mutations of the otherknown eIF2� kinases (PKR, HRI, and GCN2) do not resultin mice with diabetes, indicating that only eIF2� phos-phorylation in response to ER stress is involved in thisphenotype (52, 217, 226).

The importance of translational regulation in re-sponse to ER stress may not be restricted to rare geneticdiseases, as a recent study, using transgenic mice subtlyimpaired in their ability to phosphorylate eIF2� has re-vealed a phenotype remarkably similar to that of humantype II diabetes. PERK and all eIF2� kinases phosphory-late a single residue on their substrate: serine-51. Muta-tion of this serine to an alanine (eIF2�S51A) prevents itsphosphorylation in response to any stress. HeterozygouseIF2�S51A mice are capable of significant translationalregulation by phosphorylation of the remaining popula-tion of eIF2� molecules and are born with normal pan-creatic islets and under normal conditions do not developdiabetes (164). In contrast, eIF2�S51A homozygotes areborn with severe �-cell deficiency by late embryonic stage(165). However, when heterozygous mice are fed a high-fat diet, a new and interesting phenotype is revealed(164). These mice are more prone to obesity because of afailure to increase energy expenditure in response to theexcess calories. Remarkably, they more rapidly exhibithyperleptinemia, mild hyperinsulinemia, and raised fast-ing glucose, features of the “metabolic syndrome” cur-rently dominating the epidemic of type 2 diabetes in de-veloped countries. When islets from high-fat diet-fedeIF2�S51A mice were analyzed in vitro, they demonstratedincreased basal insulin secretion but reduced stimulatedinsulin secretion. Prolonged high-fat feeding led to disten-sion of the ER by unfolded proteins, including proinsulin,and to profound glucose intolerance. This surprising find-ing highlights the exquisite sensitivity of pancreatic�-cells even to subtle dysregulation of eIF2a phosphory-lation and raises the intriguing possibility that minor vari-ations in efficiency of eIF2� phosphorylation could, inprinciple, contribute to significant morbidity. As yet, nostudies have addressed the possible existence of suchvariations in eIF2� signaling between patient groups, forexample, obese subjects with or without associated fea-tures of metabolic syndrome.

2. Integrated stress response

The effects of eIF2� phosphorylation are not re-stricted to attenuation of protein translation, but alsoinclude activation of a transcriptional program. This isillustrated in a family of white matter hypomyelinationdisorders, termed CACH/VWM leukodystrophies. These

1136 STEFAN J. MARCINIAK AND DAVID RON

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 5: Endoplasmic Reticulum Stress Signaling in Disease

are autosomal recessive conditions with a spectrum ofseverity from congenital to adult-onset disease, involvingthe progressive loss of mental and motor faculties due todeterioration of brain white matter. In all cases the caus-ative mutations have been found within subunits of aneIF2�-interacting protein, eIF2B (46, 47, 100, 158). As partof its normal active cycle, eIF2 undergoes rounds of GTPbinding, hydrolysis, and guanine nucleotide exchange.Only in the GTP-bound state is eIF2 capable of bringingmethionyl tRNA to the ribosome. The guanine nucleotideexchange factor (GEF) for eIF2 is, in fact, eIF2B; how-ever, once phosphorylated, eIF2 binds avidly to its GEFinhibiting further exchange (27). In this way, phosphory-lation of a small pool of eIF2 can inhibit most GEF activity(Fig. 1). The mutations that cause the CACH/VWM disor-ders appear to interfere with eIF2B GEF activity, butrecent studies with cells from patients with CACH/VWM

have shown that, rather than affecting global translationrates during stress, these mutations increase signaling viaa transcription factor known as ATF4 (82, 193).

Whilst the majority of RNA transcripts experiencedecreased translation during periods of increased eIF2�phosphorylation, a small and still poorly defined subset istranslated more efficiently (53, 68, 121, 124). This para-doxical effect is due to the presence of multiple upstreamopen reading frames (uORFs) 5� to the coding sequenceinitiation codon (1, 68, 112, 194). The best-characterizedexample in mammals is the stress-inducible transcriptionfactor ATF4 (112, 194). This transcript has two uORFs,the second overlapping out of frame with the true ATF

coding sequence. During resting unstressed conditions,ribosomes scan along the mRNA translating uORF1 andthen recapacitate by rebinding eIF2/GTP/Met-tRNA ter-nary complex in time to translate inhibitory uORF2. Dur-ing stress, limiting levels of eIF2 ternary complex lead toa delay in recapacitation of these scanning ribosomes,such that they fail to reinitiate at uORF2 but instead scanto the ATF4 initiation codon. By then, a proportion havereacquired ternary complex allowing translation of theactive transcription factor.

In yeast, the single eIF2� kinase Gcn2p regulates atranscription program governing the response to aminoacid deprivation through induction of the transcriptionfactor GCN4, by a mechanism analogous to that of ATF4

(37, 68, 125). As the family of eIF2� kinases has diversi-fied to respond to numerous stresses, so too has thetranscriptional program it regulates in metazoans. It con-tinues to regulate genes essential for amino acid suffi-ciency, but now ATF4 also induces antioxidant genes andgenes of the ER protein maturation machinery (57). Be-cause eIF2� phosphorylation triggers a final commonpathway in response to many stresses, this transcriptionalprogram in combination with protein translation modula-tion has been termed the integrated stress response (ISR)(57) (Fig. 1).

Oxidative protein folding is inextricably linked to thegeneration of reactive oxygen species (ROS), in large partthrough the activity of ER oxidase 1 (ERO1), which gen-erates much of the oxidizing potential of the organelle(48, 152, 187, 188). Without ER stress-regulated activationof ATF4, PERK�/� �-cells, and those from Wolcott-Ral-lison patients, lack the prosurvival effects of the ISR withits antioxidant expression program (54, 57). It is worthremarking, therefore, that recent data indicate the oxida-tion of cysteines to form disulfide bonds leads directly tothe generation of ROS and cell death (58) and that ERstress-dependent ERO1� induction promotes ER oxida-tion (117). Insulin, with its three disulfide bonds per mol-ecule, might therefore be expected to impose a consider-able ROS load on the cell. The finding that targeted dele-tion of a widely expressed ER-associated reductase,

FIG. 1. The integrated stress response (ISR). A: basally PERK ismaintained in an inactive state, perhaps through interactions with BiP,while GTP-bound eIF2 complex efficiently supports new protein trans-lation initiation. The guanine nucleotide exchange factor (GEF) eIF2Bmaintains the GTP-bound population of eIF2 complexes. B: during en-doplasmic reticulum (ER) stress BiP dissociates from PERK, whichbecomes active. Unfolded protein might interact directly with the lumi-nal domain of PERK to promote activation. Phosphorylation of eIF2 onits �-subunit leads directly to inhibition of eIF2B activity. When GTP-bound eIF2 levels are limiting, most protein translation is inhibited,while translation of ATF4 is enhanced. ATF4 migrates to the nucleus totransactive genes of the ISR.

ENDOPLASMIC RETICULUM AND DISEASE 1137

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 6: Endoplasmic Reticulum Stress Signaling in Disease

Ncb5or, leads to a selective loss of �-cells, once moreunderlines their sensitivity to redox stress (212).

3. Recovery of protein translation

A) EIF2� PHOSPHATASES. The execution of an adaptivegene expression program like the ISR can only be per-formed if the attenuation of protein translation accompa-nying eIF2� phosphorylation is reversed (29, 114, 131).Indeed, persisting loss of protein synthetic activity wouldcompromise cell survival, regardless of a need for adap-tation to stress. For these reasons, phosphatase activityexists to relieve eIF2� phosphorylation, restoring normaltranslation initiation activity.

Viral infection leads to eIF2� phosphorylationthrough activation of PKR (10, 90). This is part of aninnate antiviral response antagonizing viral protein trans-lation. Frequently, the relationships between infectiveagents and host are characterized by the adaptations ofone party being countered by adaptations within theother. For example, viruses have evolved proteins thatdirectly inhibit the activity of PKR (21, 153); however, inthe case of herpes simplex virus (HSV), viral proteintranslation can be maintained despite continued PKR ac-tivity. This is achieved through enhanced dephosphoryla-tion of eIF2� by a virally encoded protein, ICP34.5, whichbinds to a host serine/threonine phosphatase, proteinphosphatase 1 (PP1), and directs its activity towardseIF2� (61, 62).

It is now well appreciated that protein phosphatases,including PP1, are represented within the genome byrelatively few genes and achieve their specificity throughthe formation of complexes with regulatory subunits (28).It was therefore gratifying when suppressor screens of ERsignaling revealed two mammalian genes, CReP andGADD34, with striking homology to ICP34.5 (80, 131).These proved to be responsible for reversing PERK-in-duced translational attenuation, as they too are regulatorysubunits of PP1 that specifically promote eIF2� dephos-phorylation (29, 80, 131). CReP is constitutively ex-pressed, while GADD34 is strongly induced during ERstress. The induction of GADD34 in particular appearscentral to the reversal of stress-induced translational at-tenuation (18, 87, 114, 131, 132).

B) ROLE OF CHOP. Activation of ER stress signaling hasbeen correlated with the induction of cell death in manymodels of ER stress. The mechanisms by which this mightpromote cell death remain unclear; however, CHOP, atranscription factor downstream of the PERK-ATF4 axis,was thought to be important to the process, since itsdeletion ameliorates tissue damage during ER stress (20,136, 137, 174, 229). The downstream mediators of thiseffect remained unknown until recent work demonstratedthat some of the toxic effects of CHOP are the result ofGADD34 induction (117).

The CHOP dependence of ER stress toxicity is wellillustrated in a murine model of diabetes, the Akitamouse. Mice, unlike humans, have two genes encodinginsulin (Ins1 and Ins2). These appear functionally redun-dant, as deleting both alleles of either gene has no effecton glucose homeostasis (101). It is therefore remarkablethat a spontaneous mutation in the Ins2 gene (Akita)leads to a severe diabetic phenotype inherited as a semi-dominant trait (196, 223).

Insulin is translated as a single proinsulin polypep-tide that undergoes oxidation within the ER to form threeintramolecular disulfide bonds. After excision of a shortpeptide from the prohormone, these three bonds hold theremaining two insulin chains together. The Akita muta-tion converts a conserved cysteine to tyrosine preventingformation of one of these bonds (196). This mutant insulinis not secreted, but instead degraded (8, 196). Mice har-boring the Akita mutation have apparently normal pan-creatic islets at birth, but go on to develop diabetes in theensuing weeks due to selective �-cell loss in the absenceof inflammation. This was initially attributed to nonspe-cific effects (196); however, when this mutation was in-troduced into mice with a targeted deletion of the CHOP

gene, onset of diabetes was significantly delayed due tothe preservation of �-cell numbers (136).

As indicated above, CHOP is a transcription factorhighly induced during ER stress (197, 229). The protectiveeffect of CHOP deletion appears not restricted to theAkita mouse, since it also affords protection against cy-tokine-induced �-cell death, in which nitric oxide triggersER stress by the depletion of ER calcium stores (20, 137),nor is it restricted to the pancreas, as CHOP�/� animalsare resistant to renal damage caused by the ER stress-inducing toxin tunicamycin (229). CHOP deletion alsoprotects mice from dopaminergic neuron loss following6-hydroxydopamine injection (174). This drug, which iscommonly used as a model for Parkinson’s disease sinceit selectively kills dopaminergic neurons in vivo, causesER stress in dopaminergic neurons in tissue culture (71,162).

One popular hypothesis held that CHOP, a metazoanspecific gene, evolved to promote the death of individualcells in response to insurmountable levels of ER stress,affording particular benefit to multicellular organisms(135). In accordance with this, CHOP overexpression hasbeen shown to sensitize cells to the toxicity of ER stress(118). Analysis, however, of the CHOP transcriptionalprogram failed to reveal an obvious connection to path-ways that promote cell death. Instead, CHOP appears todefend ongoing protein secretion, in part through induc-tion of GADD34 (117). In this light, it appears that theER-stressed cell is challenged with balancing the need todefend its chaperone reserve, by limiting secretory pro-tein synthesis, against the need of both cell and organismfor ongoing ER synthetic function. According to this

1138 STEFAN J. MARCINIAK AND DAVID RON

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 7: Endoplasmic Reticulum Stress Signaling in Disease

model, the protection from death afforded by deletion ofthe CHOP gene in some models of ER stress reflects,therefore, an enforced shift of this balance from syntheticfunction towards the defense of chaperone reserve (Fig.2). In accordance with this, GADD34 null animals havebeen shown to be protected from ER stress-induced neph-rotoxicity equally well as CHOP�/� animals (117).

Of course, there exists the formal possibility thatCHOP-dependent GADD34 induction evolved to promotecell death rather than ongoing protein synthesis; however,two pieces of evidence argue against this. First of all, intissue culture GADD34 mutant cells are more prone tocell death in response to ER stress induced by thapsigar-gin (132). This agent causes ER stress through depletionof ER calcium stores and is particularly effective at shut-ting down protein synthesis, indicating that recovery ofprotein translation is necessary for the adaptive prosur-vival effects of UPR signaling. Second, in murine modelsof Pelizaeus-Merzbacher leukodystrophy (due to a prote-olipid protein mutation), CHOP expression appears tohave an antiapoptotic effect, a strange phenotype for aprodeath signal (176).

It remains unclear whether the toxic effects of CHOPin models of ER stress are primarily mediated throughGADD34. Although this may be true for tunicamycin-induced nephrotoxicity, it has yet to be proven for othermodels. For example, it is possible that other CHOP targetgenes, such as ERO1�, may mediate toxicity in a tissue-

specific fashion. In cell culture, CHOP overexpressionleads to ROS production (118), and CHOP-dependentERO1� induction appears responsible for increased oxi-dation in the stressed ER (117). This may be relevant totissues known to be sensitive to redox perturbation, suchas the pancreatic �-cell.

Perhaps there are circumstances experienced bymulticellular organisms during which the cost of worsen-ing ER stress is compensated for by enhanced proteinsecretion, at least in a subset of cells. Because osteoblastsare replenishable secretory cells they may behave in thisfashion and, consistent with this, recent data suggest thatthe bones of CHOP�/� animals demonstrate defects dueto impaired osteoblast function (147). If the importance ofCHOP in the defense of protein secretion is a generalfeature of replenishable secretory cells, it is tempting tospeculate that terminally differentiated effector cells ofthe immune system might also behave thus. ChallengingCHOP�/� and GADD34 mutant animals with infectiousagents could test this hypothesis.

An exciting prediction arising from these findings isthat careful titration of GADD34 inhibitors might be oftherapeutic benefit in diseases involving ER stress,through the preservation of ER chaperone reserve. It isencouraging, therefore, that a recent screen for smallmolecules that protect cells from ER stress yielded acompound that enhances eIF2� phosphorylation and, inHSV-infected cells, appears to block eIF2� dephosphory-lation by ICP34.5 (15). If compounds can be synthesizedthat have specificity for GADD34 over CReP, there existsthe potential for selective effects on ER-stressed tissue,perhaps limiting systemic toxicity.

4. NF-�B and ER activity

A distinct antiviral response initiated by the accumu-lation of viral proteins in the ER has previously beenpostulated (140, 141, 143). This followed from observa-tions that virally encoded proteins could activate NF-�Bsignaling in a cell autonomous fashion (119, 144, 145).This is in contrast to most canonical NF-�B stimuli, e.g.,tumor necrosis factor (TNF)-�, which originate from out-side the cell. This “ER overload response” (EOR), as itwas termed, appeared not to be specific for viral proteins,since other membrane proteins, e.g., MHC class I, whenoverexpressed could also induce NF-�B signaling (145). Itwas subsequently extended to include NF-�B inductionfrom ER accumulation of other nontransmembrane pro-teins, including �1-antitrypsin polymers (67, 95). Thetransduction mechanism for this pathway remained ob-scure, although release of ER calcium and ROS were bothsuggested (140, 143).

Recent discoveries have implicated PERK signalingin the activation of NF-�B in response to ER perturba-tions, challenging the notion that the EOR is distinct from

FIG. 2. Failure of ER homeostasis. The ratio of client protein load(gray spheres) to the availability of chaperones (red caps) varies de-pending on new protein translation rates. When levels of new proteinexceed the capacity of ER chaperones to bind them, aggregation ensues.PERK promotes eIF2� phosphorylation and thus inhibits protein trans-lation, which is represented by a rightward shift in this figure. GADD34induction, through eIF2� dephosphorylation and disinhibition of eIF2BGEF activity, promotes protein translation and a leftward shift with anincreased risk of client protein aggregation. CHOP�/� and GADD34

mutant cells experience lower levels of new protein aggregation becausereduced new protein synthesis defends the reserve of ER chaperones.

ENDOPLASMIC RETICULUM AND DISEASE 1139

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 8: Endoplasmic Reticulum Stress Signaling in Disease

the UPR (36, 78). There exists a protein family of NF-�Binhibitors termed I-�B, which, in unstimulated cells, isexpressed constitutively and sequesters NF-�B inactivewithin the cytoplasm (160). Classically, NF-�B is acti-vated through the relief of this inhibition, for example,TNF receptor activation triggers I-�B phosphorylation,causing its ubiquitination and degradation by the proteo-some (146). In the case of ER stress, at least when sig-naled by PERK, disinhibition of NF-�B is achieved onceagain by reducing I�B levels, but through a novel mech-anism. I-�B has a far shorter half-life than NF-�B. Conse-quently, translational attenuation preferentially lowersI�B levels, releasing NF-�B to execute its transcriptionprogram. A similar effect can be obtained by treatment ofcells in culture with other agents that inhibit proteinsynthesis, e.g., cycloheximide (36, 78). This novel mech-anism extends to NF-�B activation in response to otherstresses, including ultraviolet (UV) irradiation and aminoacid deprivation, where GCN2 appears to be the relevantkinase (77, 78, 211). Some controversy persists within thisfield, since the existence of a PERK-dependent NF-�Bcascade does not preclude activation of NF-�B by otherER-originating signals. This will be clarified once EORstressors have been more extensively studied in PERKmutant and eIF2�S51A cell lines.

The physiological significance of ER stress-inducedNF-�B signaling likely reflects the importance of NF-�Bsignaling in other settings, namely, the coordination of in-flammatory response and promotion of cellular survival. Inbacterial infection, the induction of NF-�B by microbialcomponents sensed by Toll-like receptors leads to upregu-lation of early effectors of the innate immune response,including chemokines, proinflammatory cytokines, and im-mune receptors (7, 91, 215). These early effectors then gen-erate a further cycle of NF-�B induction and generation ofan antimicrobial response (104). The induction of such acascade by detection of viral components within the ER, andsubsequent PERK-dependent NF-�B activation, might repre-sent an analogous antiviral response.

Developmentally, NF-�B is a prosurvival signal insome cell types, for example, in the maturation of B andT lymphocytes and in the development and regenerationof the liver (6, 103, 183, 210). While many stimuli thatinduce NF-�B also promote cell death, paradoxicallyNF-�B tends to oppose signal-induced cell death by in-duction of antiapoptotic genes including c-IAP1, c-IAP2,

and FLIP (120, 195). Indeed, NF-�B can have proliferativeeffects through its targets c-myc and cylcinD1 (142, 155).It is therefore possible that NF-�B induction may also linksecretory activity to trophic signals.

B. Long-Term Change

In contrast to the continuous modulation of ER syn-thetic activity, differentiation of a cell into a highly secre-

tory phenotype requires a dramatic expansion of ER. Thisis exemplified in the development of B-lymphocytes intoplasma cells. In this regard, defects in plasma cell differ-entiation have helped reveal an additional role for theUPR in the regulation of absolute ER mass within the cell.

1. Linking ER mass to demand

Multiple myeloma is a hematological neoplasm char-acterized by presence in the blood of a monoclonal im-munoglobin or Bence Jones protein (free monoclonal �-or �-light chains). These are produced by myeloma cellswithin the bone marrow, which over time expand todisplace normal marrow leading to anemia and immuno-logical deficiency. The plasma concentration of the mono-clonal paraprotein can reach exceeding high levels (�100g/l), even leading to complications through increasedblood viscosity. The causative myeloma cells are mono-clonal expansions of plasma cells, immune cells normallycharged with large-scale immunoglobin manufacture.

To understand the genesis of plasma cells and my-elomas, the transcription factors necessary for their dif-ferentiation have been determined. Early work demon-strated that myeloma cell lines frequently express a basicleucine zipper transcription factor, XBP-1, to extremelyhigh levels (26, 204). XBP-1 was also shown to be essen-tial for normal plasma cell differentiation (156, 157, 204).Myeloma cells and mature plasma cells share the ability tosecrete huge quantities of immunoglobin, thanks to theirhighly developed secretory pathway. Histologically thesecells have characteristically basophilic cytoplasm due tothe large numbers of ribosomes associated with theirextensive ER (Fig. 3). It seems likely that XBP-1’s role inplasma cell differentiation is related to regulation of ERexpansion in response to increased demand (96). Further-more, new therapeutic agents being developed to treatmyeloma may exploit the dependence of this neoplasm onUPR activation (2, 66). Proteosome inhibitors haveproven to be highly effective in the killing of myelomacells and in early trials have achieved remarkable clinicalremissions in otherwise drug-resistant disease (81). Theirselectivity toward myelomas might reflect the reliance ofthese cells on efficient ERAD, which itself is dependenton XBP-1 signaling (96, 221). The first drug of this class,bortezomib (Velcade), has now been approved by theFDA for use in otherwise refractory disease (17).

XBP-1 regulates many genes of the unfolded proteinresponse in metazoans, including genes involved in gen-eration of the ER membrane (177). It is upregulated at theprotein level during ER stress by a remarkable mecha-nism. The XBP-1 primary transcript is translated, albeitinefficiently, to produce a protein without significanttransactivation activity (99). ER stress-dependent splicingof XBP-1 mRNA involves excision of a short intron (26ntin mammals) causing a frame shift in its open reading

1140 STEFAN J. MARCINIAK AND DAVID RON

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 9: Endoplasmic Reticulum Stress Signaling in Disease

frame, which introduces a new COOH-terminal transacti-vation domain to generate the active transcription factor(19, 99, 169, 222). The endonuclease responsible for re-moving the inhibitory intron is an ER transmembraneprotein, IRE1, whose luminal domain shares significanthomology with that of PERK, allowing it a similar regu-latory interaction with BiP (11, 12, 86, 108, 134). As withPERK, ER stress-induced dissociation of BiP allows clus-tering of IRE1 and its activation through transautophos-phorylation. In mammals there are, in fact, two IRE1isoforms. IRE1� is detectable in all tissues (185), while

IRE1� appears limited to the gut epithelium (198). Thesignificance of this tissue-dependent isoform expressionis unclear but may point toward cell type specific differ-ences in UPR signaling. IRE1��/� animals are not viable,whereas IRE1� mutants appear well, although are moreprone to chemical-induced colitis (11, 185, 192).

The IRE1-dependent arm of the UPR is the mostancient in evolutionary terms, since the entire UPR oflower eukaryotes, including that of yeast, is regulated bythis molecule (30, 31, 40, 184, 185). Although in yeast thetarget transcription factor HAC1 bares little homology toXBP1, the activation of Ire1p is identical. It was, in fact,historically the first UPR pathway to be described inmolecular terms (14, 22, 50, 85, 123, 167, 173, 202). Inyeast, after Ire1p-mediated intron excision from theHAC1 mRNA, religation by the tRNA ligase Rlg1p gener-ates an efficiently translated message (123, 161, 172). Theanalogous ligase in higher organisms has yet to be iden-tified and remains an important unanswered question.

2. Insulin resistance and ER stress

In addition to linking secretory capacity to demandthrough regulation of ER expansion, IRE1 also plays animportant role in integrating ER stress signaling withother stress and growth factor sensitive pathways.

An important feature of type 2 diabetes is peripheralinsulin resistance. Normally, activated insulin receptorsphosphorylate proximal signaling molecules, such as in-sulin receptor substrate 1 (IRS-1), on tyrosine residues,which transduce the effects of insulin through interactionwith cytosolic targets (180, 182). In obesity (genetic ordietary), this tyrosine phosphorylation is inhibited byJNK-dependent serine phosphorylation of IRS-1 (4, 5).Surprisingly, the mechanism of obesity-related JNK acti-vation appears to involve ER stress.

For unclear reasons, liver and adipose cells fromobese mice show biochemical evidence of ER stress withPERK activation, eIF2� phosphorylation, and BiP induc-tion (139). It has previously been shown that ER stress,through IRE1 activation, can directly trigger the JNK cas-cade (192). Activated IRE1 recruits the scaffolding pro-tein TRAF2 to the ER membrane (192), which triggers amitogen-activated protein (MAP) kinase cascade leadingto JNK activation (129, 130). Thus IRE1 offers a plausiblemechanistic link between obesity and peripheral insulinresistance. In this model, obesity-associated ER stresswould contribute to insulin resistance by causing JNKactivation through IRE1/TRAF2 with subsequent IRS-1serine phosphorylation (Fig. 4). Consistent with this hy-pothesis, IRE1�/� cells fail to active JNK or phosphory-late IRS-1 during ER stress unlike wild-type controls.Supporting a causal link between peripheral ER stressand insulin resistance, a recent study has demonstratedprotection against obesity-induced type 2 diabetes in mice

FIG. 3. The unfolded protein response (UPR). A: photomicrograph(�1,000) of neoplastic plasma cells stained with eosin and hematoxylinfrom a bone marrow aspirate in a case of multiple myeloma. Note thelateral displacement of the nucleus by the extensive basophilic ER.Scale bar is 10 �m. (Image kindly provided by Dr. Wendy Erber, Adden-brookes Hospital, UK.) B: ER stress causes signaling through threedistinct mediator proteins in the ER membrane offering redundancyand, perhaps, a means to adjust the UPR to fit the circumstances of thestress. Activated PERK phosphorylates eIF2� and inhibits new proteintranslation. Additionally, ATF4 induction activates genes of the ISR,including GADD34, which relieves translational attenuation, andERO1�, which promotes oxidataive protein folding. Antioxidant targetsof ATF4 act to buffer increased reactive oxygen species (ROS) producedby ERO1�. ATF6 cleavage releases soluble ATF6c allowing transactiva-tion of UPR target genes that increase the ER protein maturation ma-chinery. IRE1-mediated activation of XBP1, which is more sustained dueto XBP1 autoinduction, leads to induction of genes involved both in thematuration of ER client proteins and also genes that promote ER asso-ciated protein degradation.

ENDOPLASMIC RETICULUM AND DISEASE 1141

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 10: Endoplasmic Reticulum Stress Signaling in Disease

by overexpression of an ER chaperone, while knockdownof the same chaperone was diabetogenic (128). Further-more, in the Akita mouse, systemic rather than �-celloverexpression of an ER chaperone appeared to improveperipheral insulin sensitivity (138).

In light of this, it is possible that the peripheralinsulin resistance seen in the eIF2�S51A heterozygousmice (above) might have resulted from increased signal-ing through the IRE1-JNK arm of the UPR attempting tocompensate for impaired PERK-dependent signaling.Consistent with this speculation, PERK�/� cells, whichare impaired in eIF2� phosphorylation, experience ba-sally higher IRE1 activation (55).

It can be seen that the role of UPR signaling indiabetes is far from straightforward. At the level of thepancreatic �-cell, the UPR protects against developingtype I diabetes by ensuring that this highly secretorytissue maintains efficient function despite wide swings inER protein flux. Conversely, chronic ER stress signaling(perhaps related to obesity) appears involved in the eti-ology of peripheral insulin resistance. We can at presentonly speculate as to how the UPR progresses from regu-lation to dysregulation over the passing years. It is knownthat chronic exposure to elevated saturated free fattyacids induces ER stress (83). Consequently, obesity-re-lated ER stress may reflect the evolutionarily recent phe-nomenon of chronic nutrient excess. The dysregulation ofER stress signaling seen in this disease may reflect thelack of selection pressure to adapt this homeostatic mech-anism to chronic inappropriate activation.

3. ER stress and lipid metabolism

A) HOMOCYSTEINE AND ATHEROSCLEROSIS. Among theknown risk factors for atherosclerosis, ER stress wouldcurrently be ranked low, if at all. This may begin tochange as evidence accumulates implicating the UPR indisordered lipid metabolism. The relationship betweenthe UPR and lipid accumulation is a complex one. On theone hand, many target genes of the UPR involve lipidsynthesis, in part to allow expansion of the ER itself; onthe other hand, there is increasing evidence that pertur-bation of the lipid environment within the ER can activateUPR signaling.

Homocysteinemia is associated with the develop-ment of atherosclerosis, certainly in patients with rareinherited disorders of amino acid metabolism, but also inthe wider population, as indicated by large epidemiolog-ical studies (42, 70, 201). Although homocysteine is asulfur-containing amino acid, some of its toxic effectsappear to involve dysregulation of cholesterol and triglyc-eride biosynthesis (205). Patients with inherited hyperho-mocysteinemia or laboratory animals fed a homocysteine-rich diet develop hepatic steatosis in the absence ofmarked rises in plasma lipids. The likely explanation islocal lipid synthesis, and consistent with this, homocys-teine has been shown to activate lipogenic signaling viathe sterol regulated element-binding proteins (SREBPs)(205). Surprisingly, ER stress contributes to the activationof SREBP by homocysteine. Livers of homocysteine-fedmice contain raised levels of ER chaperones, as do cul-tured cells exposed to high levels of homocysteine invitro. The importance of ER stress in this lipogenic sig-naling was demonstrated by BiP overexpression, whichameliorated SREBP induction in response to homocys-teine.

The relationship between the UPR and SREBP, whileclearly evident, is still not entirely explained in mechanis-

FIG. 4. Insulin resistance in obesity mediated by ER stress. A: innonobese subjects, glucose homeostasis is maintained by effectors ofinsulin signaling. Insulin binding to its receptor causes tyrosine phos-phorylation of IRS1, which in turn promotes activation of effectorproteins. B: in obesity, ER stress causes insulin resistance. IRE1 activa-tion may then recruit TRAF2 to the ER membrane causing activation ofJNK, which in turn would promote phosphorylation of serine residues inIRS1. Serine phosphorylation of IRS1 inhibits its tyrosine phosphoryla-tion by activated insulin receptors, thus impairing insulin signaling.

1142 STEFAN J. MARCINIAK AND DAVID RON

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 11: Endoplasmic Reticulum Stress Signaling in Disease

tic terms. The SREBPs are transmembrane proteins that,under sterol-sufficient conditions, are retained within theER through interaction with other ER membrane pro-teins, Insig-1 and Insig-2 (213, 216). When cholesterollevels are low, this interaction is weakened, whereuponSREBP progresses to the Golgi to be cleaved sequentiallyby two serine proteases, S1P and S2P (33). This cleavageliberates the cytosolic portion of SREBP as a solubleprotein, which enters the nucleus to bind specific DNAelements, thus transactivating target genes involved inlipid synthesis. One possible mechanism of ER stress-induced SREBP activation, for which there is evidence,involves stress-dependent loss of Insig-1 from the ER (98).

B) SREBP AND ATF6. Interestingly, the third known classof ER stress sensor, exemplified by ATF6, is unrelated toPERK or IRE1. ATF6 is another ER transmembrane pro-tein, that undergoes processing similar to SREBP (24, 60,102). It is normally retained within the ER through inter-action, not with Insig proteins, but with BiP (168). DuringER stress, ATF6 is released from the ER and moves on tothe Golgi. There it is processed by the very same pro-teases that act on SREBP, to liberate a soluble transcrip-tion factor (218). Unlike SREBP, the targets of ATF6 areclassical UPR genes (133, 199, 220). Indeed, ATF6 wasisolated in a search for ligands of an ER stress responsiveDNA element (ESRE). Despite their separate transcrip-tional programs, there may still be some cross-talk be-tween ATF6 and SREBP. A recent report suggests thatwithin the nucleus, cleaved ATF6 can bind SREBP2 di-rectly to inhibit its transactivation potential and opposelipogenesis (224).

Following from the identification of ATF6, there havebeen several related transcription factors identified,which share similar domain structures and localization tothe ER [Luman (35, 113, 154), OASIS (88), CREBH (225),ATF6b (59), CREB4 (178)]. The existence of such a vari-ety of UPR signaling molecules at first seems puzzling, buta clue to their roles may come from tissue distribution. Inthe cases of OASIS and CREBH, at least, there is clearevidence for tissue specificity, OASIS being found in as-trocytes of the central nervous system, while CREBH isrestricted to the liver. This raises the possibility that ERstress may elicit tissue-specific transcriptional responses.In the case of CREBH for example, it has already beenshown that its activation by ER stress leads to inductionof genes of the acute phase response, providing a directlink between hepatic ER stress and systemic inflamma-tion (225).

C) CHOLESTEROL’S IMPACT ON THE ER. The dependence ofcholesterol metabolism on ER stress signaling may reflectthe impact of sterols on ER function. This is well illus-trated by the macrophage in development of atheroscle-rosis. During the initial phase of atherosclerotic lesionformation, foam cells are a characteristic histologicalfinding within the vessel intima. These are macrophages

that have taken up oxidized lipoprotein particles andbecome laden with cholesterol. This cholesterol is storedas esters within large lipid vesicles, which give foam cellstheir foamy appearance. Over time, death of these cells,likely due to the toxic effects of unesterified cholesterol,results in deposition of extracellular cholesterol withinthe plaque. Unlike its esters, free cholesterol efficientlyinserts into lipid bilayers and can alter membrane physi-cal properties. The ER membrane, being poor in freecholesterol (16), may be especially sensitive to choles-terol loading. Indeed, recent findings indicate that for freecholesterol to have its toxic effect it must be trafficked tothe ER (45). This trafficking results in activation of UPRsignaling and caspase activation, and eventually in mac-rophage apoptosis. In addition, ER stress caused by freecholesterol loading of macrophages promotes chemokinesecretion, and this may contribute to the formation of“vulnerable” atherosclerotic lesion, which are prone torupture as a function of their high inflammatory cell infil-trate (105).

III. CONCLUDING REMARKS

The complexity of the mammalian UPR, mediated asit is by three distinct classes of ER sensors and numeroustranscription factors, may provide a degree of redun-dancy; however, perhaps a more interesting interpreta-tion of this complexity may be to allow higher organismsa more subtle response to ER stress. For example, thedivision of UPR genes between multiple transcriptionfactors likely enables portions of the gene expressionprogram to be induced as is appropriate either to theduration or intensity of ER stress. While ATF6 and XBP-1transactivate many of the same genes, their expressionprograms show important differences (97). Binding ofATF6 to the BiP promoter is sufficient for maximal trans-activation, while XBP-I binding is sufficient for maximalinduction of EDEM (97, 221). In contrast, many genesrequire both arms to be active in order for best induction(214). It has been suggested that an early, ATF6-domi-nated, response may attempt to compensate for ER stressthrough chaperone induction alone. While XBP-1 activa-tion, which is more sustained, may, in addition, promoteprotein degradation through the induction of ERAD com-ponents. Accordingly, IRE1��/� cells have been showndefective in glycoprotein ERAD (221).

Repeatedly, models of ER dysfunction have gener-ated models of diabetes, and investigations into rare ge-netic diabetic syndromes have revealed unexpected linksto ER stress. The field remains fertile for further study,since mutations in ER stress-inducible genes such asWFS1 (Wolfram syndrome) (74, 179, 181) and P58IPK

(93), which lead to diabetic syndromes, have yet to befully understood. Furthermore, the study of apparentlyunrelated disease states including atherosclerosis, viral

ENDOPLASMIC RETICULUM AND DISEASE 1143

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 12: Endoplasmic Reticulum Stress Signaling in Disease

infection, and multiple myeloma may also benefit fromour increasing understanding of ER stress in their patho-physiology.

ACKNOWLEDGMENTS

Address for reprint requests and other correspondence:S. J. Marciniak, Cambridge Institute for Medical Research, Univ.of Cambridge, Wellcome Trust/MRC Building, Hills Road, Cam-bridge CB2 2XY, UK (e-mail: [email protected]); and D. Ron,Skirball Institute of Biomolecular Medicine, New York Univer-sity School of Medicine, New York, NY 10016 (e-mail:[email protected]).

REFERENCES

1. Abastado JP, Miller PF, Jackson BM, and Hinnebusch AG.

Suppression of ribosomal reinitiation at upstream open readingframes in amino acid-starved cells forms the basis for GCN4 trans-lational control. Mol Cell Biol 11: 486–496, 1991.

2. Adams J. Proteasome inhibition in cancer: development of PS-341.Semin Oncol 28: 613–619, 2001.

3. Adesnik M, Lande M, Martin T, and Sabatini DD. Retention ofmRNA on the endoplasmic reticulum membranes after in vivodisassembly of polysomes by an inhibitor of initiation. J Cell Biol

71: 307–313, 1976.4. Aguirre V, Uchida T, Yenush L, Davis R, and White MF. The

c-Jun NH (2)-terminal kinase promotes insulin resistance duringassociation with insulin receptor substrate-1 and phosphorylationof Ser (307). J Biol Chem 275: 9047–9054, 2000.

5. Aguirre V, Werner ED, Giraud J, Lee YH, Shoelson SE, and

White MF. Phosphorylation of Ser307 in insulin receptor sub-strate-1 blocks interactions with the insulin receptor and inhibitsinsulin action. J Biol Chem 277: 1531–1537, 2002.

6. Aifantis I, Gounari F, Scorrano L, Borowski C, and von Boeh-

mer H. Constitutive pre-TCR signaling promotes differentiationthrough Ca2� mobilization and activation of NF-kappaB and NFAT.Nat Immunol 2: 403–409, 2001.

7. Alexopoulou L, Holt AC, Medzhitov R, and Flavell RA. Rec-ognition of double-stranded RNA and activation of NF-kappaB byToll-like receptor 3. Nature 413: 732–738, 2001.

8. Allen JR, Nguyen LX, Sargent KE, Lipson KL, Hackett A, and

Urano F. High ER stress in beta-cells stimulates intracellular deg-radation of misfolded insulin. Biochem Biophys Res Commun 324:166–170, 2004.

9. Berlanga JJ, Santoyo J, and De Haro C. Characterization of amammalian homolog of the GCN2 eukaryotic initiation factor 2al-pha kinase. Eur J Biochem 265: 754–762, 1999.

10. Berry MJ, Knutson GS, Lasky SR, Munemitsu SM, and Samuel

CE. Mechanism of interferon action. Purification and substratespecificities of the double-stranded RNA-dependent protein kinasefrom untreated and interferon-treated mouse fibroblasts. J Biol

Chem 260: 11240–11247, 1985.11. Bertolotti A, Wang X, Novoa I, Jungreis R, Schlessinger K,

Cho JH, West AB, and Ron D. Increased sensitivity to dextransodium sulfate colitis in IRE1b deficient mice. J Clin Invest 107:585–593, 2001.

12. Bertolotti A, Zhang Y, Hendershot L, Harding H, and Ron D.

Dynamic interaction of BiP and the ER stress transducers in theunfolded protein response. Nat Cell Biol 2: 326–332, 2000.

13. Bole D, Hendershot L, and Kearney J. Posttranslational associ-ation of immunoglobulin heavy chain binding protein with thenascent heavy chains in nonsecreting and secreting hybridomas.J Cell Biol 102: 1558–1566, 1986.

14. Bork P and Sander C. A hybrid protein kinase-RNase in aninterferon-induced pathway? FEBS Lett 334: 149–152, 1993.

15. Boyce M, Bryant KF, Jousse C, Long K, Harding HP, Scheuner

D, Kaufman RJ, Ma D, Coen D, Ron D, and Yuan J. A selectiveinhibitor of eIF2� dephosphorylation protects cells from ER stress.Science 307: 935–939, 2005.

16. Bretscher MS and Munro S. Cholesterol and the Golgi apparatus.Science 261: 1280–1281, 1993.

17. Bross PF, Kane R, Farrell AT, Abraham S, Benson K, Brower

ME, Bradley S, Gobburu JV, Goheer A, Lee SL, Leighton J,

Liang CY, Lostritto RT, McGuinn WD, Morse DE, Rahman A,

Rosario LA, Verbois SL, Williams G, Wang YC, and Pazdur R.

Approval summary for bortezomib for injection in the treatment ofmultiple myeloma. Clin Cancer Res 10: 3954–3964, 2004.

18. Brush MH, Weiser DC, and Shenolikar S. Growth arrest andDNA damage-inducible protein GADD34 targets protein phospha-tase 1alpha to the endoplasmic reticulum and promotes dephos-phorylation of the alpha subunit of eukaryotic translation initiationfactor 2. Mol Cell Biol 23: 1292–1303, 2003.

19. Calfon M, Zeng H, Urano F, Till JH, Hubbard SR, Harding HP,

Clark SG, and Ron D. IRE1 couples endoplasmic reticulum loadto secretory capacity by processing the XBP-1 mRNA. Nature 415:92–96, 2002.

20. Cardozo AK, Ortis F, Storling J, Feng YM, Rasschaert J,

Tonnesen M, Van Eylen F, Mandrup-Poulsen T, Herchuelz A,

and Eizirik DL. Cytokines downregulate the sarcoendoplasmicreticulum pump Ca2� ATPase 2b and deplete endoplasmic reticu-lum Ca2�, leading to induction of endoplasmic reticulum stress inpancreatic beta-cells. Diabetes 54: 452–461, 2005.

21. Carroll K, Elroy-Stein O, Moss B, and Jagus R. Recombinantvaccinia virus K3L gene product prevents activation of double-stranded RNA-dependent, initiation factor 2 alpha-specific proteinkinase. J Biol Chem 268: 12837–12842, 1993.

22. Chapman RE and Walter P. Translational attenuation mediatedby an mRNA intron. Curr Biol 7: 850–859, 1997.

23. Chen JJ, Throop MS, Gehrke L, Kuo I, Pal JK, Brodsky M, and

London IM. Cloning of the cDNA of the heme-regulated eukaryoticinitiation factor 2 alpha (eIF-2 alpha) kinase of rabbit reticulo-cytes: homology to yeast GCN2 protein kinase and human double-stranded-RNA-dependent eIF-2 alpha kinase. Proc Natl Acad Sci

USA 88: 7729–7733, 1991.24. Chen X, Shen J, and Prywes R. The luminal domain of ATF6

senses endoplasmic reticulum (ER) stress and causes translocationof ATF6 from the ER to the Golgi. J Biol Chem 277: 13045–13052,2002.

25. Cheung MS, Garcia AE, and Onuchic JN. Protein folding medi-ated by solvation: water expulsion and formation of the hydropho-bic core occur after the structural collapse. Proc Natl Acad Sci USA

99: 685–690, 2002.26. Claudio JO, Masih-Khan E, Tang H, Goncalves J, Voralia M, Li

ZH, Nadeem V, Cukerman E, Francisco-Pabalan O, Liew CC,

Woodgett JR, and Stewart AK. A molecular compendium ofgenes expressed in multiple myeloma. Blood 100: 2175–2186, 2002.

27. Clemens M. Protein kinases that phosphorylate eIF2 and eIF2B,their role in eukaryotic cell translational control. In: Translational

Control, edited by Hershey J, Mathews M, and Sonenberg N. ColdSpring Harbor, NY: Cold Spring Harbor Laboratory Press, 1996, p.139–172.

28. Cohen PT. Protein phosphatase 1—targeted in many directions.J Cell Sci 115: 241–256, 2002.

29. Connor JH, Weiser DC, Li S, Hallenbeck JM, and Shenolikar

S. Growth arrest and DNA damage-inducible protein gadd34 as-sembles a novel signaling complex containing protein phosphatase1 and inhibitor 1. Mol Cell Biol 21: 6841–6850, 2001.

30. Cox JS, Shamu CE, and Walter P. Transcriptional induction ofgenes encoding endoplasmic reticulum resident proteins requires atransmembrane protein kinase. Cell 73: 1197–1206, 1993.

31. Cox JS and Walter P. A novel mechanism for regulating activityof a transcription factor that controls the unfolded protein re-sponse. Cell 87: 391–404, 1996.

32. Credle JJ, Finer-Moore JS, Papa FR, Stroud RM, and Walter

P. Inaugural article: on the mechanism of sensing unfolded proteinin the endoplasmic reticulum. Proc Natl Acad Sci USA 102: 18773–18784, 2005.

33. DeBose-Boyd RA, Brown MS, Li WP, Nohturfft A, Goldstein

JL, and Espenshade PJ. Transport-dependent proteolysis ofSREBP: relocation of site-1 protease from Golgi to ER obviates theneed for SREBP transport to Golgi. Cell 99: 703–712, 1999.

1144 STEFAN J. MARCINIAK AND DAVID RON

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 13: Endoplasmic Reticulum Stress Signaling in Disease

34. Delepine M, Nicolino M, Barrett T, Golamaully M, Lathrop

GM, and Julier C. EIF2AK3, encoding translation initiation factor2-alpha kinase 3, is mutated in patients with Wolcott-Rallison syn-drome. Nat Genet 25: 406–409, 2000.

35. DenBoer LM, Hardy-Smith PW, Hogan MR, Cockram GP, Au-

das TE, and Lu R. Luman is capable of binding and activatingtranscription from the unfolded protein response element. Bio-

chem Biophys Res Commun 331: 113–119, 2005.36. Deng J, Lu PD, Zhang Y, Scheuner D, Kaufman RJ, Sonenberg

N, Harding HP, and Ron D. Translational repression mediatesactivation of Nuclear Factor kappa B by phosphorylated transla-tion initiation factor 2. Mol Cell Biol 24: 10161–10168, 2004.

37. Dever TE, Feng L, Wek RC, Cigan AM, Donahue TF, and

Hinnebusch AG. Phosphorylation of initiation factor 2 alpha byprotein kinase GCN2 mediates gene-specific translational controlof GCN4 in yeast. Cell 68: 585–596, 1992.

38. De Virgilio M, Kitzmuller C, Schwaiger E, Klein M, Kreibich

G, and Ivessa NE. Degradation of a short-lived glycoprotein fromthe lumen of the endoplasmic reticulum: the role of N-linked gly-cans and the unfolded protein response. Mol Biol Cell 10: 4059–4073, 1999.

39. Dobson CM. Principles of protein folding, misfolding and aggre-gation. Semin Cell Dev Biol 15: 3–16, 2004.

40. Dong B and Silverman RH. Alternative function of a proteinkinase homology domain in 2�,5�-oligoadenylate dependent RNaseL. Nucleic Acids Res 27: 439–445, 1999.

41. Dorner A, Wasley L, and Kaufman R. Overexpression of GRP78mitigates stress induction of glucose regulated proteins and blockssecretion of selective proteins in Chinese hamster ovary cells.EMBO J 11: 1563–1571, 1992.

42. Eikelboom JW, Lonn E, Genest J Jr, Hankey G, and Yusuf S.

Homocyst(e)ine and cardiovascular disease: a critical review of theepidemiologic evidence. Ann Intern Med 131: 363–375, 1999.

43. Eriksson KK, Vago R, Calanca V, Galli C, Paganetti P, and

Molinari M. EDEM contributes to maintenance of protein foldingefficiency and secretory capacity. J Biol Chem 279: 44600–44605,2004.

44. Fagard R and London IM. Relationship between phosphorylationand activity of heme-regulated eukaryotic initiation factor 2 alphakinase. Proc Natl Acad Sci USA 78: 866–870, 1981.

45. Feng B, Yao PM, Li Y, Devlin CM, Zhang D, Harding HP,

Sweeney M, Rong JX, Kuriakose G, Fisher EA, Marks AR, Ron

D, and Tabas I. The endoplasmic reticulum as the site of choles-terol-induced cytotoxicity in macrophages. Nat Cell Biol 5: 781–792, 2003.

46. Fogli A, Schiffmann R, Bertini E, Ughetto S, Combes P, Ey-

mard-Pierre E, Kaneski CR, Pineda M, Troncoso M, Uziel G,

Surtees R, Pugin D, Chaunu MP, Rodriguez D, and Boespflug-

Tanguy O. The effect of genotype on the natural history of eIF2B-related leukodystrophies. Neurology 62: 1509–1517, 2004.

47. Fogli A, Wong K, Eymard-Pierre E, Wenger J, Bouffard JP,

Goldin E, Black DN, Boespflug-Tanguy O, and Schiffmann R.

Cree leukoencephalopathy and CACH/VWM disease are allelic atthe EIF2B5 locus. Ann Neurol 52: 506–510, 2002.

48. Frand AR and Kaiser CA. The ERO1 gene of yeast is required foroxidation of protein dithiols in the endoplasmic reticulum. Mol Cell

1: 161–170, 1998.49. Gardner RG, Shearer AG, and Hampton RY. In vivo action of

the HRD ubiquitin ligase complex: mechanisms of endoplasmicreticulum quality control and sterol regulation. Mol Cell Biol 21:4276–4291, 2001.

50. Gonzalez TN, Sidrauski C, Dorfler S, and Walter P. Mechanismof non-spliceosomal mRNA splicing in the unfolded protein re-sponse pathway. EMBO J 18: 3119–3132, 1999.

51. Hampton RY, Gardner RG, and Rine J. Role of 26S proteasomeand HRD genes in the degradation of 3-hydroxy-3-methylglutaryl-CoA reductase, an integral endoplasmic reticulum membrane pro-tein. Mol Biol Cell 7: 2029–2044, 1996.

52. Han AP, Yu C, Lu L, Fujiwara Y, Browne C, Chin G, Fleming

M, Leboulch P, Orkin SH, and Chen JJ. Heme-regulatedeIF2alpha kinase (HRI) is required for translational regulation andsurvival of erythroid precursors in iron deficiency. EMBO J 20:6909–6918, 2001.

53. Harding H, Novoa I, Zhang Y, Zeng H, Wek RC, Schapira M,

and Ron D. Regulated translation initiation controls stress-inducedgene expression in mammalian cells. Mol Cell 6: 1099–1108, 2000.

54. Harding H, Zeng H, Zhang Y, Jungreis R, Chung P, Plesken H,

Sabatini D, and Ron D. Diabetes mellitus and excocrine pancre-atic dysfunction in Perk�/� mice reveals a role for translationalcontrol in survival of secretory cells. Mol Cell 7: 1153–1163, 2001.

55. Harding H, Zhang Y, Bertolotti A, Zeng H, and Ron D. Perk isessential for translational regulation and cell survival during theunfolded protein response. Mol Cell 5: 897–904, 2000.

56. Harding H, Zhang Y, and Ron D. Translation and protein foldingare coupled by an endoplasmic reticulum resident kinase. Nature

397: 271–274, 1999.57. Harding H, Zhang Y, Zeng H, Novoa I, Lu P, Calfon M, Sadri N,

Yun C, Popko B, Paules R, Stojdl D, Bell J, Hettmann T,

Leiden J, and Ron D. An integrated stress response regulatesamino acid metabolism and resistance to oxidative stress. Mol Cell

11: 619–633, 2003.58. Haynes CM, Titus EA, and Cooper AA. Degradation of mis-

folded proteins prevents ER-derived oxidative stress and celldeath. Mol Cell 15: 767–776, 2004.

59. Haze K, Okada T, Yoshida H, Yanagi H, Yura T, Negishi M, and

Mori K. Identification of the G13 (cAMP-response-element-bindingprotein-related protein) gene product related to activating tran-scription factor 6 as a transcriptional activator of the mammalianunfolded protein response. Biochem J 355: 19–28, 2001.

60. Haze K, Yoshida H, Yanagi H, Yura T, and Mori K. Mammaliantranscription factor ATF6 is synthesized as a transmembrane pro-tein and activated by proteolysis in response to endoplasmic retic-ulum stress. Mol Biol Cell 10: 3787–3799, 1999.

61. He B, Gross M, and Roizman B. The gamma134 5 protein ofherpes simplex virus 1 has the structural and functional attributesof a protein phosphatase 1 regulatory subunit and is present in ahigh molecular weight complex with the enzyme in infected cells.J Biol Chem 273: 20737–20743, 1998.

62. He B, Gross M, and Roizman B. The gamma (1)345 protein ofherpes simplex virus 1 complexes with protein phosphatase 1alphato dephosphorylate the alpha subunit of the eukaryotic translationinitiation factor 2 and preclude the shutoff of protein synthesis bydouble-stranded RNA-activated protein kinase. Proc Natl Acad Sci

USA 94: 843–848, 1997.63. Helenius A and Aebi M. Roles of N-linked glycans in the endo-

plasmic reticulum. Annu Rev Biochem 73: 1019–1049, 2004.64. Hendershot L, Wei J, Gaut J, Melnick J, Aviel S, and Argon Y.

Inhibition of immunoglobulin folding and secretion by dominantnegative BiP ATPase mutants. Proc Natl Acad Sci USA 93: 5269–5274, 1996.

65. Hendershot LM, Wei JY, Gaut JR, Lawson B, Freiden PJ, and

Murti KG. In vivo expression of mammalian BiP ATPase mutantscauses disruption of the endoplasmic reticulum. Mol Biol Cell 6:283–296, 1995.

66. Hideshima T, Richardson P, Chauhan D, Palombella VJ, El-

liott PJ, Adams J, and Anderson KC. The proteasome inhibitorPS-341 inhibits growth, induces apoptosis, and overcomes drugresistance in human multiple myeloma cells. Cancer Res 61: 3071–3076, 2001.

67. Hidvegi T, Schmidt BZ, Hale P, and Perlmutter DH. Accumu-lation of mutant alpha 1 antitrypsin Z in the ER activates caspases-4and -12, NFkappa B and BAP31 but not the unfolded proteinresponse. J Biol Chem 2005.

68. Hinnebusch AG. Gene-specific translational control of the yeastGCN4 gene by phosphorylation of eukaryotic initiation factor 2.Mol Microbiol 10: 215–223, 1993.

69. Hinnebusch AG. Mechanism and regulation of initiator methionyl-tRNA binding to ribosomes. In: Translational Control of Gene

Expression, edited by Sonenberg N, Hershey JWB, and MathewsMB. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press,2000, p. 185–243.

70. Hofmann MA, Lalla E, Lu Y, Gleason MR, Wolf BM, Tanji N,

Ferran LJ Jr, Kohl B, Rao V, Kisiel W, Stern DM, and Schmidt

AM. Hyperhomocysteinemia enhances vascular inflammation andaccelerates atherosclerosis in a murine model. J Clin Invest 107:675–683, 2001.

ENDOPLASMIC RETICULUM AND DISEASE 1145

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 14: Endoplasmic Reticulum Stress Signaling in Disease

71. Holtz WA and O’Malley KL. Parkinsonian mimetics induce as-pects of unfolded protein response in death of dopaminergic neu-rons. J Biol Chem 278: 19367–19377, 2003.

72. Hosokawa N, Tremblay LO, You Z, Herscovics A, Wada I, and

Nagata K. Enhancement of endoplasmic reticulum (ER) degrada-tion of misfolded Null Hong Kong alpha1-antitrypsin by human ERmannosidase I. J Biol Chem 278: 26287–26294, 2003.

73. Hosokawa N, Wada I, Hasegawa K, Yorihuzi T, Tremblay LO,

Herscovics A, and Nagata K. A novel ER alpha-mannosidase-likeprotein accelerates ER-associated degradation. EMBO Rep 2: 415–422, 2001.

74. Inoue H, Tanizawa Y, Wasson J, Behn P, Kalidas K, Bernal-

Mizrachi E, Mueckler M, Marshall H, Donis-Keller H, Crock P,

Rogers D, Mikuni M, Kumashiro H, Higashi K, Sobue G, Oka

Y, and Permutt MA. A gene encoding a transmembrane protein ismutated in patients with diabetes mellitus and optic atrophy (Wol-fram syndrome). Nat Genet 20: 143–148, 1998.

75. Itoh N and Okamoto H. Translational control of proinsulin syn-thesis by glucose. Nature 283: 100–102, 1980.

76. Jakob CA, Burda P, Roth J, and Aebi M. Degradation of mis-folded endoplasmic reticulum glycoproteins in Saccharomyces cer-

evisiae is determined by a specific oligosaccharide structure. J Cell

Biol 142: 1223–1233, 1998.77. Jiang HY and Wek RC. GCN2 phosphorylation of eIF2alpha acti-

vates NF-kappaB in response to UV irradiation. Biochem J 385:371–380, 2005.

78. Jiang HY, Wek SA, McGrath BC, Scheuner D, Kaufman RJ,

Cavener DR, and Wek RC. Phosphorylation of the alpha subunitof eukaryotic initiation factor 2 is required for activation of NF-kappaB in response to diverse cellular stresses. Mol Cell Biol 23:5651–5663, 2003.

79. Jimenez JL, Nettleton EJ, Bouchard M, Robinson CV, Dobson

CM, and Saibil HR. The protofilament structure of insulin amyloidfibrils. Proc Natl Acad Sci USA 99: 9196–9201, 2002.

80. Jousse C, Oyadomari S, Novoa I, Lu PD, Zhang Y, Harding HP,

and Ron D. Inhibition of a constitutive translation initiation factor2a phosphatase, CReP, promotes survival of stressed cells. J Cell

Biol 163: 767–775, 2003.81. Kane RC, Bross PF, Farrell AT, and Pazdur R. Velcade: US FDA

approval for the treatment of multiple myeloma progressing onprior therapy. Oncologist 8: 508–513, 2003.

82. Kantor L, Harding HP, Ron D, Schiffmann R, Kaneski CR,

Kimball SR, and Elroy-Stein O. Heightened stress response inprimary fibroblasts expressing mutant eIF2B genes from CACH/VWM leukodystrophy patients. Hum Genet 1–8, 2005.

83. Karaskov E, Scott C, Zhang L, Teodoro T, Ravazzola M, and

Volchuk A. Chronic palmitate but not oleate exposure inducesendoplasmic reticulum stress which may contribute to INS-1 pan-creatic �-cell apoptosis. Endocrinology. In press.

84. Karplus M and Weaver DL. Protein folding dynamics: the diffu-sion-collision model and experimental data. Protein Sci 3: 650–668, 1994.

85. Kawahara T, Yanagi H, Yura T, and Mori K. Endoplasmic retic-ulum stress-induced mRNA splicing permits synthesis of transcrip-tion factor Hac1p/Ern4p that activates the unfolded protein re-sponse. Mol Biol Cell 8: 1845–1862, 1997.

86. Kimata Y, Kimata YI, Shimizu Y, Abe H, Farcasanu IC, Takeu-

chi M, Rose MD, and Kohno K. Genetic evidence for a role ofBiP/Kar2 that regulates Ire1 in response to accumulation of un-folded proteins. Mol Biol Cell 14: 2559–2569, 2003.

87. Kojima E, Takeuchi A, Haneda M, Yagi F, Hasegawa T, Ya-

maki KI, Takeda K, Akira S, Shimokata K, and Isobe KI. Thefunction of GADD34 is a recovery from a shutoff of protein syn-thesis induced by ER stress-elucidation by GADD34-deficient mice.FASEB J 17: 1573–1575, 2003.

88. Kondo S, Murakami T, Tatsumi K, Ogata M, Kanemoto S,

Otori K, Iseki K, Wanaka A, and Imaizumi K. OASIS, a CREB/ATF-family member, modulates UPR signalling in astrocytes. Nat

Cell Biol 7: 186–194, 2005.89. Koo EH, Lansbury PT Jr, and Kelly JW. Amyloid diseases:

abnormal protein aggregation in neurodegeneration. Proc Natl

Acad Sci USA 96: 9989–9990, 1999.

90. Kostura M and Mathews MB. Purification and activation of thedouble-stranded RNA-dependent eIF-2 kinase DAI. Mol Cell Biol 9:1576–1586, 1989.

91. Krappmann D, Wegener E, Sunami Y, Esen M, Thiel A, Mord-

muller B, and Scheidereit C. The IkappaB kinase complex andNF-kappaB act as master regulators of lipopolysaccharide-inducedgene expression and control subordinate activation of AP-1. Mol

Cell Biol 24: 6488–6500, 2004.92. Kurzchalia TV, Wiedmann M, Girshovich AS, Bochkareva ES,

Bielka H, and Rapoport TA. The signal sequence of nascentpreprolactin interacts with the 54K polypeptide of the signal rec-ognition particle. Nature 320: 634–636, 1986.

93. Ladiges WC, Knoblaugh SE, Morton JF, Korth MJ, Sopher BL,

Baskin CR, MacAuley A, Goodman AG, LeBoeuf RC, and

Katze MG. Pancreatic beta-cell failure and diabetes in mice with adeletion mutation of the endoplasmic reticulum molecular chaper-one gene P58IPK. Diabetes 54: 1074–1081, 2005.

94. Lande MA, Adesnik M, Sumida M, Tashiro Y, and Sabatini DD.

Direct association of messenger RNA with microsomal membranesin human diploid fibroblasts. J Cell Biol 65: 513–528, 1975.

95. Lawless MW, Greene CM, Mulgrew A, Taggart CC, O’Neill SJ,

and McElvaney NG. Activation of endoplasmic reticulum-specificstress responses associated with the conformational disease Zalpha 1-antitrypsin deficiency. J Immunol 172: 5722–5726, 2004.

96. Lee AH, Iwakoshi NN, Anderson KC, and Glimcher LH. Pro-teasome inhibitors disrupt the unfolded protein response in my-eloma cells. Proc Natl Acad Sci USA 100: 9946–9951, 2003.

97. Lee AH, Iwakoshi NN, and Glimcher LH. XBP-1 regulates asubset of endoplasmic reticulum resident chaperone genes in theunfolded protein response. Mol Cell Biol 23: 7448–7459, 2003.

98. Lee JN and Ye J. Proteolytic activation of sterol regulatory ele-ment-binding protein induced by cellular stress through depletionof Insig-1. J Biol Chem 279: 45257–45265, 2004.

99. Lee K, Tirasophon W, Shen X, Michalak M, Prywes R, Okada

T, Yoshida H, Mori K, and Kaufman RJ. IRE1-mediated uncon-ventional mRNA splicing and S2P-mediated ATF6 cleavage mergeto regulate XBP1 in signaling the unfolded protein response. Genes

Dev 16: 452–466, 2002.100. Leegwater PA, Vermeulen G, Konst AA, Naidu S, Mulders J,

Visser A, Kersbergen P, Mobach D, Fonds D, van Berkel CG,

Lemmers RJ, Frants RR, Oudejans CB, Schutgens RB, Pronk

JC, and van der Knaap MS. Subunits of the translation initiationfactor eIF2B are mutant in leukoencephalopathy with vanishingwhite matter. Nat Genet 29: 383–388, 2001.

101. Leroux L, Desbois P, Lamotte L, Duvillie B, Cordonnier N,

Jackerott M, Jami J, Bucchini D, and Joshi RL. Compensatoryresponses in mice carrying a null mutation for Ins1 or Ins2. Dia-

betes 50 Suppl 1: S150–S153, 2001.102. Li M, Baumeister P, Roy B, Phan T, Foti D, Luo S, Lee AS,

Wang Y, Shen J, Arenzana N, Tirasophon W, Kaufman RJ, and

Prywes R. ATF6 as a transcription activator of the endoplasmicreticulum stress element: thapsigargin stress-induced changes andsynergistic interactions with NF-Y and YY1. Mol Cell Biol 20: 5096–5106, 2000.

103. Li Q, Van Antwerp D, Mercurio F, Lee KF, and Verma IM.

Severe liver degeneration in mice lacking the IkappaB kinase 2gene. Science 284: 321–325, 1999.

104. Li X and Stark GR. NFkappaB-dependent signaling pathways.Exp Hematol 30: 285–296, 2002.

105. Li Y, Schwabe RF, DeVries-Seimon T, Yao PM, Gerbod-Gian-

none MC, Tall AR, Davis RJ, Flavell R, Brenner DA, and

Tabas I. Free cholesterol-loaded macrophages are an abundantsource of tumor necrosis factor-alpha and interleukin-6: model ofNF-kappaB- and map kinase-dependent inflammation in advancedatherosclerosis. J Biol Chem 280: 21763–21772, 2005.

106. Lilley BN and Ploegh HL. A membrane protein required fordislocation of misfolded proteins from the ER. Nature 429: 834–840, 2004.

107. Lippincott-Schwartz J, Bonifacino JS, Yuan LC, and Klausner

RD. Degradation from the endoplasmic reticulum: disposing ofnewly synthesized proteins. Cell 54: 209–220, 1988.

1146 STEFAN J. MARCINIAK AND DAVID RON

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 15: Endoplasmic Reticulum Stress Signaling in Disease

108. Liu CY, Xu Z, and Kaufman RJ. Structure and intermolecularinteractions of the luminal dimerization domain of humanIRE1alpha. J Biol Chem 278: 17680–17687, 2003.

109. Liu M, Li Y, Cavener D, and Arvan P. Proinsulin disulfide mat-uration and misfolding in the endoplasmic reticulum. J Biol Chem

280: 13209–13212, 2005.110. Lomas DA, Evans DL, Finch JT, and Carrell RW. The mecha-

nism of Z alpha 1-antitrypsin accumulation in the liver. Nature 357:605–607, 1992.

111. Lomas DA and Mahadeva R. Alpha1-antitrypsin polymerizationand the serpinopathies: pathobiology and prospects for therapy.J Clin Invest 110: 1585–1590, 2002.

112. Lu PD, Harding HP, and Ron D. Translation re-initiation atalternative open reading frames regulates gene expression in anintegrated stress response. J Cell Biol 167: 27–33, 2004.

113. Lu R and Misra V. Potential role for luman, the cellular homologueof herpes simplex virus VP16 (alpha gene trans-inducing factor), inherpesvirus latency. J Virol 74: 934–943, 2000.

114. Ma Y and Hendershot LM. Delineation of a negative feedbackregulatory loop that controls protein translation during endoplas-mic reticulum stress. J Biol Chem 278: 34864–34873, 2003.

115. Mancini R, Aebi M, and Helenius A. Multiple endoplasmic retic-ulum-associated pathways degrade mutant yeast carboxypeptidaseY in mammalian cells. J Biol Chem 278: 46895–46905, 2003.

116. Marciniak SJ, Garcia-Bonilla L, Hu J, Harding HP, and Ron D.

Activation-dependent substrate recruitment by the eukaryotictranslation initiation factor kinase PERK. J Cell Biol. In press.

117. Marciniak SJ, Yun CY, Oyadomari S, Novoa I, Zhang Y, Jun-

greis R, Nagata K, Harding HP, and Ron D. CHOP induces deathby promoting protein synthesis and oxidation in the stressed en-doplasmic reticulum. Genes Dev 18: 3066–3077, 2004.

118. McCullough KD, Martindale JL, Klotz LO, Aw TY, and Hol-

brook NJ. Gadd153 sensitizes cells to endoplasmic reticulumstress by down-regulating Bcl2 and perturbing the cellular redoxstate. Mol Cell Biol 21: 1249–1259, 2001.

119. Meyer M, Caselmann WH, Schluter V, Schreck R, Hofschnei-

der PH, and Baeuerle PA. Hepatitis B virus transactivator MHBst:activation of NF-kappa B, selective inhibition by antioxidants andintegral membrane localization. EMBO J 11: 2991–3001, 1992.

120. Micheau O, Lens S, Gaide O, Alevizopoulos K, and Tschopp J.

NF-kappaB signals induce the expression of c-FLIP. Mol Cell Biol

21: 5299–5305, 2001.121. Mikulits W, Pradet-Balade B, Habermann B, Beug H, Garcia-

Sanz JA, and Mullner EW. Isolation of translationally controlledmRNAs by differential screening. FASEB J 14: 1641–1652, 2000.

122. Molinari M, Calanca V, Galli C, Lucca P, and Paganetti P. Roleof EDEM in the release of misfolded glycoproteins from the cal-nexin cycle. Science 299: 1397–1400, 2003.

123. Mori K, Ogawa N, Kawahara T, Yanagi H, and Yura T. mRNAsplicing-mediated C-terminal replacement of transcription factorHac1p is required for efficient activation of the unfolded proteinresponse. Proc Natl Acad Sci USA 97: 4660–4665, 2000.

124. Mueller PP, Harashima S, and Hinnebusch AG. A segment ofGCN4 mRNA containing the upstream AUG codons confers trans-lational control upon a heterologous yeast transcript. Proc Natl

Acad Sci USA 84: 2863–2867, 1987.125. Mueller PP and Hinnebusch AG. Multiple upstream AUG codons

mediate translational control of GCN4. Cell 45: 201–207, 1986.126. Munro S and Pelham HR. An Hsp70-like protein in the ER:

identity with the 78 kd glucose-regulated protein and immunoglob-ulin heavy chain binding protein. Cell 46: 291–300, 1986.

127. Nagai K, Oubridge C, Kuglstatter A, Menichelli E, Isel C, and

Jovine L. Structure, function and evolution of the signal recogni-tion particle. EMBO J 22: 3479–3485, 2003.

128. Nakatani Y, Kaneto H, Kawamori D, Yoshiuchi K, Hatazaki M,

Matsuoka TA, Ozawa K, Ogawa S, Hori M, Yamasaki Y, and

Matsuhisa M. Involvement of endoplasmic reticulum stress ininsulin resistance and diabetes. J Biol Chem 280: 847–851, 2005.

129. Nishitoh H, Matsuzawa A, Tobiume K, Saegusa K, Takeda K,

Inoue K, Hori S, Ishikawa K, Mizusawa H, Kakizuka A, and

Ichijo H. ASK1 is essential for endoplasmic reticulum stress-in-duced neuronal cell death triggered by expanded polyglutaminerepeats. Genes Dev 16: 1345–1355, 2002.

130. Nishitoh H, Saitoh M, Mochida Y, Takeda K, Nakano H, Rothe

M, Miyazono K, and Ichijo H. ASK1 is essential for JNK/SAPKactivation by TRAF2. Mol Cell 2: 389–395, 1998.

131. Novoa I, Zeng H, Harding H, and Ron D. Feedback inhibition ofthe unfolded protein response by GADD34-mediated dephosphor-ylation of eIF2a. J Cell Biol 153: 1011–1022, 2001.

132. Novoa I, Zhang Y, Zeng H, Jungreis R, Harding HP, and Ron D.

Stress-induced gene expression requires programmed recoveryfrom translational repression. EMBO J 22: 1180–1187, 2003.

133. Okada T, Yoshida H, Akazawa R, Negishi M, and Mori K.

Distinct roles of ATF6 and PERK in transcription during the mam-malian unfolded protein response. Biochem J 366: 585–594, 2002.

134. Okamura K, Kimata Y, Higashio H, Tsuru A, and Kohno K.

Dissociation of Kar2p/BiP from an ER sensory molecule, Ire1p,triggers the unfolded protein response in yeast. Biochem Biophys

Res Commun 279: 445–450, 2000.135. Oyadomari S, Araki E, and Mori M. Endoplasmic reticulum

stress-mediated apoptosis in pancreatic beta-cells. Apoptosis 7:335–345, 2002.

136. Oyadomari S, Koizumi A, Takeda K, Gotoh T, Akira S, Araki

E, and Mori M. Targeted disruption of the Chop gene delaysendoplasmic reticulum stress-mediated diabetes. J Clin Invest 109:525–532, 2002.

137. Oyadomari S, Takeda K, Takiguchi M, Gotoh T, Matsumoto M,

Wada I, Akira S, Araki E, and Mori M. Nitric oxide-induced apo-ptosis in pancreatic b-cells is mediated by the endoplasmic reticulumstress pathway. Proc Natl Acad Sci USA 98: 10845–10850, 2001.

138. Ozawa K, Miyazaki M, Matsuhisa M, Takano K, Nakatani Y,

Hatazaki M, Tamatani T, Yamagata K, Miyagawa J, Kitao Y,

Hori O, Yamasaki Y, and Ogawa S. The endoplasmic reticulumchaperone improves insulin resistance in type 2 diabetes. Diabetes

54: 657–663, 2005.139. Ozcan U, Cao Q, Yilmaz E, Lee AH, Iwakoshi NN, Ozdelen E,

Tuncman G, Gorgun C, Glimcher LH, and Hotamisligil GS.

Endoplasmic reticulum stress links obesity, insulin action, and type2 diabetes. Science 306: 457–461, 2004.

140. Pahl H and Baeuerle P. The ER-overload response: activation ofNF-kB. Trends Biol Sci 22: 63–67, 1997.

141. Pahl H and Baeuerle P. A novel signal transduction pathway fromthe endoplasmic reticulum to the nucleus is mediated by the tran-scription factor NF-kB. EMBO J 14: 2580–2588, 1995.

142. Pahl HL. Activators and target genes of Rel/NF-kappaB transcrip-tion factors. Oncogene 18: 6853–6866, 1999.

143. Pahl HL and Baeuerle PA. The ER-overload response: activationof NF-kappa B. Trends Biochem Sci 22: 63–67, 1997.

144. Pahl HL and Baeuerle PA. Expression of influenza virus hemag-glutinin activates transcription factor NF-kappa B. J Virol 69:1480–1484, 1995.

145. Pahl HL, Sester M, Burgert HG, and Baeuerle PA. Activation oftranscription factor NF-kappaB by the adenovirus E3/19K proteinrequires its ER retention. J Cell Biol 132: 511–522, 1996.

146. Palombella VJ, Rando OJ, Goldberg AL, and Maniatis T. Theubiquitin-proteasome pathway is required for processing the NF-kappa B1 precursor protein and the activation of NF-kappa B. Cell

78: 773–785, 1994.147. Pereira RC, Stadmeyer L, Marciniak SJ, Ron D, and Canalis

E. C/EBP homologous protein is necessary for normal osteoblasticfunction. J Cell Biochem 97: 633–640, 2005.

148. Petkova AT, Ishii Y, Balbach JJ, Antzutkin ON, Leapman RD,

Delaglio F, and Tycko R. A structural model for Alzheimer’sbeta-amyloid fibrils based on experimental constraints from solidstate NMR. Proc Natl Acad Sci USA 99: 16742–16747, 2002.

149. Pilon M, Romisch K, Quach D, and Schekman R. Sec61p servesmultiple roles in secretory precursor binding and translocation into theendoplasmic reticulum membrane. Mol Biol Cell 9: 3455–3473, 1998.

150. Plemper RK, Bohmler S, Bordallo J, Sommer T, and Wolf DH.

Mutant analysis links the translocon and BiP to retrograde proteintransport for ER degradation. Nature 388: 891–895, 1997.

151. Plemper RK, Bordallo J, Deak PM, Taxis C, Hitt R, and Wolf

DH. Genetic interactions of Hrd3p and Der3p/Hrd1p with Sec61psuggest a retro-translocation complex mediating protein transportfor ER degradation. J Cell Sci 112: 4123–4134, 1999.

ENDOPLASMIC RETICULUM AND DISEASE 1147

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 16: Endoplasmic Reticulum Stress Signaling in Disease

152. Pollard MG, Travers KJ, and Weissman JS. Ero1p: a novel andubiquitous protein with an essential role in oxidative protein fold-ing in the endoplasmic reticulum. Mol Cell 1: 171–182, 1998.

153. Poppers J, Mulvey M, Khoo D, and Mohr I. Inhibition of PKRactivation by the proline-rich RNA binding domain of the herpessimplex virus type 1 Us11 protein. J Virol 74: 11215–11221, 2000.

154. Raggo C, Rapin N, Stirling J, Gobeil P, Smith-Windsor E,

O’Hare P, and Misra V. Luman, the cellular counterpart of herpessimplex virus VP16, is processed by regulated intramembrane pro-teolysis. Mol Cell Biol 22: 5639–5649, 2002.

155. Rayet B and Gelinas C. Aberrant rel/nfkb genes and activity inhuman cancer. Oncogene 18: 6938–6947, 1999.

156. Reimold AM, Iwakoshi NN, Manis J, Vallabhajosyula P, Szo-

molanyi-Tsuda E, Gravallese EM, Friend D, Grusby MJ, Alt F,

and Glimcher LH. Plasma cell differentiation requires the tran-scription factor XBP-1. Nature 412: 300–307, 2001.

157. Reimold AM, Ponath PD, Li YS, Hardy RR, David CS,

Strominger JL, and Glimcher LH. Transcription factor B celllineage-specific activator protein regulates the gene for humanX-box binding protein 1. J Exp Med 183: 393–401, 1996.

158. Richardson JP, Mohammad SS, and Pavitt GD. Mutations caus-ing childhood ataxia with central nervous system hypomyelinationreduce eukaryotic initiation factor 2B complex formation and ac-tivity. Mol Cell Biol 24: 2352–2363, 2004.

159. Rothenburg S, Deigendesch N, Dittmar K, Koch-Nolte F, Haag

F, Lowenhaupt K, and Rich A. A PKR-like eukaryotic initiationfactor 2alpha kinase from zebrafish contains Z-DNA binding do-mains instead of dsRNA binding domains. Proc Natl Acad Sci USA

102: 1602–1607, 2005.160. Rothwarf DM and Karin M. The NF-kappa B activation pathway:

a paradigm in information transfer from membrane to nucleus.Science’s STKE (1999), http://stke.sciencemag.org/cgi/content/full/sigtrans;1999/5/re1.

161. Ruegsegger U, Leber JH, and Walter P. Block of HAC1 mRNAtranslation by long-range base pairing is released by cytoplasmicsplicing upon induction of the unfolded protein response. Cell 107:103–114, 2001.

162. Ryu EJ, Harding HP, Angelastro JM, Vitolo OV, Ron D, and

Greene LA. Endoplasmic reticulum stress and the unfolded pro-tein response in cellular models of Parkinson’s disease. J Neurosci

22: 10690–10698, 2002.163. Saliba RS, Munro PM, Luthert PJ, and Cheetham ME. The

cellular fate of mutant rhodopsin: quality control, degradation andaggresome formation. J Cell Sci 115: 2907–2918, 2002.

164. Scheuner D, Mierde DV, Song B, Flamez D, Creemers JW,

Tsukamoto K, Ribick M, Schuit FC, and Kaufman RJ. Control ofmRNA translation preserves endoplasmic reticulum function in betacells and maintains glucose homeostasis. Nat Med 11: 757–764, 2005.

165. Scheuner D, Song B, McEwen E, Gillespie P, Saunders T,

Bonner-Weir S, and Kaufman RJ. Translational control is re-quired for the unfolded protein response and in-vivo glucose ho-meostasis. Mol Cell 7: 1165–1176, 2001.

166. Schiene C and Fischer G. Enzymes that catalyse the restructuringof proteins. Curr Opin Struct Biol 10: 40–45, 2000.

167. Shamu CE and Walter P. Oligomerization and phosphorylation ofthe Ire1p kinase during intracellular signaling from the endoplas-mic reticulum to the nucleus. EMBO J 15: 3028–3039, 1996.

168. Shen J, Chen X, Hendershot L, and Prywes R. ER stress regu-lation of ATF6 localization by dissociation of BiP/GRP78 and un-masking of Golgi localization signals. Dev Cell 3: 99–111, 2002.

169. Shen X, Ellis RE, Lee K, Liu CY, Yang K, Solomon A, Yoshida

H, Morimoto R, Kurnit DM, Mori K, and Kaufman RJ. Com-plementary signaling pathways regulate the unfolded protein re-sponse and are required for C. elegans development. Cell 107:893–903, 2001.

170. Shi Y, An J, Liang J, Hayes SE, Sandusky GE, Stramm LE, and

Yang NN. Characterization of a mutant pancreatic eIF-2alpha ki-nase, PEK, and co-localization with somatostatin in islet delta cells.J Biol Chem 274: 5723–5730, 1999.

171. Shi Y, Vattem KM, Sood R, An J, Liang J, Stramm L, and Wek

RC. Identification and characterization of pancreatic eukaryoticinitiation factor 2 alpha-subunit kinase, PEK, involved in transla-tional control. Mol Cell Biol 18: 7499–7509, 1998.

172. Sidrauski C, Cox JS, and Walter P. tRNA ligase is required forregulated mRNA splicing in the unfolded protein response. Cell 87:405–413, 1996.

173. Sidrauski C and Walter P. The transmembrane kinase Ire1p is asite-specific endonuclease that initiates mRNA splicing in the un-folded protein response. Cell 90: 1031–1039, 1997.

174. Silva RM, Ries V, Oo TF, Yarygina O, Jackson-Lewis V, Ryu EJ,

Lu PD, Marciniak SM, Ron D, Przedborski S, Kholodilov N,

Greene LA, and Burke RE. CHOP/GADD153 is a mediator of apo-ptotic death in substantia nigra dopamine neurons in an in vivoneurotoxin model of parkinsonism. J Neurochem 95: 974–986, 2005.

175. Sood R, Porter AC, Ma K, Quilliam LA, and Wek RC. Pancre-atic eukaryotic initiation factor-2alpha kinase (PEK) homologuesin humans, Drosophila melanogaster and Caenorhabditis elegans

that mediate translational control in response to endoplasmic re-ticulum stress. Biochem J 346: 281–293, 2000.

176. Southwood CM, Garbern J, Jiang W, and Gow A. The unfoldedprotein response modulates disease severity in Pelizaeus-Mezbacher Disease. Neuron 36: 585–596, 2002.

177. Sriburi R, Jackowski S, Mori K, and Brewer JW. XBP1: a linkbetween the unfolded protein response, lipid biosynthesis and biogen-esis of the endoplasmic reticulum. J Cell Biol 167: 35–41, 2004.

178. Stirling J and O’Hare P. CREB4, a transmembrane bZip tran-scription factor and potential new substrate for regulation andcleavage by S1P. Mol Biol Cell 17: 413–426, 2006.

179. Strom TM, Hortnagel K, Hofmann S, Gekeler F, Scharfe C,

Rabl W, Gerbitz KD, and Meitinger T. Diabetes insipidus, dia-betes mellitus, optic atrophy and deafness (DIDMOAD) caused bymutations in a novel gene (wolframin) coding for a predictedtransmembrane protein. Hum Mol Genet 7: 2021–2028, 1998.

180. Sun XJ, Rothenberg P, Kahn CR, Backer JM, Araki E, Wilden

PA, Cahill DA, Goldstein BJ, and White MF. Structure of theinsulin receptor substrate IRS-1 defines a unique signal transduc-tion protein. Nature 352: 73–77, 1991.

181. Takeda K, Inoue H, Tanizawa Y, Matsuzaki Y, Oba J, Wa-

tanabe Y, Shinoda K, and Oka Y. WFS1 (Wolfram syndrome 1)gene product: predominant subcellular localization to endoplasmicreticulum in cultured cells and neuronal expression in rat brain.Hum Mol Genet 10: 477–484, 2001.

182. Tamemoto H, Kadowaki T, Tobe K, Yagi T, Sakura H, Haya-

kawa T, Terauchi Y, Kohjiro Ueki, Kaburagi Y, Satoh S, Seki-

hara H, Yoshioka S, Horikoshi H, Furuta Y, Ikawa Y, Kasuga

M, Yazaki Y, and Aizawa S. Insulin resistance and growth retar-dation in mice lacking insulin receptor substrate-1. Nature 372:182–186, 1994.

183. Tanaka M, Fuentes ME, Yamaguchi K, Durnin MH, Dalrymple

SA, Hardy KL, and Goeddel DV. Embryonic lethality, liver de-generation, and impaired NF-kappa B activation in IKK-beta-defi-cient mice. Immunity 10: 421–429, 1999.

184. Tirasophon W, Lee K, Callaghan B, Welihinda A, and Kaufman

RJ. The endoribonuclease activity of mammalian IRE1 autoregu-lates its mRNA and is required for the unfolded protein response.Genes Dev 14: 2725–2736, 2000.

185. Tirasophon W, Welihinda AA, and Kaufman RJ. A stress re-sponse pathway from the endoplasmic reticulum to the nucleusrequires a novel bifunctional protein kinase/endoribonuclease(Ire1p) in mammalian cells. Genes Dev 12: 1812–1824, 1998.

186. Tokunaga F, Brostrom C, Koide T, and Arvan P. Endoplasmicreticulum (ER)-associated degradation of misfolded N-linked gly-coproteins is suppressed upon inhibition of ER mannosidase I.J Biol Chem 275: 40757–40764, 2000.

187. Tu BP, Ho-Schleyer SC, Travers KJ, and Weissman JS. Bio-chemical basis of oxidative protein folding in the endoplasmicreticulum. Science 290: 1571–1574, 2000.

188. Tu BP and Weissman JS. The FAD and O2-dependent reactioncycle of Ero1-mediated oxidative protein folding in the endoplas-mic reticulum. Mol Cell 10: 983–994., 2002.

189. Tu BP and Weissman JS. Oxidative protein folding in eukaryotes:mechanisms and consequences. J Cell Biol 164: 341–346, 2004.

190. Tzamarias D, Roussou I, and Thireos G. Coupling of GCN4mRNA translational activation with decreased rates of polypeptidechain initiation. Cell 57: 947–954, 1989.

1148 STEFAN J. MARCINIAK AND DAVID RON

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 17: Endoplasmic Reticulum Stress Signaling in Disease

191. Tzamarias D and Thireos G. Evidence that the GCN2 proteinkinase regulates reinitiation by yeast ribosomes. EMBO J 7: 3547–3551, 1988.

192. Urano F, Wang X, Bertolotti A, Zhang Y, Chung P, Harding H,

and Ron D. Coupling of stress in the endoplasmic reticulum toactivation of JNK protein kinases by transmembrane protein kinaseIRE1. Science 287: 664–666, 2000.

193. Van der Voorn JP, van Kollenburg B, Bertrand G, Van Haren

K, Scheper GC, Powers JM, and van der Knaap MS. Theunfolded protein response in vanishing white matter disease.J Neuropathol Exp Neurol 64: 770–775, 2005.

194. Vattem KM and Wek RC. Reinitiation involving upstream ORFsregulates ATF4 mRNA translation in mammalian cells. Proc Natl

Acad Sci USA 101: 11269–11274, 2004.195. Wang CY, Mayo MW, and Baldwin AS Jr. TNF- and cancer

therapy-induced apoptosis: potentiation by inhibition of NF-kap-paB. Science 274: 784–787, 1996.

196. Wang J, Takeuchi T, Tanaka S, Kubo SK, Kayo T, Lu D,

Takata K, Koizumi A, and Izumi T. A mutation in the insulin 2gene induces diabetes with severe pancreatic beta-cell dysfunctionin the Mody mouse. J Clin Invest 103: 27–37, 1999.

197. Wang XZ, Lawson B, Brewer J, Zinszner H, Sanjay A, Mi L,

Boorstein R, Kreibich G, Hendershot L, and Ron D. Signalsfrom the stressed endoplasmic reticulum induce C/EBP homolo-gous protein (CHOP/GADD153). Mol Cell Biol 16: 4273–4280, 1996.

198. Wang XZ, Harding HP, Zhang Y, Jolicoeur EM, Kuroda M, and

Ron D. Cloning of mammalian Ire1 reveals diversity in the ERstress responses. EMBO J 17: 5708–5717, 1998.

199. Wang Y, Shen J, Arenzana N, Tirasophon W, Kaufman RJ, and

Prywes R. Activation of ATF6 and an ATF6 DNA binding site bythe ER stress response. J Biol Chem 275: 27013–27020, 2000.

200. Wei J and Hendershot LM. Characterization of the nucleotidebinding properties and ATPase activity of recombinant hamsterBiP purified from bacteria. J Biol Chem 270: 26670–26676, 1995.

201. Welch GN and Loscalzo J. Homocysteine and atherothrombosis.N Engl J Med 338: 1042–1050, 1998.

202. Welihinda AA and Kaufman RJ. The unfolded protein responsepathway in Saccharomyces cerevisiae. Oligomerization and trans-phosphorylation of Ire1p (Ern1p) are required for kinase activa-tion. J Biol Chem 271: 18181–18187, 1996.

203. Welsh M, Scherberg N, Gilmore R, and Steiner DF. Transla-tional control of insulin biosynthesis. Evidence for regulation ofelongation, initiation and signal-recognition-particle-mediatedtranslational arrest by glucose. Biochem J 235: 459–467, 1986.

204. Wen XY, Stewart AK, Sooknanan RR, Henderson G, Hawley

TS, Reimold AM, Glimcher LH, Baumann H, Malek LT, and

Hawley RG. Identification of c-myc promoter-binding protein andX-box binding protein 1 as interleukin-6 target genes in humanmultiple myeloma cells. Int J Oncol 15: 173–178, 1999.

205. Werstuck GH, Lentz SR, Dayal S, Hossain GS, Sood SK, Shi

YY, Zhou J, Maeda N, Krisans SK, Malinow MR, and Austin

RC. Homocysteine-induced endoplasmic reticulum stress causesdysregulation of the cholesterol and triglyceride biosynthetic path-ways. J Clin Invest 107: 1263–1273, 2001.

206. Wicksteed B, Herbert TP, Alarcon C, Lingohr MK, Moss LG,

and Rhodes CJ. Cooperativity between the preproinsulin mRNAuntranslated regions is necessary for glucose-stimulated transla-tion. J Biol Chem 276: 22553–22558, 2001.

207. Wille H, Michelitsch MD, Guenebaut V, Supattapone S, Ser-

ban A, Cohen FE, Agard DA, and Prusiner SB. Structuralstudies of the scrapie prion protein by electron crystallography.Proc Natl Acad Sci USA 99: 3563–3568, 2002.

208. Williams NP, Hinnebusch AG, and Donahue TF. Mutations inthe structural genes for eukaryotic initiation factors 2 alpha and 2beta of Saccharomyces cerevisiae disrupt translational control ofGCN4 mRNA. Proc Natl Acad Sci USA 86: 7515–7519, 1989.

209. Wolcott C and Rallison M. Infancy-onset diabetes mellitus andmultiple epiphyseal dysplasia. J Pediatr 80: 292–297, 1972.

210. Wu M, Lee H, Bellas RE, Schauer SL, Arsura M, Katz D,

FitzGerald MJ, Rothstein TL, Sherr DH, and Sonenshein GE.

Inhibition of NF-kappaB/Rel induces apoptosis of murine B cells.EMBO J 15: 4682–4690, 1996.

211. Wu S, Tan M, Hu Y, Wang JL, Scheuner D, and Kaufman RJ.

Ultraviolet light activates NFkB through translational inhibition ofIkBa synthesis. J Biol Chem 279: 34898–34902, 2004.

212. Xie J, Zhu H, Larade K, Ladoux A, Seguritan A, Chu M, Ito S,

Bronson RT, Leiter EH, Zhang CY, Rosen ED, and Bunn HF.

Absence of a reductase, NCB5OR, causes insulin-deficient diabetes.Proc Natl Acad Sci USA 101: 10750–10755, 2004.

213. Yabe D, Brown MS, and Goldstein JL. Insig-2, a second endo-plasmic reticulum protein that binds SCAP and blocks export ofsterol regulatory element-binding proteins. Proc Natl Acad Sci USA

99: 12753–12758, 2002.214. Yamamoto K, Yoshida H, Kokame K, Kaufman RJ, and Mori K.

Differential contributions of ATF6 and XBP1 to the activation ofendoplasmic reticulum stress-responsive cis-acting elementsERSE, UPRE and ERSE-II. J Biochem (Tokyo) 136: 343–350, 2004.

215. Yamamoto M, Yamazaki S, Uematsu S, Sato S, Hemmi H,

Hoshino K, Kaisho T, Kuwata H, Takeuchi O, Takeshige K,

Saitoh T, Yamaoka S, Yamamoto N, Yamamoto S, Muta T,

Takeda K, and Akira S. Regulation of Toll/IL-1-receptor-mediatedgene expression by the inducible nuclear protein IkappaBzeta.Nature 430: 218–222, 2004.

216. Yang T, Espenshade PJ, Wright ME, Yabe D, Gong Y, Aeber-

sold R, Goldstein JL, and Brown MS. Crucial step in cholesterolhomeostasis: sterols promote binding of SCAP to INSIG-1, a mem-brane protein that facilitates retention of SREBPs in ER. Cell 110:489–500, 2002.

217. Yang YL, Reis LF, Pavlovic J, Aguzzi A, Schafer R, Kumar A,

Williams BR, Aguet M, and Weissmann C. Deficient signaling inmice devoid of double-stranded RNA-dependent protein kinase.EMBO J 14: 6095–6106, 1995.

218. Ye J, Rawson RB, Komuro R, Chen X, Dave UP, Prywes R,

Brown MS, and Goldstein JL. ER stress induces cleavage ofmembrane-bound ATF6 by the same proteases that processSREBPs. Mol Cell 6: 1355–1364, 2000.

219. Ye Y, Shibata y Yun C, Ron D, and Rapoport TA. A membraneprotein complex mediates retro-translocation from the ER lumeninto the cytosol. Nature 429: 841–847, 2004.

220. Yoshida H, Haze K, Yanagi H, Yura T, and Mori K. Identificationof the cis-acting endoplasmic reticulum stress response elementresponsible for transcriptional induction of mammalian glucose-regulated proteins. Involvement of basic leucine zipper transcrip-tion factors. J Biol Chem 273: 33741–33749, 1998.

221. Yoshida H, Matsui T, Hosokawa N, Kaufman RJ, Nagata K,

and Mori K. A time-dependent phase shift in the mammalianunfolded protein response. Dev Cell 4: 265–271, 2003.

222. Yoshida H, Matsui T, Yamamoto A, Okada T, and Mori K. XBP1mRNA is induced by ATF6 and spliced by IRE1 in response to ERstress to produce a highly active transcription factor. Cell 107:881–891, 2001.

223. Yoshioka M, Kayo T, Ikeda T, and Koizumi A. 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, 1997.224. Zeng L, Lu M, Mori K, Luo S, Lee AS, Zhu Y, and Shyy JY. ATF6

modulates SREBP2-mediated lipogenesis. EMBO J 23: 950–958, 2004.225. Zhang K, Shen X, Wu J, Sakaki K, Saunders T, Rutkowski DT,

Back SH, and Kaufman RJ. Endoplasmic reticulum stress acti-vates cleavage of CREBH to induce a systemic inflammatory re-sponse. Cell 124: 587–599, 2006.

226. Zhang P, McGrath BC, Reinert J, Olsen DS, Lei L, Gill S, Wek

SA, Vattem KM, Wek RC, Kimball SR, Jefferson LS, and

Cavener DR. The GCN2 eIF2alpha kinase is required for adaptationto amino acid deprivation in mice. Mol Cell Biol 22: 6681–6688, 2002.

227. Zhang Y, McLaughlin R, Goodyer C, and LeBlanc A. Selectivecytotoxicity of intracellular amyloid beta peptide1–42 through p53and Bax in cultured primary human neurons. J Cell Biol 156:519–529, 2002.

228. Zhou M and Schekman R. The engagement of Sec61p in the ERdislocation process. Mol Cell 4: 925–934, 1999.

229. Zinszner H, Kuroda M, Wang X, Batchvarova N, Lightfoot RT,

Remotti H, Stevens JL, and Ron D. CHOP is implicated inprogrammed cell death in response to impaired function of theendoplasmic reticulum. Genes Dev 12: 982–995, 1998.

ENDOPLASMIC RETICULUM AND DISEASE 1149

Physiol Rev • VOL 86 • OCTOBER 2006 • www.prv.org

on February 11, 2015

Dow

nloaded from

Page 18: Endoplasmic Reticulum Stress Signaling in Disease

doi:10.1152/physrev.00015.2006 86:1133-1149, 2006.Physiol RevStefan J. Marciniak and David RonEndoplasmic Reticulum Stress Signaling in Disease

You might find this additional info useful...

222 articles, 122 of which can be accessed free at:This article cites /content/86/4/1133.full.html#ref-list-1

100 other HighWire hosted articles, the first 5 are:This article has been cited by

  [PDF] [Full Text] [Abstract]

, May 16, 2014; 289 (20): 13758-13768.J. Biol. Chem.Ick Young KimJea Hwang Lee, Joon Hyun Kwon, Yeong Ha Jeon, Kwan Young Ko, Seung-Rock Lee andp97(VCP) during Endoplasmic Reticulum-associated Degradation

of Selenoprotein S Are Essential Residues for Interaction with183 and Pro178Pro 

[PDF] [Full Text] 2014; 34 (7): 1329-1332.Arterioscler Thromb Vasc Biol

Renu Virmani, Michael Joner and Kenichi Sakakura: CalcificationATVBRecent Highlights of

  [PDF] [Full Text] [Abstract]

, July , 2014; 86 (1): 20-27.Mol PharmacolJohn R. Bracamontes, Ping Li, Gustav Akk and Joe Henry SteinbachHave Opposite Effects on Surface Expression

Subunitsδ and 4α Receptor AMutations in the Main Cytoplasmic Loop of the GABA 

[PDF] [Full Text] [Abstract], February , 2015; 35 (2): 645-651.Anticancer Res

CHRISTOS KAZAZISNATALIA G. VALLIANOU, ANGELOS EVANGELOPOULOS, NIKOS SCHIZAS andPotential Anticancer Properties and Mechanisms of Action of Curcumin 

[PDF] [Full Text] [Abstract], February , 2015; 45 (2): 365-376.Eur Respir J

Marciniak and Pieter S. HiemstraEmily F.A. van't Wout, Annemarie van Schadewijk, David A. Lomas, Jan Stolk, Stefan J.

-antitrypsin deficiency1αFunction of monocytes and monocyte-derived macrophages in

including high resolution figures, can be found at:Updated information and services /content/86/4/1133.full.html

can be found at:Physiological Reviewsabout Additional material and information http://www.the-aps.org/publications/prv

This information is current as of February 11, 2015. 

website at http://www.the-aps.org/.MD 20814-3991. Copyright © 2006 by the American Physiological Society. ISSN: 0031-9333, ESSN: 1522-1210. Visit ourpublished quarterly in January, April, July, and October by the American Physiological Society, 9650 Rockville Pike, Bethesda

provides state of the art coverage of timely issues in the physiological and biomedical sciences. It isPhysiological Reviews

on February 11, 2015

Dow

nloaded from