0163-769X/05/$20.00/0 Printed in U.S.A. Endocrine Reviews 26(4):479 –503 Copyright © 2005 by The Endocrine Society doi: 10.1210/er.2004-0010 Beyond the Signal Sequence: Protein Routing in Health and Disease Cecilia Castro-Fernández,* Guadalupe Maya-Núñez,* and P. Michael Conn Oregon National Primate Research Center (P.M.C., C.C-F., G.M-N.) and Departments of Physiology and Pharmacology and Cell and Developmental Biology (P.M.C.), Oregon Health and Science University, Beaverton, Oregon 97006; and Research Units in Developmental Biology (C.C-F., G.M-N.), and Reproductive Medicine (P.M.C.), Instituto Mexicano del Seguro Social, Mexico Distrito Federal 06725, Mexico Receptors, hormones, enzymes, ion channels, and structural components of the cell are created by the act of protein synthesis. Synthesis alone is insufficient for proper function, of course; for a cell to operate effectively, its components must be correctly compartmentalized. The mechanism by which proteins maintain the fidelity of localization warrants attention in light of the large number of different molecules that must be routed to distinct subcellular loci, the potential for error, and resultant disease. This review summarizes diseases known to have etiologies based on defective protein folding or failure of the cell’s quality control apparatus and presents approaches for therapeutic intervention. (Endocrine Reviews 26: 479 –503, 2005) I. Introduction II. Protein Processing and the Role of Chaperones A. ER processing and quality control (QC) B. Golgi apparatus C. Mitochondria D. Chaperones: the role of folding in QC and routing III. Diseases Caused by Defective Routing A. G Protein-Coupled Receptors (GPCRs) in disease states B. Wild-type protein degradation and oligomerization in protein processing C. Other proteins IV. Diseases Caused by Conformational Errors A. AD and neuropathies B. Parkinson’s disease (PD) C. Huntington’s disease (HD) D. Viral prions E. Cataracts F. ␣1-AT deficiency V. Rescue of Defective Proteins A. Temperature B. Chemical chaperones C. Nonspecific agents VI. Conclusions I. Introduction I N 1971 GÜNTER BLOBEL presented a first version of the “signal hypothesis.” He postulated that, in the case of secretory proteins, there is an intrinsic signal that governs protein movement to and across membranes. Blobel’s laboratory and others subsequently showed that other distinct “address tags” direct proteins to intracellular organelles. These principles are universal, operating similarly in yeast, plant, and animal cells; errors in these tags can result in hereditable disease. In 1999, the Nobel Prize recognized Blobel’s work. Recent evidence, coming largely from mutations that result in human disease, led to another inescapable conclusion: the successful intracellular routing of many proteins is also governed by a sensitive quality control (QC) system that recognizes particular structural motifs, then retains and degrades defective molecules (1). Abnormal proteins are extremely dangerous to cells because they interfere with normal functions and eventually may result in cell death. The cell limits the accumulation of aberrant conformations with a chaperone system to assist the overall folding process (2, 3). A picture is now emerging that, for some wild-type (wt) proteins, such as the opiate receptor and human (but not rodent) GnRH receptor (GnRHR), only a modest percentage of the newly synthesized protein is routed correctly; the balance is destroyed. The reason for this apparent inefficiency is unclear but suggests the possibility of an unappre- First Published Online November 9, 2004 * C.C-F. and G.M.-N. contributed equally to the preparation of this manuscript and are designated cofirst authors. Abbreviations: AD, Alzheimer’s disease; ␣1-AT, ␣1-antitrypsin; APP, amyloid precursor protein; AQP2, aquaporin-2; BiP, Ig-binding protein; BSE, bovine spongiform encephalopathy; CF, cystic fibrosis; CFTR, CF transmembrane conductance regulator; CJD, Creutzfeldt-Jackob disease; COPII, coat protein complex II; DA, dopamine; ER, endoplasmic reticulum; ERAD, ER-associated degradation; ERGIC, ER-Golgi intermediate compartment; GABABR, ␥-aminobutyric acid B receptor; GnRHR, GnRH receptor; GPCR, G protein-coupled receptor; GRP, glucose-regulated protein; HC, hemochromatosis; HD, Huntington’s disease; HFE, hemochromatosis gene; HH, hypogonadotrophic hypogonadism; Hsp, heat shock protein; LB, Lewy body; LDL, low-density lipoprotein; LDLR, LDL receptor; 2M, 2-microglobulin; MHC, major histocompatibility complex; NCCT, NaCl cotransporter; NDI, nephrogenic diabetes insipidus; PD, Parkinson’s disease; PDI, protein disulfide isomerase; PS1, presenilin 1; QC, quality control; RAP, receptor-associated protein; rBAT, rat liver bile acid coenzyme A: amino acid Nacetyltransferase; Rh, rhodopsin; RP, retinitis pigmentosa; TSE, transmissible spongiform encephalopathy; UGT, uridine diphosphateglucose:glycoprotein glycosyltransferase; UPR, unfolded protein response; V2R, vasopressin type 2 receptor; vCJD, variant CJD. Endocrine Reviews is published bimonthly by The Endocrine Society (http://www.endo-society.org), the foremost professional society serving the endocrine community. 479 480 Endocrine Reviews, June 2005, 26(4):479 –503 ciated level of regulation that may provide receptors for up-regulation, enzymes for feedback adjustment, or ion channels to adapt to rapidly changing conditions. In some cases, protein oligomerization is requisite for export of the nascent protein from the endoplasmic reticulum (ER); mutant or defective proteins may oligomerize with and retain wt proteins, leading to a dominant-negative phenotype. Diseases caused by abnormal proteins include circumstances in which a specific protein, or protein complex, fails to fold correctly [e.g., cystic fibrosis (CF)], or is not sufficiently stable to perform its normal function (e.g., some forms of cancer). They also include conditions in which abnormal folding results in the failure of protein to be correctly positioned within the cell [e.g., hypogonadotrophic hypogonadism (HH), ␣1-antitrypsin (␣1-AT) deficiency, and some forms of retinitis pigmentosa (RP)]. Other misfolding diseases, amyloidoses, are caused by conformational changes coupled to the aggregation and cytotoxicity of misfolded proteins outside the cell (Table 1). This review addresses the molecular mechanisms of protein misfolding and protein aggregation and the newly appreciated role of this event in health and disease. II. Protein Processing and the Role of Chaperones A. ER processing and quality control (QC) The ER is a lamellar organelle associated with ribosomal protein synthesis, co- and posttranslational modification, and protein folding (Fig. 1). Proteins enter the ER in an unfolded state via an aqueous channel, the Sec61 translocon, to receive posttranslational modifications such as disulfide bond formation, signal peptide cleavage, N-linked glycosylation, glycophosphatidyl inositol anchor addition, and palmitoylation/myristoylation (3, 4). If the protein is not properly folded/matured, it will remain in the ER and is eventually degraded without reaching its normal cellular site of action (5, 6). TABLE 1. Examples of diseases caused by abnormal proteins Disease Protein(s) Caused by defective routing HH GnRHR RP Rhodopsin NDI Arginine vasopressin AQP2 Cancer p53 Cystinuria rBAT CF CFTR Gitelman’s syndrome Thiazide-sensitive NaCl transporter HC HFE Hypercholesterolemia LDLR Caused by conformational errors AD Amyloid- peptide -Protein PD ␣-Synuclein Parkin Ubiquitin-carboxyl-hydrolase L1 HD Huntingtin TSEs Prion protein Cataracts ␣B-crystallin ␣1-AT deficiency ␣1-ATZ Castro-Fernández et al. • Protein Routing in Health and Disease TABLE 2. Chaperones involved in QC Family Hsp70 Hsp90 Hsp40 Hsp60 Lectins Oxidoreductases Members BiP, Grp78 Grp94 ERdj1-ERdj5, Sec63 Hsp60 Calnexin, calreticulin PDI, ERp57, ERp72, etc. Hsp100 Hsp100 Small Hsp Hsp27, ␣-crystallin Function Chaperone Chaperone Cofactor, can regulate BiP in vitro Mitochondrial chaperone Glycoprotein QC Formation of disulfide bonds and isomerization and reduction Mitochondrial chaperone Holding denatured proteins The ER has QC machinery that discriminates between properly folded proteins and misfolded ones; sometimes aberrant forms acquire their correct conformation after several passages through the folding chaperone system. The ER functions as a guardian that prevents the premature exit of folding intermediates or incompletely assembled proteins from the ER. This QC mechanism includes a wide range of molecular chaperones and folding catalysts and appears to have evolved to assist the folding process (7, 8) (Table 2). All proteins are subjected to a “primary” QC that monitors their architectural design through ubiquitous folding sensors that are abundant in the ER and recognize common structural and biophysical features which distinguish native from nonnative protein conformations (7, 8). The Ig-binding protein (BiP, discussed below), calnexin, calreticulin, glucoseregulated protein (GRP)-94, and the thiol-disulfide oxidoreductases protein disulfide isomerase (PDI) and Erp57 are molecular chaperones and sensors found in the ER (9 – 11). A misfolded or incompletely folded protein can be bound by these factors, and these provide assisted folding. Some proteins utilize mostly one system; others use them sequentially or even simultaneously, and the chaperone systems can also serve as retention anchors for immature proteins (7, 12) or protein switches that either properly route a protein or direct it to a degradation pathway (13, 14). The secondary QC, which is often cell type specific, contains selective mechanisms that regulate the export of proteins, which are accompanied out of the ER by specialized chaperones and enzymes (12). The effectiveness of the QC system depends, in large part, on the ability of functional proteins to avoid capture because their hydrophobic regions are generally buried. However, when these hydrophobic surfaces are exposed, ER resident chaperones interact with them, retaining them in the ER (8). It appears that the glucosyl transferases recognize misfolded proteins when hydrophobic residues situated on the carboxyl side of the glycosylation site are present. The QC mechanism, which involves a remarkable series of glycosylation and deglycosylation reactions, enables correctly folded proteins to be distinguished from misfolded ones (15) (Fig. 2). However, if the glycoprotein is misfolded it results in retention in the ER where it can undergo a new cycle of reglycosylation by a uridine diphosphate-glucose:glycoprotein Castro-Fernández et al. • Protein Routing in Health and Disease Endocrine Reviews, June 2005, 26(4):479 –503 481 FIG. 1. The QC process in the secretory pathway. The ER is a lamellar organelle associated with ribosomal protein synthesis, co- and posttranslational modification, and protein folding. Proteins enter the ER in an unfolded state via the Sec61 translocon to receive posttranslational modifications; if the protein is not properly folded/matured, it will remain in the ER and is eventually degraded without reaching its normal cellular site of action. The discrimination between properly folded proteins and misfolded ones takes place in the ER by the QC machinery; sometimes aberrant forms acquire their correct conformation after several passages through the folding chaperone system. Terminally misfolded proteins are retrotranslocated across the ER membrane to be degraded by proteasomes in a process known as ER-associated degradation. Correctly folded proteins are allowed to exit the ER and enter the secretory pathway, through exit sites located at the ER membrane. Proteins are packed into COPII-coated vesicles and leave the ER to be transported to the Golgi apparatus, where they fuse with the target membrane elements in the ERGIC, near the Golgi. Once the proteins reach the Golgi complex, they are sorted to peripheral compartments such as vacuoles, plasma membrane, and secretory vesicles. Retrieval of misfolded proteins from the Golgi complex to the ER occurs; also defective proteins can be transported to the endosomal system for lysosomal or vacuolar degradation. SV, Secretory vesicle. glycosyltransferase (UGT) (3, 16). Another protein, ER degradation-enhancing 1,2-mannosidase I (a recently discovered lectin), binds misfolded proteins that are released from calnexin and promotes their degradation, rather than allowing entry into another UGT-calnexin cycle (13). It was initially believed that ER-resident proteinases and peptidases degraded misfolded and orphan proteins (17). However, terminally misfolded molecules are ”retrotranslocated“ or ”dislocated“ across the ER membrane to be degraded by cytosolic proteasomes (Fig. 3). The misfolded or unassembled proteins are retained; proteins that normally form associations with other proteins cannot progress through the secretory pathway until they are bound to their partners. Instead, the misfolded proteins are recognized by ER chaperones or by other factors; they are then again translocated to the cytosol through the translocon Sec61 (18). This process, known as ER-associated degradation (ERAD), selects the QC substrates to be degraded, targets them to the translocon, and transports them back to the cytosol, where the proteolytic components of the protein degradation pathway such as ubiquitin, ubiquitin-conjugating enzymes, and the cytosolic 26S proteasome are used (6). After retrotranslocation from the ER, misfolded proteins are polyubiquitylated before degradation. Lack of polyubiquitylation prevents proteasomal degradation and leads to failure in transport of the protein from the ER lumen or membrane out to the cytoplasm, causing the protein to remain in the ER. Ubiquitination may facilitate the recruitment of factors that actively extract the polypeptide from the ER membrane. The proteins that undergo proteasomal degradation by the proteolytic complex 26S proteasome are marked by the binding of a ubiquitin chain to a lysine residue (19 –21). The 26S proteasome is composed of a core 20S particle. These proteins are digested into short peptides and one or two 19S regulatory particles, which are responsible for substrate recognition and transport into the core particle. The entry channel is quite narrow, and even in its open state, it allows entry of only unfolded proteins (22, 23). Although excluding normal globular proteins, this architecture requires complex ATPdependent mechanisms to recognize, unfold, and linearize substrates and to inject them into the 20S proteasome (23). After proteolysis, the polyubiquitin chain is released and hydrolyzed into single ubiquitin moieties that go through a new cycle of protein degradation (3). When aberrant proteins accumulate in the ER due to ineffective ER degradation machinery, there is a different response that occurs. This “unfolded protein response” (UPR) leads to synthesis of ER-resident chaperones and enzymes. 482 Endocrine Reviews, June 2005, 26(4):479 –503 Castro-Fernández et al. • Protein Routing in Health and Disease FIG. 2. The calnexin/calreticulin cycle. The QC mechanism involves a series of glycosylation and deglycosylation reactions that enable correctly folded proteins to be distinguished from misfolded ones. It appears that the glucosyl transferases recognize misfolded proteins when hydrophobic residues situated on the carboxyl side of the glycosylation site are present. When proteins translocate the ER membrane, a core of oligosaccharides (Glc3Man9-GlcNAc2) is attached to asparagine residues in the Asn-X-Ser/Thr consensus sequence (step 1). After the folding process, two of the outermost glucose residues are eliminated by the glucosidases I and II (step 2). Disulfide bonds are transiently formed by the Erp57 (a thiol oxidoreductase chaperone), which interacts specifically with calnexin and calreticulin, and the monoglucosylated protein (step 3). After the last glucose residue is eliminated by glucosidase II, the complex dissociates and releases a protein with a carbohydrate structure of Man9GlcNAc2 (step 4). Correctly folded proteins are then transported out of the ER to the secretory pathway. When the glycoprotein is not correctly folded, the N-glycan is reglycosylated by the UGT (steps 5 and 6), which induces a new round of calnexin/calreticulin binding preventing the exit of the protein from the ER. Drugs that block N-linked glycosylation or disulfide bond formation or alter calcium homeostasis produce this response. If the response fails to clear the ER, apoptosis is induced (8). The UPR and ERAD are partially overlapping responses that lead to elimination of misfolded secretory proteins. Induction of the UPR up-regulates multiple genes involved in ERAD/proteasomal protein degradation. Defects in ERAD constitutively activate the UPR, and simultaneous deficiency of both pathways decreases cell viability. In contrast, correctly folded proteins are transported out of the ER and do not go through this QC system, which only retains and degrades misfolded proteins (7, 24). Defective components can cause the QC system to lead to destruction of recoverable proteins at a great cost to energy input. A disturbed QC leads to disease and eventually to cell death. Proteins are allowed to exit the ER and enter the secretory pathway when they are properly folded, through exit sites located at the ER membrane (25–28). Misfolded proteins and ER chaperones do not usually exit the ER; however, it has been observed that some misfolded proteins that are exported from the ER shuttle between the ER and the Golgi apparatus during their maturation. The UPR also induces the expression of genes required for ER export (28, 29). Proper function of the QC machinery in the ER is essential for a healthy cellular metabolism. B. Golgi apparatus Coated vesicles mediate intracellular transport of proteins. Correctly folded proteins are packed into coat protein com- plex II (COPII)-coated vesicles and leave the ER at the exit sites and are transported to the Golgi apparatus, where they fuse with the target membrane elements in the ER-Golgi intermediate compartment (ERGIC), near the Golgi (25–28). ER exit can be by bulk flow or it can be facilitated by transport receptors, such as ERGIC-53, a transmembrane lectin that cycles between ER, ERGIC, and Golgi. ERGIC-53 possesses an ER exit determinant in its cytosolic domain that mediates binding to COPII-coated vesicles. Once in the Golgi, the proteins are sorted to peripheral compartments of the celllike vacuoles, plasma membrane, and secretory vesicles (28, 30). Cells also have a back-up QC system, which prevents transport of misfolded proteins beyond the ER. The Golgi complex also makes conformation-based discrimination of abnormal proteins, which are targeted for degradation. In certain cases, retrieval of misfolded proteins from the Golgi complex to the ER occurs (31). The recycling of misfolding proteins might be required by ERAD, assisting the ER QC machinery (7, 28, 32–34). It appears that misfolded soluble proteins require a modification obtained in the Golgi, presumably in the cis-Golgi, to be recognized as ERAD substrates. It is possible that retrotranslocation occurs from a specialized compartment that soluble proteins can reach only via the Golgi apparatus. It is not yet clear how the QC system operates in the Golgi complex, because the chemical environment of the Golgi differs from that in the ER. The abundant chaperones and folding enzymes that are found in the ER are lacking in the Castro-Fernández et al. • Protein Routing in Health and Disease FIG. 3. Ubiquitin proteasome pathway. Terminally misfolded proteins are retrotranslocated across the ER membrane to be degraded by proteasomes. The misfolded protein is targeted to the Sec61 translocon and then is transported back to the cytosol (step 1). After retrotranslocation from the ER, misfolded proteins are polyubiquitylated before degradation. The lack of polyubiquitylation prevents proteasomal degradation and leads to failure in transport of the protein from the ER lumen or membrane out to the cytoplasm, causing the protein to remain in the ER. The ubiquitin is attached to a cysteine residue of an ubiquitin-activating enzyme (E1), and then transferred to a cysteine of an ubiquitin-conjugating enzyme (E2), and finally to the substrate by the ubiquitin-protein ligase (E3, step 2). The proteins that undergo proteasomal degradation by the proteolytic complex 26S proteasome are marked by the binding of a ubiquitin chain to a lysine residue. The 26S proteasome is composed of a core 20S particle and one or two 19S regulatory particles. The entry channel is quite narrow and, even in its open state, it allows entry of only unfolded proteins in a complex ATP-dependent mechanisms to recognize, unfold, and linearize substrates and to inject them into the 20S proteasome (step 3). These proteins are digested to small peptides, which in turn are rapidly hydrolyzed to amino acids by peptidases (step 4). After proteolysis, the polyubiquitin chain is released and hydrolyzed into single ubiquitin moieties that go through a new cycle of protein degradation. Golgi; the Golgi has a different ionic milieu and less soluble protein content and is more acidic than the ER. However, the intermediate compartment has calreticulin, glucosyltransferase, and glucosidase II, indicating a potential for glycoprotein folding to occur (35, 36). A Golgi-specific endo-␣mannosidase (“endomannosidase”) has been demonstrated to function as a trimming endoglycosidase that has been found to work in conjunction with calreticulin. It has been suggested that this endomannosidase has a role in the QC of N-glycosylation. However, it is not yet clear whether the endomannosidase will dissociate the calreticulin-glycoprotein complex of properly folded glycoproteins and/or of improperly folded ones that have escaped the ER (37). There is a QC mechanism that resides in the Golgi complex, which routes defective proteins to the endosomal system to be degraded (38). When the quantity of misfolded proteins increases in the ER, some move to the intermediate compart- Endocrine Reviews, June 2005, 26(4):479 –503 483 ment and to the cis-Golgi (30). Proteins that are targets for lysosomal or vacuolar degradation can either be targeted directly to the endosomal system or can be first routed to the plasma membrane before they are delivered to the endosomal system (12). It is possible that the QC mechanism for some misfolded or aggregated proteins is initiated in the endosomes, thereby precluding these proteins from reaching the cell surface (39 – 41). Golgi complex can be a reservoir for misfolded membrane proteins, but whether the proteins are degraded or remain in the Golgi for a long period of time is unknown (42). Misfolded membrane proteins tend to expose polar residues, which are recognized as defects by the QC. Retrieval of membrane proteins from Golgi to the ER can be triggered by the presence of these polar residues in the transmembrane domains. These polar residues can also cause entry from the Golgi or cell surface into the endosomal system. A multispanning membrane protein of the Golgi, Tul1, which is a putative ubiquitin ligase, is an enzyme that recognizes polar residues and might separate membrane proteins that are incorrectly folded in the transmembrane region or not completely oligomerized for vacuolar targeting (7). Evidence based on lysosomal enzyme inhibitors suggests that the turnover of secretory pathway membrane proteins may happen by targeting from the trans-Golgi network to the endosome/lysosome system (43). The QC mechanisms in the Golgi complex became particularly important for exportable proteins that achieved their posttranslational modification in the Golgi apparatus. Native or mutant proteins can be directed to degradation to maintain the local environment of the secretory pathway. C. Mitochondria The mitochondrion is an organelle that consists of an outer and inner membrane that separates the matrix from the intermembrane space. Mitochondria provide the majority of the ATP to the cell by coupling the oxidative phosphorylation with respiration. The mitochondrial respiratory chain is composed of five enzyme complexes: reduced nicotinamide adenine dinucleotide-Coenzyme Q reductase (complex I), succinate Coenzyme Q reductase (complex II), ubiquinolcytochrome c reductase (complex III), cytochrome c oxidase (complex IV), and F1F0-ATPase (complex V). In the respiratory enzyme complexes these components transfer electrons to each other and ultimately to molecular oxygen; they also translocate protons across the inner mitochondrial membrane. This proton gradient activates the ATP synthase that catalyzes the conversion of ADP to ATP, completing then the process of oxidative phosphorylation (44, 45). Studies in yeast suggest that a protein QC system exists in mitochondria similar to that in the ER (46). Proteins that cross the mitochondrial membranes do so in an extended conformation, being coupled to the processes of protein unfolding in the cytosol and protein refolding in the matrix. Precursor proteins destined for the mitochondrion are escorted to their proper cellular location and are maintained in a conformation compatible with import by molecular chaperones within the cytosol, such as the heat shock protein 70 (Hsp70) and mitochondrial import stimulation factor. Both of these chap- 484 Endocrine Reviews, June 2005, 26(4):479 –503 erones exploit the inherent instability of the precursor in unfolding the protein so that it can properly interact with the import machinery. Once the precursor polypeptide translocation into the matrix is complete, the precursor must adopt its mature conformation to be able to function in the organelle. Molecular chaperones in the mitochondria are involved in translocation, refolding, and assembly of imported proteins and proteins encoded in the mitochondria, and protein degradation (47, 48). After completion of translocation, folding routes may differ for individual proteins. Some proteins can fold directly after release from the membrane-associated form of Hsp70, or they can be transferred from there directly to Hsp60, chaperonins that mediate the folding of proteins to their native state, or into monomers that subsequently undergo oligomeric assembly (49). There are several other matrix proteins, in addition to the molecular chaperones, that are also involved in protein maturation and folding, such as peptidases and peptidyl prolylisomerases. However, the peptidases have no access to the precursors while they are attached to chaperones. When there is an unfavorable condition that provokes proteins to become unfolded and denatured, the molecular chaperones, in addition to binding proteins that are folding de novo, bind to (preexistent) proteins that have become unfolded and maintain them in a soluble state before the favorable conditions are restored. This situation is also observed when proteins are not able to fold correctly because of a chemical modification, lack of assembly partners, or because of mutation. These misfolded proteins will be targeted for degradation. The chaperones help to maintain these misfolded proteins in a soluble state, to prevent them from aggregation. If there is a lack of correct chaperone function, then these proteins aggregate and cannot be degraded. The ATP-dependent degradation of misfolded proteins in the mitochondrial matrix is accomplished by the proteolysis in mitochondria 1 protease (47). D. Chaperones: the role of folding in QC and routing Molecular chaperones are proteins that interact transiently with unfolded or partially folded intermediates, stabilizing exposed residues with particular physicochemical characteristics and preventing them from forming inappropriate intra- or intermolecular contacts (50). Chaperones increase the efficiency of the overall folding process by reducing the probability of aggregation and by protecting proteins as they fold and by rescuing misfolded and even aggregated proteins and enable them to have a second chance to fold correctly (1, 51, 52). The concentrations of many chaperones are substantially increased during cellular stress. Chaperones frequently utilize a cycle of ATP-driven conformational changes to fold or refold their targets. Protein folding typically takes place in the cytoplasm, for cytoplasmic proteins, or in the ER, for transmembrane and secretory proteins (53). The Hsps have been divided into groups according to their mol wt. They are classified in two main groups: high mol wt ATP-dependent Hsp that requires cochaperones to modulate their conformation and ATP binding. This group includes four major families (Hsp70, Hsp60, Hsp90, and Hsp100). The Castro-Fernández et al. • Protein Routing in Health and Disease second group of chaperones includes the small ATP-independent Hsp, such as Hsp27, that trap denatured proteins on their surfaces. The basic reactivity of all chaperones is similar, but their structural characteristics differ; they participate in very diverse cellular processes, and they can be targeted to different subcellular compartments (53, 54). Misfolded proteins can interact with one or more of the Hsp70 class of chaperones, which protects them against aggregation (55, 56). Hsp70 molecules form multichaperone complexes with cochaperones such as the Hsp40, and they can function in an ATP-dependent process to catalyze the refolding of denatured or partially denatured molecules into enzymatically active forms (51, 57, 58). Certain combinations of chaperones, e.g., Hsp70, Hsp104, and Hsp40, are capable of disassembly of intracellular protein aggregates and acceleration of the refolding of the insoluble molecules into soluble, native species (21, 59 – 61). The molecular chaperones that are used in QC are abundant in the ER. Retention-based, chaperone-mediated QC in the ER is well characterized in the case of glycoproteins. The two homologous lectins, calnexin, which is a transmembrane protein, and calreticulin, which is a soluble luminal protein, operate as chaperones in the QC of newly synthesized glycoproteins, ensuring that correctly folded glycoproteins leave the ER, and they can also mediate glycoprotein degradation of the misfolded protein (15, 62– 66). The lectinassisted glycoprotein folding involves cycling of de- and reglycosylation until correct folding is achieved. If the protein is correctly folded, it will leave the ER. Misfolded it will be recognized by the ER enzyme UGT and reglycosylated, allowing it to reassociate with the QC machinery. The protein will be able to get completely folded as long as it goes through the calnexin-calreticulin cycle repeatedly. Both calnexin and calreticulin suppress aggregation of misfolded proteins via a polypeptide-binding site in their globular domains. Calnexin and calreticulin form complexes with ERp57, a member of the PDI family, which form transient disulfide bonds with calnexin- and calreticulin-bound glycoproteins, facilitating correct disulfide bond formation. PDI is one of the chaperones transcriptionally up-regulated when unfolded proteins accumulate in the cell (67). One of the most abundant ER molecular chaperones is BiP, the main protein of the Hsp70 family. It stabilizes newly synthesized polypeptides during folding, prevents aggregation by binding to and stabilizing exposed hydrophobic regions of unfolded proteins, mediates retention in the ER, suppresses formation of nonnative disulfide bonds, and has a role in the translocation of newly synthesized proteins across the ER membrane and in degradation of misfolded proteins. BiP is also up-regulated when misfolded proteins accumulate in the ER (7). Chaperone ERGIC-53 is a lectin that serves as an escort protein carrying high-mannose N-linked glycans, which cycles between the ER and the Golgi complex; this molecule also helps to prevent proteins from interacting with ligands during the early secretory pathway. When there is an excess of misfolded proteins due to physical or chemical stresses or to mutations, there is an increase in chaperone and proteolytic activities, to reduce the presence of damaging nonnative proteins. Several chaper- Castro-Fernández et al. • Protein Routing in Health and Disease ones can stabilize proteins undergoing denaturation and allow their refolding when the conditions are back to normal. However, when the chaperone’s function on protein stabilization is not optimal, it results in protein aggregation. III. Diseases Caused by Defective Routing A. G protein-coupled receptors (GPCRs) in disease states 1. HH. HH is characterized by absent or decreased gonadal function and manifests as delayed or disrupted sexual development in affected individuals; it constitutes one of the most common causes of hereditary hypogonadism. Mutations in the GnRHR gene are one etiology of HH. Heterologous expression of these GnRHR alleles reveals that the majority encode misfolded receptors with an increased propensity for targeted degradation or intracellular retention. Mutations are predominantly in transmembrane segments, yet (somewhat surprisingly) are widely dispersed across the entire sequence of the GnRHR (68 –70). In HH due to GnRH resistance, affected individuals are either compound heterozygous or homozygous for the mutation. To date, 16 inactivating mutations in the human (h)GnRHR gene and a splice junction mutation at the intron 1-exon 2 boundary have been described as causes of HH and were isolated from patients with phenotypes from partial to complete hypogonadism. The majority of these misfolded translation products do not route correctly to the cell membrane, where the mutant protein is otherwise fully functional (71). The expression of 13 naturally occurring hGnRHR point mutations shows that 11 of these mutants become nonfunc- Endocrine Reviews, June 2005, 26(4):479 –503 485 tional misrouted proteins that are rescuable by genetic or pharmacological approaches (Fig. 4; Refs. 68, 72, and 73). The coexistence of mutant and wt receptors may yield multimeric complexes that are impeded from attaining a conformation consistent with cell surface transport (74, 75). GnRHR heterooligomerization of the misrouted protein with the wt receptor (76) may explain the dominant-negative effect of the mutant form (75). Mutant receptors are capable of pharmacological rescue and, when rescued, these receptors no longer exert a dominant-negative effect. Pharmacological chaperone treatment approximately doubles the expression of wt hGnRHR (but not rat GnRHR) at the plasma membrane. Based on this observation, it has been suggested that more than half of the synthesized hGnRHR never arrives at the plasma membrane. The disparity between the human and rat sequences appears related to specific structural differences in the human sequence that result in instability of the human sequence (77). One mutant that results in HH is the E90K mutation of the hGnRHR gene; this mutant may be rescued by deleting the amino acid K191, an “extra” amino acid in primate GnRHR (not found in rodents) that appears to destabilize the hGnRHR, thus precluding movement of much of the translation product to the plasma membrane (72). Patients bearing the E90K defect express a severe phenotype of hypogonadotrophic hypogonadism (78). Peptidomimetic, cell membrane-permeant GnRH antagonists from three different chemical classes have been shown to serve as templates (pharmacological chaperones or “pharmacoperones”) for naturally occurring and “designer” (misfolded) GnRHR mutants and thereby effect rescue by stabilizing FIG. 4. hGnRHR mutant rescue with a complex indole and peptidomimetics antagonist of the GnRHR “IN3” (264). HH is characterized by absent or decreased gonadal function and manifests as delayed or disrupted sexual development in affected individuals and constitutes one of the most common causes of HH. Mutations in the GnRHR gene are one etiology of HH. Heterologous expression of these GnRHR alleles reveals that the majority encode misfolded receptors with an increased propensity for targeted degradation or intracellular retention. Mutations are predominantly in transmembrane segments, yet (somewhat surprisingly) widely dispersed across the entire sequence of the GnRHR. The majority of these misfolded translation products do not route correctly to the cell membrane, where the mutant protein is otherwise fully functional. The expression of 13 naturally occurring hGnRHR point mutations, shows that 11 of these mutants become nonfunctional misrouted proteins that are rescuable by a cell membrane-permeant antagonist of GnRH, IN3. Most of misfolded hGnRHR mutants are rescued by IN3, which allows them to escape the ER and avoid a degradation pathway. 486 Endocrine Reviews, June 2005, 26(4):479 –503 the GnRHR (68, 69, 71). Unexpectedly, even the use of green fluorescent protein or “hemagglutin-tags” on receptor mutations, has a dramatic effect on the wt receptor and resulted in altered routing (79), a finding that, presumably, is not unique to this receptor. 2. RP. Rhodopsin (Rh) is the visual pigment protein of vertebrate rod cells and is an integral membrane glycoprotein. Rh synthesis requires membrane insertion of the polypeptide opsin, and it has a covalently bound reverse agonist, the chromophore 11-cis-retinal, which interacts with amino acid residues located in transmembrane ␣-helices (80, 81). To date, more than 150 opsin mutants have been identified that cause RP, a heterogeneous group of inherited disorders that cause progressive retinal degeneration and that lead to photoreceptor death and subsequent severe loss of peripheral and night vision. In humans, more than 25% of all RP cases are caused by dominant mutations in the gene encoding Rh, producing misfolded Rh, which, presumably, interferes with the maturation of wt molecules in the ER (82). The opsin mutants are divided into three different classes depending on their distinct biochemical features. Class I mutants that accumulate at the cell surface; Class II mutants retained in the ER; and Class III mutants mainly retained in the ER but also found on the cell surface (80, 81). The most common mutation linked to RP is the P23H mutation, which is a Class III type mutant, located in the N-terminal tail within one of the -strands in close proximity to residues that form the retinal plug, which keeps the chromophore in its proper position. P23H-opsin is retained intracellularly, predominantly in a perinuclear region. The P23H mutant leads to formation of conjugated hydrogen bonding that exposes the chromophore-binding site. Mutations located within this region result in improper folding of opsin and poor binding of the chromophore. The P23H mutant has a structural alteration that decreases the stability of the protein, leading to aberrant folding and subsequent aggregation. However, there is a pharmacological chaperone, an 11-cis-retinal analog that facilitates proper folding and stabilization of P23Hopsin (80, 81). 3. Nephrogenic diabetes insipidus (NDI). In nephrogenic diabetes insipidus (NDI), a disease that can be inherited or acquired, the kidney is unable to concentrate urine in response to the antidiuretic hormone arginine vasopressin, despite normal or elevated plasma concentrations of such hormone. NDI describes only those conditions in which arginine vasopressin release fails to induce the expected increase in the permeability of the cortical and medullary collecting ducts to water. Vasopressin regulates body water, resulting in water reabsorption from urine, by fusion of vesicles containing aquaporin-2 (AQP2) water channels with the apical membrane of renal collecting ducts, this being the final step in the antidiuresis-signaling pathway that leads to increased water permeability in the duct. Mutations in either the vasopressin type 2 receptor (V2R) gene or the AQP2 gene can cause congenital NDI. Mutations in the V2R gene cause the X-linked form of inheritance of NDI, and mutations in the AQP2 gene cause a rare non-X-linked form of inheritance of NDI (83– 85). Castro-Fernández et al. • Protein Routing in Health and Disease There are 155 known mutations within the V2R and, as in the case of the GnRHR, the majority are in the transmembrane segments of the V2R, but are broadly distributed in the sequence. More than 90% of these mutations failed to properly fold and therefore have defective intracellular transport, being retained in intracellular compartments instead of being transported to the cell membrane. In some cases, accumulation of V2R mutants away from the plasma membrane, possibly in the ER, occurs because there is little maturation (sugar trimming) of the glycosyl-groups (86). In the AQP2 gene there have been several point mutations identified that are associated with NDI. Autosomal-recessive NDI-causing AQP2 mutations have a significant defective protein trafficking that results in retention of the mutant protein in the ER, although they do not present major differences in intrinsic water permeability or protein stability. A mutation in the AQP2 gene, in which there was a substitution of a lysine for a glutamic acid at position 258 (AQP2E258K), has been shown to result in an autosomal-dominant form of inherited NDI. This mutant is localized mainly in the Golgi complex region in Xenopus oocytes, and it is stable as the wt AQP2 (85). The AQP2-E258K mutant forms mixed oligomers with the wt AQP2 and has a dominant-negative effect on the function of the wt form, which might explain the dominant inheritance of NDI, whereas in the recessive inheritance of NDI there appears to be a lack of oligomerization of the mutant AQP2 protein with the wt AQP2 (84, 85). The AQP2-E258K mutant exits the ER, after which it heterotetramerizes with wt AQP2. AQP2 mutants seem to have an impaired routing and thereby differ in their retention site. For example, the T126M AQP2 mutant is retained in the ER and rapidly degraded by the proteasome; in contrast, the E258K AQP2 mutant is retained in the Golgi and then degraded by proteasomes and lysosomes (87). B. Wild-type protein degradation and oligomerization in protein processing There is evidence showing that in many cells, approximately 30 – 60% of the newly synthesized proteins never achieve their completely folded native structure and are thereby targeted for degradation. This is documented in the case of the GnRH and the opioid receptors, where exit from the ER is the limiting step in processing of the newly synthesized receptors. Frequently, receptor folding intermediates are retained in the ER until the correct conformation is attained. In the case of the opioid receptors, there is an apparent inefficiency of the maturation process because only 40% of the precursor receptor was converted to the mature form and reached the cell surface, whereas the other 60% is retained in the ER and degraded (88). For the human GnRHR, about half of the synthesized receptor is expressed at the plasma membrane. This is not the case for the rodent GnRHR, which is more stable and is almost completely routed to the cell surface. It has been shown that in the presence of pharmacological chaperones, there is a higher expression of hGnRHR at the plasma membrane (75). For many years the GPCRs have been believed to be mo- Castro-Fernández et al. • Protein Routing in Health and Disease nomeric structures, each one being independently activated by the binding of a single ligand. These receptors then activate G proteins. However, this classical model may be oversimplified because an increasing number of studies have demonstrated that many GPCRs exist (in the plasma membrane) as oligomers or heterooligomers (89). The existence of a dimeric GPCR was suggested by using bivalent antibodies directed against a peptide antagonist of the GnRHR; this antibody converted the antagonist into an agonist and promoted receptor activation (90). It has been demonstrated that several GPCRs form constitutive dimers during biosynthesis. For example, the dimerization of the metabotropic glutamate receptor (mGluR1a) takes place in the ER in the HEK293 cell line (91). Also it has been shown that oxytocin, vasopressin V1a, and V2 receptors are present as dimers in an ER-enriched fraction (92). Several studies also suggested that some GPCRs form dimers in the presence of agonists and that sometimes the formation of a constitutive dimer is dependent on receptor density (93–95). It has been assumed that GPCR mutants do not have an effect on the function of the wt receptor. However, it has been demonstrated that heterodimerization between the wt and the mutant receptors contributes, at least in part, to the dominant-negative phenotypes associated with heterozygous mutations. A loss-of-function CCR2(Y139F) mutant acts as a dominant-negative agent, blocking signaling through the wt CCR2 (93). Other examples include the V2R (96) and D2 receptors (97). In the case of GnRHR and the fly Frizzled family of Wnt receptors, the mutant forms heterooligomers between the wt and the dominant-negative receptors, trapping them both in the ER (76, 98). Sometimes, the simultaneous presence of two different GPCR moieties is necessary to constitute a functional receptor. For example, neither type 1 nor type 2 ␥-aminobutyric acid B receptor (GABABR1 nor GABABR2) when expressed individually, activates GIRK-type potassium channels. However, the combined expression of GABABR1 and GABABR2 produces a oligomer that is capable of stimulation of channel activity upon binding agonist by the complex (99). The heterodimerization of the GABABR1 and GABABR2 in the ER is necessary for exporting GABABR1 to the cell surface (100, 101). In the heterodimeric complex GABABR2 is required to activate G proteins, and GABABR1 is responsible for the binding of its ligand (100). The retention of the wt receptor proteins as well as the formation of oligomers seems to modify the overall processing of the newly synthesized proteins. The retention of the wt and mutant receptors by oligomerization in the ER could be associated with different diseases. Further analysis of such mechanisms will likely lead to a better understanding of the regulation of GPCRs. C. Other proteins 1. Cancer (p53). The tumor suppressor p53 activates numerous genes and leads to cell cycle arrest or apoptosis, protecting the cell against tumor growth. Because of its growthinhibitory activity, p53 must be maintained at low levels to ensure cell survival and proper organism development; this is regulated by MDM2, which inhibits the function of p53 by Endocrine Reviews, June 2005, 26(4):479 –503 487 targeting it for ubiquitin-dependent proteasomal degradation (102). The mutant form of p53 loses the ability to induce cell cycle arrest or apoptosis. Approximately 50% of the human cancers lose p53 function mainly by point mutations in the core domain, which results in loss of sequence-specific DNA binding and transcription function. These mutations mainly affect residues that are directly involved in contacting DNA and that are important for maintaining the structural stability of the highly conserved DNA-binding domain by thermodynamically destabilizing p53, which results in its unfolding and inactivation (Fig. 5; Ref. 103). The core domain is relatively unstable, because the majority of residues in the p53 core domain are targets of mutation in tumors. A structural destabilization and partial global unfolding of the core domain of p53, or a local distortion of the conformation required for DNA binding, or both, can result in malfunction of the p53 tumor-associated mutants in different ways, depending on the position and nature of each mutation (104). Misfolded mutants have a dominant-negative effect by forming functionally inactive heterooligomers with the wt p53, as the mutant forms expose extended hydrophobic surfaces; such surfaces aggregate with associated wt subunits and disturb their structure. The R249S mutant form of p53, one of the most common cancer-associated mutations in this gene, is located in the third loop, causing local structural changes around the mutation site (105). Missense single-point mutations are the most common alterations in the p53 gene; however, many cancers lose p53 activity by inhibition of the wt protein rather than by mutation. Tumor growth suppression can be achieved by reintroduction of wt p53, and intermittent threshold levels of p53 activity may be sufficient, functioning as an important target in the development of cancer therapies. Various strategies FIG. 5. Mutant p53 pathways that result in destabilization, unfolding, and inactivation. The p53 point mutations mainly affect residues that are localized in the core domain. A, Unfolded p53 loses sequencespecific DNA binding and transcriptional function by blocking interaction with target DNA. B, Misfolded mutants have a dominantnegative effect by aggregating with wt p53, thus forming functionally inactive oligomers that are incapable of entering the nucleus. C, External small synthetic compounds can bind to p53, stabilizing the DNA-binding conformation, allowing nuclear transport and normal function. 488 Endocrine Reviews, June 2005, 26(4):479 –503 have been designed to restore function to mutant p53. The restoration of p53 function, based upon molecular modeling, introduced additional DNA contacts or enhanced the stability of the folded state. One strategy to reactivate destabilized p53 mutants leading to thermodynamic stabilization and activity restoration is by using a small molecule that binds the native, but not the denatured, state and drives the conformational equilibrium toward the native state. Furthermore, the introduction of second-site suppressor mutations, by site-directed mutagenesis of single amino acid substitutions can restore, at least partially, the conformation for DNA binding or the stability of the protein (103, 106). Moreover, it has been shown that specific DNA binding of mutant p53 as well as p53-dependent apoptosis in tumor cells can be restored by synthetic peptides that have been derived from the p53 C terminus. 2. Cystinuria. The membrane glycoprotein rat liver bile acid coenzyme A: amino acid N-acetyltransferase (rBAT) is localized at kidney and small intestine membranes, which are sites for high-affinity l-cystine absorption. This glycoprotein produces a large amount of amino acid transport. Mutations of the membrane glycoprotein rBAT have been found to cause the inheritable disorder cystinuria type I, which is an autosomal-recessive failure of dibasic amino acid transport across the luminal membrane of kidney cells and also affects intestinal absorption of cysteine and provokes a low solubility of cysteine, leading to a severe development of cysteine “calculi,” which are abnormal stony masses in the kidney commonly known as kidney stones (107). So far, 32 cystinuria-specific mutations of rBAT have been described in patients with cystinuria, which include missense and nonsense mutations, deletions, and insertions. The most common point mutation is M467T, localized in the third transmembrane domain and found in 26% of cystinuria type I genes (108). There is a diminution of membrane-expressed proteins M467T compared with wt rBAT, and the intracellular amounts of the mutant form are higher, which suggests that there might be a defect in the delivery of the mutants to the plasma membrane because of their retention in the ER, probably due to improper folding. rBAT is assembled with other proteins to constitute the complete amino acid carrier complex, and this assembly has slower or impaired activity if rBAT is improperly folded. The transport out of the ER is accomplished after assembly of proteins into native homo- or heterooligomers (108). Another mutant of rBAT, R365W, is located in the extracellular domain of rBAT and was found in some cystinuria patients. This mutant behaves as a temperature-sensitive folding mutant. It is probable that the cystinuria phenotype related to this mutant form of rBAT is a consequence of a severe trafficking defect. Most cystinuria-specific rBAT mutations are trafficking mutants, supporting the proposed role of rBAT in the routing of the holotransporter, formed by two different subunits (the rBAT heavy chain and a rBAT light chain), to the plasma membrane (109). 3. CF. CF is an autosomal-recessive disorder affecting the apical membrane of epithelial cells of the pancreas, bronchial Castro-Fernández et al. • Protein Routing in Health and Disease glands, biliary tree, intestinal glands, reproductive organs, and sweat glands. It is caused by mutations in the CF transmembrane conductance regulator (CFTR) protein, which functions as a Cl⫺ channel and regulator of other ion channels, resulting in Cl⫺-impermeable epithelial cells in the affected organs (Fig. 6; Refs. 110 and 111). Several mutations in the CFTR gene have been identified that exert negative effects resulting in errors in transcription and production of unstable mRNAs that are rapidly degraded, causing failure of the cells to synthesize sufficient quantities of the CFTR protein. Moreover, genetic errors can also cause production of a CFTR protein that fails to mature and fold properly, which results in being trapped in the ERGIC or ER, thereby failing to move further the maturation pathway and reaching the plasma membrane (112, 113). Unlike the hGnRHR and V2R, most patients share a common CFTR mutation, this being a deletion of phenylalanine at position 508 (⌬508). This site is normally in the cytoplasmic domain of CFTR; the mutant remains in the ER and is ultimately degraded by the ubiquitin-proteasome system. The disease in the homozygous patient is quite severe, even FIG. 6. Model of the lung epithelium in CF. In CF, the apical membrane of epithelial cells of the respiratory tract, along with other organs, lose plasma membrane CFTR channel expression. Mutations in the CFTR channel cause the disease, the most common being the ⌬508 CFTR mutation. A, In the normal healthy state of the lung, the CFTR is localized to the apical membrane of the lung epithelium, and Cl⫺, and thus Na⫹ and water, pass freely from the bloodstream into the lungs. B, In a CF patient, the mutant channel remains largely trapped in the cell, thereby impeding the normal flux of Cl⫺, and thus Na⫹ and water, across the epithelium; as a result the airway surface dries out and the mucus accumulates and becomes condensed and thick, impeding the passage of air. C, In the presence of a chemical chaperone or a nonspecific agent such as glycerol, the mutant CFTR is relocalized to the cell surface, recovering the Cl⫺, and thus Na⫹ and water, flux, rehydrating the airway membrane and allowing mucus clearing. Castro-Fernández et al. • Protein Routing in Health and Disease though it has been demonstrated that ⌬F508 CFTR retains some function as a Cl⫺ channel, although less than the wt protein. The CF phenotype associated with ⌬F508 CFTR appears to originate from mislocalization and premature degradation, rather that the altered functional properties or decreased stability of mutant CFTR (110). Misfolding CFTR causes this protein to be targeted to the degradative apparatus before it can be translocated to the plasma membrane. There is evidence for a temperature sensitivity of ⌬F508 CFTR maturation, although there may be no apparent decrease in thermal lability of the mutant protein (114). Calnexin binds to the newly synthesized CFTR and dissociates from it when it reaches its native conformation. In contrast, the ⌬F508 CFTR remains associated to calnexin therefore, after a prolonged interaction leading the ⌬F508 CFTR to ERAD (115). The ⌬F508 CFTR mutant is functional once it reaches the native state. Growth conditions that favor proper folding of the mutant CFTR allow increased recovery of function. Identification of a compound that corrects the ⌬F508 CFTR folding defect would be of great potential therapeutic benefit, and some studies that utilize small molecules identified from screening techniques are underway. CFTR activators may also correct the impaired chloride transport in cells expressing mutant CFTRs in CF patients (116). 4. Gitelman’s syndrome. Gitelman’s syndrome is an autosomal-recessive disease characterized by renal salt wasting with hypokalemic metabolic alkalosis, hypomagnesemia, and hypocalciuria (117). The first clinical presentation usually occurs at 6 yr of age or older and may include muscular pain and cramping with exercise and fatigue and weakness. The disease is generally mild, and some patients even remain without symptoms (118, 119). This disease is caused by mutations in the SLC12A3 gene, which encodes the thiazidesensitive NaCl cotransporter (NCCT; Fig. 7). The mutation in the NCCT gene results in salt wasting from the distal convoluted tubule, increasing aldosterone secretion, which augments Na⫹ reabsorption. Because of the indirect coupling of Na⫹ reabsorption to K⫹ and H⫹ secretion in the distal nephron, the augmented Na⫹ reabsorption activity occurs at FIG. 7. Model of the NCCT. The tertiary structure of NCCT contains 12 transmembrane domains and intracellular amino- and carboxy-terminal regions. The mutations identified in Gitelman’s syndrome are located throughout the entire coding sequence of the protein. Thirteen human NCCT mutations have been functionally tested in Xenopus laevis oocytes. The missense mutations G439S, T649R, and G741R were partially glycosylated and impaired in their route to the plasma membrane, and instead they are retained in the ER/pre-Golgi complex. In contrast, the rest of the mutations impair their insertions of a functional protein into the plasma membrane. Endocrine Reviews, June 2005, 26(4):479 –503 489 the expense of increased K⫹ and H⫹ secretion, accounting for the observed hypokalemia and metabolic alkalosis (120, 121). The tertiary structure of NCCT contains 12 transmembrane domains and intracellular amino- and carboxy-terminal regions (122). The mutations identified in Gitelman’s syndrome are located throughout the entire coding sequence of the protein and include missense, frameshift, nonsense, and splice-site mutations (123). There are multiple mechanisms by which mutations could reduce or abolish NCCT activity: impaired synthesis or routing, disturbed transport activity, and increased retrieval or degradation. Two different classes of NCCT mutations have been distinguished based on immunocytochemical analyses and the metolazone-sensitive Na⫹ uptake rates: the ERretained mutants and partly functional mutants. The ERretained mutants did not exhibit significant metolazonesensitive Na⫹ uptake, and the immunocytochemical analyses demonstrate that these mutant proteins were expressed predominantly in the cytoplasm and were absent at the plasma membrane. These mutant transporters lacked the typical complex-glycosylated band of 130 to 140 kDa. Together, these findings suggest that these partly glycosylated mutants are impaired in their routing to the plasma membrane and are retained in the ER/pre-Golgi complex (121). The C terminus of NCCT bound to chaperone Grp58, indicating that regulatory proteins may be involved in transporting NCCT ER to other compartments in the cell for further processing (124). In contrast, the partly functional mutants are present at the plasma membrane, as demonstrated in immunocytochemical analyses, but are also clearly located in the cytoplasm (unlike the wt protein). These findings indicate that the process of routing NCCT to the cell surface is only partly impaired, suggesting that different mechanisms are involved in the disturbed trafficking for the two mutant classes, but the molecular mechanisms responsible remain to be established (121). 5. Hemochromatosis (HC). Hereditary HC is an autosomalrecessive disorder characterized by an abnormal increase in the absorption of iron in the gastrointestinal tract. The organs in which iron deposits primarily are liver, pancreas, heart, 490 Endocrine Reviews, June 2005, 26(4):479 –503 joints, and endocrine organs (125). Clinical consequences of iron accumulation in these organs include cirrhosis of the liver, hepatocellular carcinoma, diabetes, heart failure, arthritis, and hypogonadism (126). The majority of North American patients with HC have a mutation in cysteine 282 (to tyrosine) in a major histocompatibility complex (MHC) class I-like gene, called hemochromatosis gene (HFE) (127, 128). A second mutation (H63D) was reported to be also present in HC patients who are heterozygous for the C282Y mutation (129). The human HFE protein is a membrane protein homologous to MHC class I molecules, which contain an extracellular peptide-binding region (␣1 and ␣2 loops), an Ig-like domain (␣3), a transmembrane region, and a short cytoplasmic tail. HFE contains intramolecular disulfide bridges in the ␣2- and ␣3-domains that stabilize its tertiary structure (127). The disulfide bridge in the ␣3-domain of MHC class I molecules is required for their association with 2-microglobulin (2M), leading to efficient intracellular processing and transport to the plasma membrane (130, 131). The C282Y mutation prevents the interaction between HFE with 2M and eliminates cell surface presentation. The role of 2M/heavy chain interaction is to facilitate and stabilize the folding of the heavy chain during biosynthesis (130, 132). The process by which HFE may affect iron uptake is poorly understood. Recent evidence demonstrated that HFE diminishes the binding affinity of iron-loaded transferrin to its receptor (133, 134). In fact, the binding of HFE and transferrin receptor is required for HFE transport to endosomes (125). However, the physiological relevance of this finding remains unclear. The association of the heavy chain with the 2M occurs in the ER and is then transported through the cis-Golgi network, to the middle and trans-Golgi cisternae where the glycosyl side chain is modified to a more complex form en route to the plasma membrane (135, 136). Class I molecules that fail to assemble properly are recycled between the ER and Golgi, rather than being retained exclusively in the ER (137). 6. Hypercholesterolemia. Plasma low-density lipoproteins (LDL) bind to the LDL receptor (LDLR), a transmembrane glycoprotein that mediates LDL cellular uptake, internalization, and its delivery to lysosomes. Loss of LDLR function leads to decreased LDL catabolism, and its cholesterol is released, resulting in elevated levels of plasma cholesterol. Mutations in the LDLR gene results in protein misfolding and ER retention, causing familial hypercholesterolemia, which is a common autosomal-dominant inherited disorder (138, 139). To date, there are more than 600 mutations that have been characterized in the LDLR; however, 50% of these mutations in familial hypercholesterolemia patients lead to a class 2 mutation, which are proteins that are partially or totally retained in the ER or have an ER-Golgi-defective transport (140). The BiP chaperone has been shown to bind to LDLR by coimmunoprecipitation studies, where an interaction of Grp78 with either the wt or the mutant LDLR was observed, and there was no interaction with other proteins. Moreover, the interaction of Grp78 with mutant LDLR was shown to be a prolonged one compared with that of the wt LDLR, demonstrating that Grp78 is the chaperone respon- Castro-Fernández et al. • Protein Routing in Health and Disease sible for retention of the misfolded receptors in the ER. The mechanisms responsible for this LDLR retention in the ER are not yet clear (138). A prolonged retention of misfolded or incompletely folded proteins in the ER leads to their degradation. The proteasomal degradation pathway mediates the LDLR mutant degradation, but the precise mechanism for the LDLR is still unknown. The receptor-associated protein (RAP) has been identified to function as a specialized chaperone in the folding and trafficking of LDLR along the early secretory pathway. RAP, which binds only to members of the LDLR family, promotes the receptor folding and maturation by preventing the formation of nonproductive intermolecular disulfide bonds and premature ligand binding within the early secretory pathway (141). It is still not known whether RAP is also involved in ER retention and/or the proteasomal degradation of the LDLR mutants (139). IV. Diseases Caused by Conformational Errors The so-called conformational diseases are disorders in which the constituent protein involves change in size or fluctuations in shape with resultant self-association and tissue deposition as amyloid fibrils (142). This diverse group of diseases includes disorders such as Alzheimer’s and Parkinson’s diseases, amyloidoses, ␣1-antitrypsin deficiency, and the prion encephalopathies. These diseases result when a specific protein is subjected to a conformational rearrangement that results in aggregation and deposition within tissues or cellular compartments (143, 144). These disorders have common features: they have either sporadic or inherited patterns, they appear later in life (generally after the fourth or fifth decade), and all are characterized by neuronal loss and synaptic abnormalities. In contrast, the clinical symptoms and disease progression are very different among such disorders (145, 146). Amyloid structures can be recognized because of Congo red and thioflavin S binding and a core structure based on the presence of highly organized -sheets. This characteristic makes them very resistant to proteolytic degradation (147, 148). When these aggregates are formed, they are stable for long periods, forming a bed for the progressive accumulation of more deposits in tissue and eventual conversion into amyloid fibrils (1). The deposits in amyloidoses are extracellular but also include some intracellular aggregates. These can often be observed as thin fibrillar structures, sometimes assembled further into larger aggregates or plaques (Fig. 8; Refs. 146 and 148). The capacity of polypeptide chains to form amyloid structures seems to be a generic feature of polypeptide chains, and it is not restricted to the relatively small number of proteins associated with recognized clinical disorders (149, 150). The polypeptides involved in amyloid diseases include full-length proteins (e.g., lysozyme or Ig light chains), biological peptides (amylin, atrial natriuretic factor), and fragments of larger proteins produced as a result of specific processing (e.g., the Alzheimer -peptide). These disorders may also result from fragments due to general degradation [e.g., polyglutamine stretches cleaved from proteins with polyglutamine extensions such as huntingtin, ataxins, and the androgen receptor (148)]. Castro-Fernández et al. • Protein Routing in Health and Disease Endocrine Reviews, June 2005, 26(4):479 –503 491 FIG. 8. Schematic representation of the possible mechanism of formation of amyloid fibrils. The so-called conformational diseases are disorders in which the constituent protein involves change in size or fluctuations in shape with resultant self-association and tissue deposition as amyloid fibrils. When a specific protein is subjected to a conformational rearrangement that promotes aggregation, it induces the deposition within tissues or cellular compartments resulting in these disorders. The core structure of amyloid fibrils contains highly organized -sheets, which makes them very resistant to proteolytic degradation. When these aggregates are formed, they are stable for long periods, forming a bed for the progressive accumulation of more deposits in tissue and eventual conversion into amyloid fibrils. These can often be observed as thin fibrillar structures, sometimes assembled further into larger aggregates or plaques. The impairment of cellular function is the result of direct interaction between the aggregated proteins with some cellular components. However, it has been demonstrated that, in many cases, the prefibrillar aggregates are more toxic to the cells than mature fibrils. In the conformational disorders the impairment of cellular function is usually the result of direct interactions between the aggregated proteins with a particular cellular component (151, 152). In addition, it has been demonstrated that in many cases the prefibrillar aggregates (also referred to as amorphous aggregated protein micelles or protofibrils) are more toxic to the cells than mature fibrils (for review, see Ref. 148). The absence of a direct correlation between the density of fibrillar plaques in the brains of victims of Alzheimer’s disease (AD) and the severity of the clinical symptoms can be explained by the toxicity exhibited by early aggregates (153). The molecular chaperones and other QC mechanisms are remarkably efficient in ensuring that prefibrillar aggregates are neutralized before they can do any damage. If the efficiency of the QC mechanisms is impaired, the probability of pathogenic behavior is increased. Most of the amyloid diseases are associated with old age, and the explanation could be a decrease in the efficiency of these protective mechanisms, when there is an increased tendency for protein to become misfolded or damaged, coupled with a decreased efficiency of the molecular chaperone and UPRs (154). The increase in human life expectancy is associated with augmentation in the prevalence of amyloid diseases, in particular in highly developed countries; this can be explained by the inefficiency of the QC mechanisms to prevent the protein aggregation (148, 155). A. AD and neuropathies AD is, at present, an incurable neurodegenerative disease and the most common form of dementia (156). The clinical manifestations in AD are the gradual loss of memory, followed by a progressive deterioration of higher cognitive functions and behavior (157), caused by a massive loss of neurons (158). Prevalence is estimated to affect 5– 8% of persons over 65 yr of age, and the incidence of inherited forms ranges from 1–30% (159). Primarily, extracellular plaques and intracellular neurofibrillary tangles characterize the pathology of AD. Plaques are composed mainly of the amyloid--peptide (A), whereas tangles are composed of the cytoskeletal protein (142). The formation of tangles is considered as a secondary event, i.e., a consequence of the formation of the A-plaques (160, 161). The diseases called “tautopathies” are different from AD, and the brain lesions in these patients are the intracellular aggregation of -proteins, without accumulation of Aplaques (162). To date, mutant forms of three genes, -amyloid precursor protein (APP) gene (located in chromosome 21), presenilin 1 (PS1, on chromosome 14), and presenilin 2 (PS2, chromosome 1), have been shown to cause early-onset familial AD (158). In addition, polymorphisms in four genes, i.e., apolipoprotein E (apo E), ␣-2 microglobulin, very low density lipoprotein receptor, and low density lipoprotein receptor-related protein, are shown to be risk factors for AD pathogenesis (158, 163). Amyloid formation by the Alzheimer’s A-protein is the most extensively studied. The A-peptides are degradation products of a transmembrane cell surface glycoprotein APP (Fig. 9). Three different proteases called ␣-, -, and ␥-secretases can cleave APP. When APP is concomitantly hydrolyzed by -secretase at the N terminus and by the ␥-secretase within the membrane, the two main products, A(1– 40) and A(1– 42), migrate outside the cell and produce fibrils (164). Peptides containing the 40- or 42-residue form of A, and shorter derivatives, form amyloid-like fibrils in vitro, with characteristics similar to fibrillar aggregates extracted from AD amyloid plaques (146, 165, 166). The amyloid cascade hypothesis predicts that the overproduction of A, or the failure to eliminate this peptide, induces AD primarily through amyloid deposition, which is postulated to be involved in neurofibrillary tangle formation (167–169). These lesions are then associated with cell death and consequent 492 Endocrine Reviews, June 2005, 26(4):479 –503 Castro-Fernández et al. • Protein Routing in Health and Disease FIG. 9. Hypothetical mechanism of conformational change for A-peptides. Three different proteases called ␣-, -, and ␥-secretases can cleave APP. When APP is concomitantly hydrolyzed by -secretase at the N terminus and by the ␥-secretase within the membrane, the two main products, A(1– 40) and A(1– 42), migrate outside the cell and produce amyloid fibrils. memory impairment (142, 170, 171). The precise involvement of amyloid in sporadic cases of AD and the formation of neurofibrillary tangles should be explained. Presenilins (PS1 and PS2) have functions similar to ␥-secretases and cleave APP to release A; these proteins also cleave the Notch receptor to release the Notch intracellular domain, activating a signal transduction pathway by their cytoplasmic domain (172). Thus, presenilins play an important role in AD pathogenesis. Tangle formation is also coupled by further increase of production, hyperphosphorylation, and loss of microtubule binding in the affected neurons. When the C terminus of is subjected to proteolysis, it generates cytotoxic fragments (161, 173). These and other fragments are thought to nucleate into tangle formations (174). B. Parkinson’s disease (PD) PD is the second most common, late-onset neurodegenerative disorder. The degeneration of dopaminergic neurons in the substantia nigra resulted in the typical characteristics of PD, which include muscular rigidity, postural instability, and resting tremor. The clinical hallmark of PD is the deposition in brain cells of intracytoplasmic inclusion bodies called Lewy bodies (LBs) (158, 159, 175). Inherited forms of PD are caused by mutations in three genes, ␣-synuclein, parkin, and ubiquitin carboxyl-terminal hydrolase L1(158, 175, 176). Autosomal-dominant PD results in mutations in the ␣-synuclein gene, and autosomal-recessive PD is due to mutations in the parkin gene. Among other factors, apolipoprotein E ⑀2/⑀4 genotype and exposure to pesticides might be risk factors for sporadic PD (177, 178). The ␣-synuclein is a highly conserved 140-amino acid protein very abundant in the central nervous system, particularly in the presynaptic terminals (179, 180), but its physiological functions are largely unknown. This protein was originally identified as a precursor of the nonamyloid component of the senile plaques in AD (181) and is a major constituent of LBs and other disease-specific lesions such as the abnormally shaped neurites known as Lewy neurites (182–184). The ␣-synuclein is also found in the intracellular inclusions of other neurodegenerative disorders, such as dementia with LB and the LB variant of AD commonly referred to as a “␣-synucleinopathies,” and also in the filamentous glial and neuronal inclusions of multiple-system atrophy (175, 185). It has been reported that ␣-synuclein inhibits phospholipase D2 (186), a plasma membrane protein that converts the neutral phosphatidyl choline to phosphatidic acid and promotes cytoskeletal reorganization and directed movement of vesicles (187). It also binds protein , promoting its protein kinase A-dependent phosphorylation (188). This protein also has been reported to act as a modulator of the activity of tyrosine hydroxylase, the rate-limiting enzyme in dopamine (DA) synthesis, and as a regulator of the DA transporter (162, 175). It is possible that ␣-synuclein may protect and help neurotransmitter vesicles to transport them from cell body to synapse, and its aggregation may exert its toxicity by creating pores in the lipid membrane, leading to loss of DA from vesicle to cytoplasm (158). In a German family with autosomal-dominant form of PD a mutation in the gene ubiquitin carboxyl-terminal hydrolase L1 was found (189). This mutation appears to be extremely rare and is characterized by a relative early-onset appearance in life (⬃50 yr). No histopathological analysis is yet available, and it is not known whether LBs form. Parkin is a highly conserved 465-amino acid protein with a molecular mass of 52 kDa, mainly expressed in brain (especially in midbrain DA cells) and muscle (190, 191). Parkin is reported to function as an E3 ubiquitin ligase and disease-linked mutations that indeed seem to compromise its enzymatic activity. C. Huntington’s disease (HD) The expanding CAG repeats coding for polyglutamine are the cause of nine neurodegenerative disorders (192). They include spinal and bulbar muscular atrophy, HD, dentatorubral and pallidoluysian atrophy, and several forms of spinocerebellar ataxia. Each one is characterized by selective neuronal death in specific regions of the brain (193). HD is an autosomal dominantly inherited neurodegenerative disorder characterized by the progressive development of mood disturbances, behavioral changes, involuntary choreiform movements, and cognitive impairments. Onset is most commonly in adulthood, with a typical duration of 15–20 yr before premature death (194). The genetic defect in HD is the expansion of an unstable Castro-Fernández et al. • Protein Routing in Health and Disease CAG repeat encoding glutamines (Q), close to the 5⬘-end of the gene for huntingtin protein (194). The disease phenotype is influenced by CAG repeat length: the presence of a trinucleotide stretch to 34 –39 CAG repeats in one allele confers risk of developing HD, but the extension above 39 repeats conditions the appearance of the disease including age of onset and the extent of DNA fragmentation in striatal neurons (Fig. 9C; Refs. 195–197). HD also shows the anticipation characteristic resulting from instability of the repeat size during transmission (198). Huntingtin is predominantly cytoplasmic; its normal function is not well understood but may involve cytoskeletal function or vesicle recycling. It has been proposed that the protein may be transported to the nucleus and have a role in the regulation of gene transcription, but this is uncertain (199). The toxicity of huntingtin in the cytoplasm may be caused by mutant full-length protein or by cleaved protein, and the toxic species may be soluble monomers or oligomers or insoluble aggregates. It has been proposed that the toxicity of this protein may involve inhibition of the proteasome or activation of caspases directly or via mitochondrial effects. It is possible that aggregation of huntingtin induces oxidative and ER stress and proteosomal and mitochondrial dysfunction (200, 201). Cytoplasmic aggregates accumulate in perinuclear or neuritic regions and are ubiquitinated. The mutant protein translocates to the nucleus, where it forms intranuclear inclusions; however, nuclear toxicity is believed to be caused by interference with gene transcription (193, 194). The Huntingtin gene is expressed in all cells but it affects only a subset of neurons. This cell type specificity may suggest the requirement for another protein for its function or it may be specifically blocking the trophic factor support for striatal neurons (158). D. Viral prions Transmissible spongiform encephalopathies (TSEs), or prion diseases, are degenerative disorders of the central nervous system that lead to motor dysfunction, dementia, and death. Prion diseases include scrapie of sheep, bovine spongiform encephalopathy (BSE) in cattle, and such human diseases as Creutzfeldt-Jackob disease (CJD), GertstmannSträussler Scheinker syndrome, and fatal familial insomnia. More recently, variant CJD (vCJD), attributed to consumption of BSE-contaminant products, have claimed more than 100 victims (202). Neither humoral nor cellular immunological responses have been detected in prion diseases (203). The precursor protein PrP 33-to 35-kDa (PrPc) is coded by a normal cellular gene and not by a putative nucleic acid carried within the prion (204). The PrPc is found predominantly on the outer surface of neurons, attached by a glycosylphosphatidyl inositol anchor (205, 206). In prion diseases a large protease-resistant aggregated form of PrP, designated PrPsc, that accumulates mainly in the brain, is believed to be the principal or only constituent of the prion (203, 207). No differences in the primary structure of PrPc and PrPsc have been detected, suggesting that they differ only in their conformation (208). The tertiary structure of PrPc has been elucidated, and the -sheet content is low; the structure Endocrine Reviews, June 2005, 26(4):479 –503 493 of the PrPsc is not currently available but the -sheet content is high (209 –211). The refolding and the nucleation models have been proposed to explain the mechanisms by which PrP aggregates. In the refolding model, the PrPc unfolds to some extent and refolds under the influence of a PrPsc molecule, and an activation energy barrier separates the two states. On the other hand, the nucleation model proposes that PrPc is in equilibrium with PrPsc, which is only stable when it forms multimers. Once a multimer is present, monomer addition follows rapidly (212, 213). PrPc has been implicated in a chaperone function for nucleic acids, but its potential role for aggregation of other proteins has not been studied (214). TSE diseases are transmissible by inoculation or ingestion of material contaminated by infected tissues. Incubation periods before the appearance of clinical symptoms range from months to years and, in the case of some kuru patients, may have been as long as 40 yr (215). 1. Kuru and iatrogenic CJD. In both diseases the patients are directly exposed to the TSE agent by contact with brain tissues or extracts contaminated by the TSE agent. Kuru is transmitted during the cultural handling (or ingestion) of brain tissue of relatives who died of this disease (215). Iatrogenic CJD has been induced by transplantation of corneal or dural tissue from patients with TSE, or by neurosurgery using instruments not completely sterilized after use on TSE patients (216). It has also been detected after inoculation of GH extracted from pituitary glands pooled from large groups of individuals (217). 2. vCJD. In 1996, authorities in the United Kingdom described vCJD, a new CJD that is now believed to be the human form of BSE [mad cow disease (202)]. vCJD can be distinguished from sporadic CJD, which is the usual form of human TSE disease, by the early age on onset, absence of electroencephalogram changes, and distinct neuropathological features (218). Because there is no association of occupational exposure of vCJD patients to cattle on farms or in abattoirs, it is likely that spread may have occurred through consumption of BSE-contaminated meat products (215). 3. Sporadic CJD. The age of onset is usually between 50 –70 yr, but can range as low as 16 yr and as high as 80 yr. The primary clinical symptom is usually dementia, which often presents with cognitive disturbances such as confusion, memory loss, and bizarre behavior (219, 220). The clinical symptoms are not associated with a mutant PrP allele; there is also no epidemiological evidence for exposure to TSE agent although contact with people or animals with TSE disease is present (216). At present, the means of acquisition of the TSE agent in these patients is unknown. Several hypotheses include exposure to an unidentified virus, spontaneous generation of nonviral agent through some somatic cell mutation of PrP in each disease individual, and stochastic initiation of spontaneous PrPsc formation without PrP mutation. However, there is no evidence for spontaneous PrPsc formation in any animal or human disease. 4. Familial TSEs. Familial TSE is an autosomal-dominant disease caused by a mutation in the PrP gene (221). These 494 Endocrine Reviews, June 2005, 26(4):479 –503 diseases include familial CJD, Gertstmann-Sträussler Scheinker syndrome, and fatal familial insomnia. For most PrP mutations, all positive individuals eventually develop the disease. In addition to point mutations in the PrP gene, insertions in the octapeptide repeat coding region of the PrP have been associated with degenerative brain disease (215). 5. BSE. In the past decade, the BSE epidemic in the United Kingdom has brought international attention to the TSE family of diseases. Feeding of protein supplements contaminated with tissues of BSE-positive cattle spread BSE. This encephalopathy has also been transmitted to other species by feeding of contaminated meat and bone meal to ungulates and large felines in zoos and possibly to domestic cats. Transmission to humans has also been suggested by the appearance of vCJD in humans primarily in the United Kingdom. At the molecular level, the interactions between BSE (bovine) PrPsc, and human PrP are rather inefficient (222). E. Cataracts Crystallins are the major structural proteins in the lens and form complex protein-protein interactions with each other. ␣-Crystallins act as molecular chaperones to prevent the aggregation and precipitation of the crystallins by forming large assemblies predicted to contain as many as 33 subunits (223). -Crystallins contain aggregates ranging in size from dimers to octomers, whereas ␥-crystallins exist as monomers. ␣B-Crystallin is a major lens protein that has a structural role in maintaining the transparency of the lens and confers protection to cells against thermal, osmotic, and oxidative damages (224, 225). This protein also functions as a regulator of apoptosis in fibroblast L-929 cells (226). The nonlenticular function of ␣B-crystallin may be related to its chaperone-like property, which allows it to bind irreversibly to denaturing proteins, preventing the formation of large light-scattering aggregates (227). In the lens, this function is highly important in preventing cataracts, as proteins are subject to osmotic and oxidative stresses over the lifetime of the organism and are very susceptible to denaturation and aggregation (228). Interestingly, ␣B-crystallin is overexpressed in a number of disorders such as AD, diffuse LB disease, and in CJD (229 –231). In many of these diseases, ␣B-crystallin is found to be associated with intermediate filaments in cytoplasmic inclusion bodies (232), but the role that ␣B-crystallin may play in these diseases in currently unknown. A missense mutation in ␣B-crystallin, R120G, has been shown to cosegregate with a desmin-related myopathy in a French family (233). The R120G mutation in ␣B-crystallin significantly altered the secondary, tertiary, as well as quaternary structure of the ␣B-crystallin molecule. The mutation of R120G in ␣B-crystallin may lead to imperfect interactions with proteins such as demin in their native state, which results in aggregations of desmin with R120G ␣B-crystallin (234). F. ␣1-AT deficiency ␣1-AT is a member of the serpentin family of protease inhibitors and is secreted in liver cells to control the proteolytic activity of a variety of circulating enzymes. It is the most Castro-Fernández et al. • Protein Routing in Health and Disease abundant serine protease inhibitor in plasma (110). In homozygous protease inhibitor phenotype ZZ (PIZZ) ␣1-AT deficiency, a misfolded but functionally active mutant ␣1ATZ molecule is characterized by a single nucleotide substitution, which results in substitution of lysine for glutamine 342. This abnormally folded mutant protein is polymerized and retained in the ER of hepatocytes rather than secreted into the circulation, causing a reduction in plasma concentrations of ␣1-AT to 80 –90% of normal concentrations (143, 235). The loop where the amino acid substitution takes place in the ␣1-ATZ molecule folds in a manner that leads to formation of large aggregates (110). The mutant protein retains approximately 80% of the wt functional activity, causing a decrease in number of ␣1-AT molecules in the lung, therefore allowing neutrophil elastase to degrade lung parenchyma and increasing the risk of deficient individuals of developing emphysema. In a subset of patients, homozygotes are prone to develop cirrhosis and hepatoma. In children, there is a hepatotoxic effect in liver cells because of the mutant ␣1-AT molecule retained in the ER that originates chronic liver disease (236, 237). There are still doubts regarding which degradation pathway is followed by this ␣1-ATZ mutant form in the ER. One of these pathways could be by binding of ␣1-ATZ to calnexin, ubiquinylation of the cytoplasmic tail of complexed calnexin, and degradation of the ␣1-ATZ-polyubiquitinated calnexin complex by the proteasome. However, it has been suggested that in hepatocytes there might be degradation by a nonlysosomal proteolytic system. Partial correction of the secretory defect of ␣1-AT, the functional activity of which is retained by the mutant ␣1ATZ, is needed to prevent liver and lung disease in ␣1-AT deficiency. Relatively small increases in plasma ␣1-AT levels are needed to inhibit destructive proteolysis in deficient individuals. Compounds that enhance secretion of ␣1-AT can be of great use in patients with this deficiency (237). V. Rescue of Defective Proteins A. Temperature The defect in protein maturation and ER exit of the ⌬F508 CFTR mutant appears to be temperature sensitive. It was observed that a small population of ⌬F508 CFTR was transported to the plasma membrane and was able to confer Cl⫺ transport by lowering the temperature to 30 C or below (112). Lowering the temperature to 27 C is associated with a decrease in degradation of ␣1-ATZ without any changes in secretion; in contrast, elevation of the temperature to 42 C results in decreased degradation and increased secretion, preventing the retention of ␣1-ATZ in the ER by correcting the secretory defect in ␣1-AT deficiency (235). B. Chemical chaperones Genetic rescue of the E90K GnRHR mutant-defective receptor was affected by deleting amino acid K191, a deletion that enhances expression of the wt human receptor at the plasma membrane. Membrane targeting and function of the E90K mutant were rescued as assessed by ligand binding and Castro-Fernández et al. • Protein Routing in Health and Disease by measurement of coupling to an effector system (72). These observations first presented the possibility that E90K was a misrouted receptor, but otherwise competent, and first suggested the possibility that chemical chaperones could be useful in rescuing mutants of the GnRHR. The peptidomimetic, cell membrane-permeant antagonist of GnRH, IN3, rescues most of the misfolded hGnRHR mutants and allows them to escape the QC machinery that degrades them. IN3 serves as a nucleus for proper folding of the mutants and rescues the mutant receptor from a degradation pathway. When mutants were coexpressed with wt hGnRHR, the wt was retained in the ER, leading to a dominant-negative effect by the mutants. However, there were three mutants (E90K, C200Y, or L266R), that were functionally recovered when complexed with the wt receptor, suggesting that the pharmacological agent could rescue both mutants and the wt GnRHR retained in the ER (75). Additional membranepermeant, chemically distinct, GnRHR peptidomimetic agents that allow properly folding and routing to the plasma membrane of GnRHR mutants have been described. However, the mutants S168R and S217R could not be rescued by any drug tested. This suggested that these were potentially irreversibly misfolded or that loss of function resulted from the inability to bind ligand (69). It has been demonstrated that the RP P23H mutant can be induced to properly fold in the presence of the pharmacological chaperone that is a seven-membered ring variant of 11-cis-retinal, 11-cis-7-ring retinal, forming significant amounts of pigment. The rescued 7-P23H-Rh pattern is diffuse but has a greater localization at the cell surface, in contrast to the mutated form located intracellularly, demonstrating that the rescued P23H-Rh was correctly glycosylated and formed a pigment, and was correctly routed to the cell membranes. The 11-cis-7-ring retinal leads to increased amounts of the correctly folded mutant (80, 81). The small peptidomimetic V2R antagonists, SR121463A and VPA-985, have been shown to rescue V2R mutants because they permeate the cell and bind to incomplete folded mutant receptors, thus acting as pharmacological chaperones promoting the proper folding and maturation of receptors and their targeting to the cell surface (238). A class of small synthetic compounds stabilizes p53 core domain. Rescue of locally distorted mutants, such as R249S, could be effected by using the small molecule, FL-CDB3, which binds preferentially to the native state and shifts the conformational equilibrium toward that of the wt in a chaperone mechanism (239). Another small compound, the ellipticine alkaloid, might be acting by directly binding to mutant p53 and stabilizing its DNA-binding conformation. It seems that this compound provokes a modification of the p53 mutants toward the wt by new protein synthesis, and it might be rescuing the folding of only the newly synthesized p53. The ellipticine could be inducing the correct folding of p53 mutants with the assistance of molecular chaperones (104). The adenosine receptor antagonist 8-cyclopentyl-1,3dipropylxanthine as well as N-acetyl-l-cysteine have been shown to activate the Cl⫺ movement in cells expressing the ⌬F508 CFTR protein; however, the mode of action of these Endocrine Reviews, June 2005, 26(4):479 –503 495 compounds remains unclear. 8-Cyclopentyl-1,3-dipropylxanthine seems to have an action only on the mutated CFTR protein, because it does not further increase Cl⫺ efflux of the wt protein. In contrast, N-acetyl-l-cysteine can exert its effect on the wt or the ⌬F508 form of CFTR. Furthermore, cells treated with butyrate show an increase in the overall expression of CFTR, and low levels of the ⌬F508 CFTR protein appear to reach the plasma membrane. This effect could be due either by allowing the mutant protein to escape from the ER and move to the plasma membrane, or by directly influencing the folding of the mutant protein and allowing it to be properly processed (112). Therapeutic strategies that target amyloid formation have been proposed. The overstabilization concept might be combined with simultaneous destabilization of the pathological -sheet conformation of misfolded proteins. -Sheet breaking peptides have been designed to block the conformational changes and aggregation of both A and PrP (240, 241). These peptides can inhibit and reverse protein misfolding and aggregation and might be able to be used both prophylactically and therapeutically. 4⬘-iodo-4⬘-Deoxydoxorubicin and tetracycline have also been reported to inhibit protein misfolding and to dissolve preformed aggregates of several proteins, showing that the reversal of protein misfolding and aggregation is a possible target (242, 243). Peptidomimetics, based on the peptide LVFFA from A, block both A seeding and growth. Unfortunately, the pharmacological profile of these compounds is far from ideal (244). These compounds are typically inhibiting the aggregation of A and several other amyloidogenic polypeptides. Their lack of specificity might be a problem because it has been shown that mammalian melanocytes and related cell types make amyloid aggregation for functional purposes (245). Decreasing the concentration of A below the critical concentration also prevents aggregation (246). Several -secretase inhibitors are now available but none are therapeutically useful. Some ␥-secretase inhibitors are certain nonsteroidal antiinflammatory drugs such as ibuprofen and indomethacin that shift ␥-secretase activity toward the production of shorter less amyloidogenic A forms such as A(1–38). However, the actual mechanism by which these compounds partially inhibit ␥-secretase is unknown (159, 247, 248). When APP transgenic mice overexpressing A are actively immunized with A peptides, deposits that would ordinarily accumulate are cleared (249). Passive immunotherapy with iv A antibody also prevents the formation of A deposits in animals that would otherwise be deposit laden. Unfortunately, active immunization of patients with AD using A(1– 42) revealed antibodies crossing the blood-brain barrier causing meningoencephalitis-like inflammation in 4% of the trial participants, presumably originating from microglial activation (250). This event resulted in the discontinuation of the clinical trial. Monoclonal antibodies that specifically recognize the elongated polyglutamine stretch in the Huntington protein suppressed the in vitro aggregation of this protein (251). It has also been observed that some testified aggregation inhibitors such as Congo red, thioflavine S, chrysamine G, and Direct fast yellow inhibits Huntington protein aggregation; however, some other A aggregation inhibitors such as melato- 496 Endocrine Reviews, June 2005, 26(4):479 –503 nin and rifampicin were not effective at inhibiting this protein aggregation, indicating some protein selectivity. Huntington protein is cleaved by caspases with the liberation of toxic fragments (252). Caspase inhibitors have been shown to inhibit the proteolysis of the Huntington protein and to decrease its toxicity (253). A recent study in a transgenic mouse model of HD indicated that the caspase inhibitor minocycline slows disease progression (254). The efficacy of caspase inhibitors in humans has yet to be determined. A variety of compounds, including anthracyclines, porphyrins, and diazo dyes, block prion replication when administered with PrPsc into animal models (255, 256). Unfortunately, this is not a clinically relevant model for therapeutic intervention, because subclinical disease exists for months in mice and years in humans. Because quinacrine is an approved pharmaceutical, clinical trials are being carried out to test the usefulness of this molecule in patients with CJD. Medicinal chemistry efforts are underway to identify more potent agents. For example, some bisacridines are 10- to 20-fold more potent than quinacrine. Recently, Zahn and colleagues (257) have used nuclear magnetic resonance spectroscopy and chemical-shift tritiation to show that quinacrine binds specifically to PrPc. Antibodies that bind to the PrPc and either stabilize the normal protein conformation or inhibit aggregation were shown to clear prions in cultured mouse neurons, indicating that aggregation inhibitors have potential in the treatment of prion disease (258). Another potential therapeutic strategy for the treatment of amyloidogenic diseases is the inhibition of associated oxidative stress. A variety of antioxidants have shown clinical benefit especially for the treatment of AD. Melatonin and indole-3-propionic acid are potent antioxidants that appear to both inhibit amyloid formation and decrease oxidative stress (259). Vitamins A, E, C, and -carotene all appear to have some potential clinical utility for the treatment of AD (260). Vitamin E has been shown to prevent A-induced oxidative damage in cell culture and delays memory deficits in animal models (261). Although there has been less success in using antioxidant for the treatment of neurodegenerative diseases other than AD, there is potential therapeutic utility in this strategy. If oxidative stress is central to amyloidinduced neurodegeneration, then antioxidants should be a benefit to multiple diseases. Coenzyme Q10, a potent antioxidant and mitochondrial support agent, has shown some clinical benefit in HD (262). The chemical chaperone 4-phenylbutyric acid mediates increases in secretion of ␣1-ATZ that are retained in the ER without affecting its degradation, although this compound does not bind to ␣1-AT specifically (235). Similarly, the glucosidase inhibitor castanospermine mediates increased secretion of ␣1-ATZ. These compounds have the potential to prevent lung injury, because they restore the secretory defect that causes liver and lung disease in ␣1-AT deficiency (237, 263). C. Nonspecific agents Incubation of AQP2-transfected mammalian cells with glycerol and related compounds cause a remarkable redistribution of some AQP2 mutants from ER to plasma mem- Castro-Fernández et al. • Protein Routing in Health and Disease brane and endosomes, conferring increased cellular water permeability (83). Second-site suppressor mutants have been identified by genetic screen or site-directed mutagenesis that restores mutant p53 activity by stabilizing the global fold of the p53 core, or by restoring wt stability and recovering the DNA conformation for DNA binding (103). Other compounds such as glycerol and trimethylamine N-oxide can restore the function of mutant p53 by interacting with the molecule and promoting its proper folding by still unknown mechanisms (104). Furthermore, glycerol, deuterated water, and trimethylamine N-oxide were found to partially correct the processing defect associated with the ⌬F508 CFTR protein. Dimethylsulfoxide can restore cAMP-dependent Cl⫺ transport to the cells, by probably allowing the correct maturation of ⌬F508 CFTR. These compounds might be influencing a population of ⌬F508 CFTR protein to adopt its properly folded state (112). Glycerol mediates increases in secretion of ␣1-ATZ without affecting the degradation, possibly up-regulating the chaperone system. This is accomplished without influencing the secretion efficiency of other proteins or by decreasing proteasome activity (235, 263). VI. Conclusions The ER is the first intracellular organelle in which newly synthesized proteins begin the folding process that leads to formation of mature proteins; such mature proteins properly route within the cell and become functional as receptors, hormones, enzymes, ion channels, and structural components. Retention and QC systems are also found in other compartments of the early secretory pathway, including the cis-Golgi; these pathways serve as a back-up QC mechanism for misfolded proteins that were not retained in the ER. Surprisingly, given the rigor of this extensive network of QC mechanisms along the secretory pathway, incompletely folded proteins are sometimes transported to the plasma membrane or even secreted. It is not yet clear whether these proteins have special characteristics that allow them to escape the QC apparatus, or whether their (apparently undetected) transport results from saturation of the QC system. Proteins that first enter the ER are exposed to chaperones; such molecules stabilize the unfolded or partially folded intermediates, retrotranslocating them to the cytosol for proteasomal degradation. Degradation of properly folded proteins is avoided because the hydrophobic regions of these proteins, which are recognized by the regulatory components of the degradative machinery, are generally buried within folded proteins and because the proteolytic sites themselves are sequestered within internal chambers that are not directly accessible to proteins in the surrounding medium. The stability of the proteins that are processed in the ER is enhanced by disulfide-bond formation, N-glycosylation, and other posttranslational modifications. If the protein folding and maturation process fails, improperly folded molecules are either retained/degraded/corrected in the ER or may lead to buildup of toxic products. Particular diseases can Castro-Fernández et al. • Protein Routing in Health and Disease be explained in terms of buildup of toxins, loss of function of (misrouted) proteins, and defects in the QC/chaperone machinery itself. Understanding the underlying basis of folding and the QC system is proving useful for the design of new therapeutic interventions in human and animal disease. This approach has already led to the use of pharmacological chaperones, small molecules, and temperature-shift techniques to rescue proteins that are potentially functional but are retained intracellularly or by forming aggregates. Chemical chaperones and ligand-induced transport presents options for designing specific drugs to control protein misfolding or transport. Although most studies have been done in vitro, understanding this process and the expectation of increased access to small molecule libraries (promised by the NIH Roadmap plan), presents the reasonable expectation of agents that will be active in vivo and thereby of considerable therapeutic utility. A companion review that focuses more narrowly on rescue of disease-associated mutants of the GnRH receptor is available (264). Endocrine Reviews, June 2005, 26(4):479 –503 497 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. Acknowledgments We appreciate the title suggestion from Dr. Don Pfaff of the Rockefeller University. We thank Eliot Spindel, from the Oregon National Primate Research Center, OHSU, for commenting on a draft of the manuscript. We also thank Shaun Brothers for reading and commenting on drafts of the manuscript. Address all correspondence and requests for reprints to: P. Michael Conn, Oregon National Primate Research Center/Oregon Health and Science University, 505 NW 185th Avenue, Beaverton, Oregon 97006. E-mail: [email protected] This work was supported by National Institutes of Health Grants HD-19899, RR-00163, HD-18185, and TW/HD-00668. 22. 23. 24. 25. 26. 27. References 1. Dobson CM 2003 Protein folding and misfolding. Nature 426:884 – 890 2. Kopito RR 2000 Aggregosomes, inclusion bodies and protein aggregation. Trends Cell Biol 10:524 –530 3. Kostova Z, Wolf DH 2003 For whom the bell tolls: protein quality control of the endoplasmic reticulum and the ubiquitin-proteasome connection. EMBO J 22:2309 –2317 4. Haigh NG, Johnson AE 2002 Protein sorting at the membrane of endoplasmic reticulum. In: Dalbey RE, Von Heijne G, eds. Protein targeting, transport and translocation. London: Academic Press; 74 –106 5. Klausner RD, Sitia R 1990 Protein degradation in the endoplasmic reticulum. Cell 62:611– 614 6. Chevet E, Cameron PH, Pelletier MF, Thomas DY, Bergeron JJM 2001 The endoplasmic reticulum: integration of protein folding, quality control, signaling and degradation. Curr Opin Struct Biol 11:120 –124 7. Ellgaard L, Helenius A 2003 Quality control in the endoplasmic reticulum. Nat Rev 4:181–191 8. Sitia R, Braakman I 2003 Quality control in the endoplasmic reticulum protein factory. Nature 426:891– 894 9. Fra AM, Fagioli C, Finazzi D, Sitia R, Alberini CM 1993 Quality control of ER synthesized proteins: an exposed thiol group as a three-way switch mediating assembly, retention and degradation. EMBO J 12:4755– 4761 10. Helenius A, Trombetta ES, Hebert DN, Simons JF 1997 Calnexin, 28. 29. 30. 31. 32. 33. 34. 35. 36. calreticulin and the folding of glycoproteins. Trends Cell Biol 7:193– 200 Hellman R, Vanhove M, Lejeune A, Stevens FJ, Hedershot LM 1999 The in vivo association of BiP with newly synthesized proteins is dependent on the rate and stability of folding and not simply on the presence of sequences that can bind to BiP. J Cell Biol 144:21–30 Ellgaard L, Molinari M, Helenius A 1999 Setting the standards: quality control in the secretory pathway. Science 286:1882–1888 Oda Y, Hosokawa N, Wada I, Nagata K 2003 EDEM as an acceptor of terminally misfolded glycoproteins released from calnexin. Science 299:1394 –1397 Molinari M, Calanca V, Galli C, Lucca P, Paganetti P 2003 Role of EDEM in the release of misfolded glycoproteins from the calnexin cycle. Science 299:1397–1400 Hammond C, Helenius A 1995 Quality control in the secretory pathway. Curr Opin Cell Biol 7:523–529 Cabral CM, Liu Y, Sifers RN 2001 Dissecting glycoprotein quality control in the secretory pathway. Trends Biochem Sci 26:619 – 624 Bonifacino JS, Klausner RD 1994 Degradation of proteins retained in the endoplasmic reticulum. Modern Cell Biol 15:137–159 Fewell SW, Travers KJ, Weissman JS, Brodsky JL 2001 The action of molecular chaperones in the early secretory pathway. Annu Rev Genet 35:149 –191 Coux O, Tanaka K, Goldenberg AL 1996 Structure and functions of the 20S and 26S proteasomes. Annu Rev Biochem 65:801– 847 Hershko A, Ciechanover A 1998 The ubiquitin system. Annu Rev Biochem 67:425– 479 Sherman MY, Goldberg AL 2001 Cellular defenses against unfolded proteins: a cell biologist thinks about neurodegenerative diseases. Neuron 29:15–32 Groll M, Bajorek M, Kohler A, Moroder L, Rubin DM, Huber R, Glickman MH, Finley D 2000 A gated channel into the proteasome core particle. Nat Struct Biol 7:1062–1067 Bernaroudj N, Zwickl P, Seemüller E, Baumester W, Goldberg AL 2003 ATP hydrolysis by the proteasome regulatory complex PAN serves multiple functions in protein degradation. Mol Cell 11:69 –78 Wickner S, Maurizi MR, Gottesman S 1999 Posttranslational quality control: folding, refolding, and degrading proteins. Science 286: 1888 –1893 Rothman JE, Wieland FT 1996 Protein sorting by transport vesicles. Science 272:227–234 Schekman R, Orci L 1996 Coat proteins and vesicle budding. Science 271:1526 –1533 Kuehn MJ, Hermann JM, Schekman R 1998 COPII-cargo interactions direct protein sorting into ER-derived transport vesicles. Nature 391:187–190 Taxis C, Vogel F, Wolf DH 2002 ER-Golgi traffic is a prerequisite for efficient ER degradation. Mol Biol Cell 13:1806 –1818 Mezzacasa A, Helenius A 2002 The transitorial ER defines a boundary for quality control in the secretion of tsO45 VSV glycoprotein. Traffic 3:833– 849 Hammond C, Helenius A 1994 Quality control in the secretory pathway: retention of a misfolded viral membrane glycoprotein involves cycling between the ER, intermediate compartment, and Golgi apparatus. J Cell Biol 126:41–52 Arvan P, Zhao X, Ramos-Castaneda J, Chang A 2002 Secretory pathway quality control operating in Golgi, plasmalemmal, and endosomal systems. Traffic 3:771–780 VanSlyke JK, Deschenes SM, Musil LS 2000 Intracellular transport, assembly, and degradation of wild-type and disease-linked mutant gap junction proteins. Mol Biol Cell 11:1933–1946 Caldwell SR, Hill KJ, Cooper AA 2001 Degradation of endoplasmic reticulum (ER) quality control substrates requires transport between the ER and Golgi. J Biol Chem 276:23296 –23303 Vashist S, Kim W, Belden WJ, Spear ED, Barlowe D, Ng DT 2001 Distinct retrieval and retention mechanisms that are required for the quality control of endoplasmic reticulum protein folding. J Cell Biol 155:355–368 Zuber C, Spiro MJ, Guhl B, Spiro RG, Roth J 2000 Golgi apparatus immunolocalization of endomannosidase suggests post-endoplasmic reticulum glucose trimming. Implications for quality control. Mol Biol Cell 11:4227– 4240 Zuber C, Fan JY, Guhl B, Parodi A, Fessler JH, Parker C, Roth J 498 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. Endocrine Reviews, June 2005, 26(4):479 –503 2001 Immunolocalization of UDP-glucose:glycoprotein glucosyltransferase indicates involvement of pre-Golgi intermediates in protein quality control. Proc Natl Acad Sci USA 98:10710 –10715 Roth J, Ziak M, Zuber C 2003 The role of glucosidase II and endomannosidase in glucose trimming of asparagine-linked oligosaccharides. Biochimie 85:287–294 Busca R, Martinez M, Vilella E, Pognonoec P, Deeb S, Auwerx J, Reina M, Vilaro S 1996 The mutation Gly142Glu in human lipoprotein lipase produces a missorted protein that is diverted to lysosomes. J Biol Chem 271:2139 –2146 Millar CA, Meerloo T, Martin S, Hickson GRX, Shimwell NJ, Wakelam MJO, James DE, Gould GW 2000 Adipsin and the glucose transporter GLUT4 traffic to the cell surface via independent pathways in adipocytes. Traffic 1:141–151 Turner M, Arvan P 2000 Protein traffic from the secretory pathway to the endosomal system in pancreatic -cells. J Biol Chem 275: 14025–14030 Harsay F, Schekman R 2002 A subset of yeast vacuolar protein sorting mutants is blocked in one branch of the exocytic pathway. J Cell Biol 156:271–286 Munro S 1995 An investigation of the role of transmembrane domains in Golgi protein retention. EMBO J 14:4695– 4705 Wolins N, Bosshart H, Kuster H, Bonifacino JS 1997 Aggregation as a determinant of protein fate in post-Golgi compartments: role of the luminal domain of furin in lysosomal targeting. J Cell Biol 139:1735–1745 Stock D, Gibbons C, Arechaga I. Leslie AG, Walker JE 2000 The rotary mechanism of ATP synthase. Curr Opin Struct Biol 10:672– 679. Eckert A, Keil U, Marques A, Bonert A, Frey C, Schüssel, Müller WE 2003 Mitochondrial dysfunction, apoptotic cell death, and Alzheimer⬘s disease. Biochem Pharmacol 66:1627–1634 Langer T, Neupert W 1996 Regulated protein degradation in mitochondria. Experientia 52:1069 –1076 Martin J 1997 Molecular chaperones and mitochondrial protein folding. J Bioenerg Biomembr 29:35– 43 Koehler CM 2000 Protein translocation pathways of the mitochondrion. FEBS Lett 476:27–31 Voos W, Rörrgers K 2002 Molecular chaperones as essential mediators of mitochondrial biogenesis. Biochim Biophys Acta 1592: 51– 62 Bross P, Corydon TJ, Andersen BS, Jørgensen MM, Bolund L, Gregersen N 1999 Protein misfolding and degradation in genetic diseases. Hum Mutat 14:186 –198 Bukau B, Horwich AL 1998 The Hsp70 and Hsp60 chaperone machines. Cell 92:351–366 Hartl FU, Hayer-Hartl M 2002 Molecular chaperones in the cytosol: from nascent chain to folded protein. Science 295:1852–1858 Stirling PC, Lundin VF, Leroux MR 2003 Getting a grip on nonnative proteins. EMBO Rep 4:565–570 Parcellier A, Gurbuxani S, Schmitt E, Solary E, Garrido C 2003 Heat shock proteins, cellular chaperones that modulate mitochondrial cell death pathways. Biochem Biophys Res Commun 304: 505–512 Feldman DE, Frydman J 2000 Protein folding in vivo: the importance of molecular chaperones. Curr Opin Struct Biol 10:26 –33 Molinari M, Helenius A 2000 Chaperone selection during glycoprotein translocation into the endoplasmic reticulum. Science 288: 331–333 Frydman J, Hohfeld J 1997 Chaperones get in touch: the Hip-Hop connection. Trends Biochem Sci 22:87–92 Kelley WL 1998 The J-domain family and the recruitment of chaperone power. Trends Biochem Sci 23:222–227 Parsell D, Kowal A, Singer M, Lindquist S 1994 Protein dissaggregation mediate by heat shock protein Hsp104. Nature 372:475– 478 Patino MM, Liu JJ, Gloover JR, Lindquist S 1996 Support of the prion hypothesis for inheritance of a phenotypic trait in yeast. Science 273:622– 626 Glover JR, Lindquist S 1998 Hsp104, Hsp70 and Hsp40: a novel chaperone system that rescues previously aggregated proteins. Cell 94:73– 82 Castro-Fernández et al. • Protein Routing in Health and Disease 62. Parodi AJ 2000 Protein glycosylation and its role in protein folding. Annu Rev Biochem 69:611– 614 63. Ou W, Cameron P, Thomas D, Bergeron J 1993 Association of folding intermediate of glycoproteins with calnexin during protein maturation. Nature 364:771–776 64. McCracken AA, Brodsky JC 1996 Assembly of ER-associated protein degradation in vitro: dependence on cytosol, calnexin, and ATP. J Cell Biol 132:291–298 65. Ware FE, Vassilakos A, Peterson PA, Jackson MR, Lehrman MA, Williams DB 1995 The molecular chaperone calnexin binds Glc1Man9GlcNAc2 oligosaccharide as an initial step in recognizing unfolded glycoproteins. J Biol Chem 270:4697– 4704 66. Liu Y, Choudhury P, Cabral CM, Siters RN 1999 Oligosaccharide modification in the early secretory pathway directs the selection of a misfolded glycoprotein for degradation by the proteasome. J Biol Chem 274:5861–5867 67. Molinari M, Helenius A 1999 Glycoproteins from mixed disulfides with oxidoreductases during folding in living cells. Nature 402: 90 –93 68. Janovick JA, Maya-Núñez G, Conn PM 2002 Rescue of hypogonadotropic hypogonadism-causing and manufactured GnRH receptor mutants by a specific protein-folding template: misrouted proteins as a novel disease etiology and therapeutic target. J Clin Endocrinol Metab 87:3255–3262 69. Janovick JA, Goulet M, Bush E, Greer J, Wettlaufer DG, Conn PM 2003 Structure-activity relations of successful pharmacologic chaperones for rescue of naturally occurring and manufactured mutants of the gonadotropin-releasing hormone receptor. J Pharmacol Exp Ther 305:608 – 614 70. Tan CM, Brady AE, Nickols HH, Wang Q, Limbird LE 2004 Membrane trafficking of G protein-coupled receptors. Annu Rev Pharmacol Toxicol 44:559 – 609 71. Conn PM, Leanos-Miranda A, Janovick JA 2002 Protein origami: therapeutic rescue of misfolded gene products. Mol Interv 2:308 – 316 72. Maya-Núñez G, Janovick JA, Ulloa-Aguirre A, Söderlund D, Conn PM, Méndez JP 2002 Molecular basis of hypogonadotropic hypogonadism: restoration of mutant (E90K) GnRH receptor function by a deletion at a distant site. J Clin Endocrinol Metab 87: 2144 –2149 73. Leaños-Miranda A, Janovick JA, Conn PM 2002 Receptormisrouting: an unexpectedly prevalent and rescuable etiology in GnRHR-mediated hypogonadotropic hypogonadism. J Clin Endocrinol Metab 87:4825– 4828 74. Ulloa-Aguirre A, Janovick JA, Leaños-Miranda A, Conn PM 2003 Misrouted cell surface receptors as a novel disease aetiology and potential therapeutic target: the case of hypogonadotropic hypogonadism due to gonadotropin-releasing hormone resistance. Expert Opin Ther Targets 7:175–185 75. Leaños-Miranda A, Ulloa-Aguirre A, Ji TH, Janovick JA, Conn PM 2003 Dominant-negative action of disease-causing gonadotropin-releasing hormone receptor (GnRHR) mutants: a trait that potentially coevolved with decreased plasma membrane expression of GnRHR in humans. J Clin Endocrinol Metab 88:3360 –3367 76. Brothers SP, Cornea A, Janovick JA, Conn PM 2004 Human ‘lossof-function’ GnRH receptor mutants retain wild type receptor in the endoplasmic reticulum: molecular basis of the dominantnegative effect. Mol Endocrinol 18:1787–1797 77. Janovick, J, Ulloa-Aguirre A, Conn PM 2003 Evolved regulation of gonadotropin releasing hormone receptor plasma membrane expression. Endocrine 22:317–328 78. Söderlund D, Canto P, De la Chesnaye E, Ulloa-Aguirre A, Méndez JP 2001 A novel homozygous mutation in the second transmembrane domain of the gonadotropin releasing hormone receptor gene. Clin Endocrinol (Oxf) 54:493– 498 79. Brothers SP, Janovick JA, Conn PM 2003 Unexpected effects of epitope and chimeric tags on gonadotropin-releasing hormone receptors: implications for understanding the molecular etiology of hypogonadotropic hypogonadism. J Clin Endocrinol Metab 88: 6107– 6112 80. Noorwez SM, Kuksa V, Imanishi Y, Zhu L, Filipek S, Palczewski K, Kaushal S 2003 Pharmacological chaperone-mediated in vivo folding and stabilization of the P23H-opsin mutant associated with Castro-Fernández et al. • Protein Routing in Health and Disease 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. autosomal dominant retinitis pigmentosa. J Biol Chem 278:14442– 14450 Noorwez SM, Malhotra R, McDowell JH, Smith KA, Krebs MP, Kaushal S 2004 Retinoids assist the cellular folding of the autosomal dominant retinitis pigmentosa opsin mutant P23H. J Biol Chem 279:16278 –16284 Colley NJ, Cassill JA, Baker EK, Zuker CS 1995 Defective intracellular transport is the molecular basis of rhodopsin-dependent dominant retinal degeneration. Proc Natl Acad Sci USA 92:3070 – 3074 Tamarappoo BK, Verkman AS 1998 Defective aquaporin-2 trafficking in nephrogenic diabetes insipidus and correction by chemical chaperones. J Clin Invest 101:2257–2267 Kamsteeg EJ, Wormhoudt TAM, Rijss JPL, van Os CH, Deen PMT 1999 An impaired routing of wild-type aquaporin-2 after tetramerization with an aquaporin-2 mutant explains dominant nephrogenic diabetes insipidus. EMBO J 18:2394 –2400 Kamsteeg EJ, Bichet DG, Konings IBM, Nivet H, Lonergan M, Arthus MF, van Os CH, Deen PMT 2003 Reversed polarized delivery of an aquaporin-2 mutant causes dominant nephrogenic diabetes insipidus. J Cell Biol 163:1099 –1109 Morello JP, Bichet DG 2001 Nephrogenic diabetes insipidus. Annu Rev Physiol 63:607– 630 Hirano K, Zuber C, Roth J, Ziak M 2003 The proteasome is involved in the degradation of different aquaporin-2 mutants causing nephrogenic diabetes insipidus. Am J Pathol 163:111–120 Petäjä-Repo UE, Hogue M, Laperrière A, Walker P, Bouvier M 2000 Export from the endoplasmic reticulum represents the limiting step in the maturation and cell surface expression of the human ␦ opioid receptor. J Biol Chem 275:13727–13736 Bai M 2004 Dimerization of G-protein-coupled receptors: roles in signal transduction. Cell Signal 16:175–186 Conn PM, Rogers DC, Stewart JM, Niedel J, Sheffield T 1982 Conversion of a gonadotropin-releasing hormone antagonist to an agonist. Nature 296:653– 655 Robbins MJ, Ciruela F, Rhodes A, McIllhinney RA 1999 Characterization of the dimerization of metabotropic glutamate receptors using an N-terminal truncation of mGluR1␣. J Neurochem 72:2539 –2547 Terrillon S, Durroux T, Mouillac B, Breit A, Ayoub MA, Taulan M, Jockers R, Barberis C, Bouvier M 2003 Oxytocin and vasopressin V1a and V2 receptors form constitutive homo- and heterodimers during biosynthesis. Mol Endocrinol 17:677– 691 Rodriguez-Frade JM, Vila-Coro AJ, de Ana AM, Albar JP, Martinez AC, Mellado M 1999 The chemokine monocyte chemoattractant protein-1 induces functional responses through dimerization of its receptor CCR2. Proc Natl Acad Sci USA 96:3628 –3633 Vila-Coro AJ, Mellado M, Martin de Ana A, Lucas P, del Real G, Marinez AC, Rodriguez-Frade JM 2000 HIV-1 infection through the CCR5 receptor is blocked by receptor dimerization. Proc Natl Acad Sci USA 97:3388 –3393 Cornea A, Janovick JA, Maya-Núñez G, Conn PM 2001 Gonadotropin-releasing hormone receptor microaggregation. Rate monitored by fluorescence resonance energy transfer. J Biol Chem 276: 2153–2158 Zhu X, Wess J 1998 Truncated V2 vasopressin receptors as negative regulators of wild-type V2 receptor function. Biochemistry 37:15773–15784 Lee SP, O’Dowd BF, Ng GY, Varghese G, Akil H, Mansour A, Nguyen T, George SR 2000 Inhibition of cell surface expression by mutant receptors demonstrates that D2 dopamine receptors exist as oligomers in the cell. Mol Pharmacol 58:120 –128 Kaykas A, Yang-Snyder J, Heroux M, Shah KV, Bouvier M, Moon RT 2004 Mutant Frizzled 4 associated with vitreoretinopathy traps wild-type Frizzled in the endoplasmic reticulum by oligomerization. Nat Cell Biol 6:52–58 Jones KA, Borowsky B, Tamm JA, Craig DA, Durkin M, Dai M, Yao WJ, Johnson M, Gunwaldsen C, Huang LY, Tang C, Shen O, Salon JA, Morse K, Laz T, Smith KE, Nagarathnam D, Noble SA, Branchek TA, Gerald C 1998 GABA(B) receptors function as a heteromeric assembly of the subunits GABA(B)R1 and GABA(B)R2. Nature 396:674 – 679 Galvez T, Duthery B, Kniazeff J, Blahos J, Rovelli G, Bettler B, Endocrine Reviews, June 2005, 26(4):479 –503 499 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. Prezeau L, Pin JP 2001 Allosteric interactions between GB1 and GB2 subunits are required for optimal GABA(B) receptor function. EMBO J 202152–2159 Pagano A, Rovelli G, Mosbacher J, Lohmann T, Duthey B, Stauffer D, Ristig D, Schuler V, Meigel I, Lampert C, Stein T, Prezeau L, Blahos J, Pin J, Froestl W, Kuhn R, Heid J, Kaupmann K, Bettler B 2001 C-terminal interaction is essential for surface trafficking but not for heteromeric assembly of GABA(b) receptors. J Neurosci 21:1189 –1202 Meulmeester E, Frenk R, Stad R, de Graaf P, Marine JC, Vousden KH, Jochemsen AG 2003 Critical role for a central part of Mdm2 in the ubiquitylation of p53. Mol Cell Biol 23:4929 – 4938 Nikolova PV, Wong KB, DeDecker B, Henckel J, Fersht AR 2000 Mechanism of rescue of common p53 cancer mutations by secondsite suppressor mutations. EMBO J 19:370 –378 Peng Y, Li C, Chen L, Sebti S, Chen J 2003 Rescue of mutant p53 transcription function by ellipticine. Oncogene 22:4478 – 4487 Wong KB, DeDecker BS, Freund SM, Proctor MR, Bycroft M, Fersht AR 1999 Hot-spot mutants of p53 core domain evince characteristic local structural changes. Proc Natl Acad Sci USA 90: 8586 – 8590 Bykov VJN, Issaeva N, Shilov A, Hultcrantz M, Pugacheva E, Chumakov P, Bergman J, Wiman KG, Selivanova G 2002 Restoration of the tumor suppressor function to mutant p53 by a lowmolecular weight compound. Nat Med 8:282–288 Pfeiffer R, Loffing J, Rossier G, Bauch C, Meier C, Eggermann T, Loffing-Cueni D, Kühn LC, Verrey F 1999 Luminal heterodimeric amino acid transporter defective in cystinuria. Mol Biol Cell 10: 4135– 4147 Chillarón J, Estévez R, Samarzija I, Waldegger S, Testar X, Lang F, Zorzano A, Busch A, Palacı́n M 1997 An intracellular trafficking defect in type I cystinuria rBAT mutants M467T and M467K. J Biol Chem 272:9543–9549 Pineda M, Wagner CA, Bröer A, Stehberger PA, Kaltenbach S, Gelpı́ JLl, Martı́n del Rı́o R, Zorzano A, Palacı́n M, Lang F, Bröer S 2004 Cystinuria-specific rBAT (R365W) mutation reveals two translocation pathways in the amino acid transporter rBAT-b0,⫹AT. Biochem J 377:665– 674 Rutishauser J, Spiess M 2002 Endoplasmic reticulum storage diseases. Swiss Med Wkly 132:211–222 Springsteel MF, Galietta LJV, Ma T, By K, Berger GO, Yang H, Dicus CW, Choung W, Quan C, Shelat AA, Guy RK, Verkman AS, Kurth MJ, Nantz MH 2003 Benzaflavone activators of the cystic fibrosis transmembrane conductance regulator: towards a pharmacophore model for the nucleotide-binding domain. Bioor Med Chem 11:4113– 4120 Randell Brown C, Hong-Brown LQ, Welch WJ 1997 Strategies for correcting the ⌬F508 CFTR protein-folding defect. J Bioenerg Biomembr 29:491–502 Gilbert A, Jadot M, Leontieva E, Wattiaux-De Coninck S, Wattiaux R 1998 DF508 CFTR localizes in the endoplasmic reticulumGolgi intermediate compartment in cystic fibrosis cells. Exp Cell Res 242:144 –152 Qu BH, Strickland E, Thomas PJ 1997 Cystic fibrosis: a disease of altered protein folding. J Bioenerg Biomembr 29:483– 490 Okiyoneda T, Harada K, Takeya M, Yamahira K, Wada I, Shuto T, Suico MA, Hashimoto I, Kai H 2004 ⌬F508 CFTR pool in the endoplasmic reticulum is increased by calnexin overexpression. Mol Biol Cell 15:563–574 Zeitllin PL 1999 Novel pharmacologic therapies for cystic fibrosis. J Clin Invest 103:447– 452 Gitelman HJ, Graham JB, Welt LG 1966 A new familial disorder characterized by hypokalemia and hypomagnesemia. Trans Assoc Am Physicians 79:221–235 Bettinelli A, Vezzoli G, Colussi G, Bianchetti MG, Sereni F, Casari G 1998 Genotype-phenotype correlations in normotensive patients with primary renal tubular hypokalemic metabolic alkalosis. J Nephrol 11:61– 69 Cruz DN, Shaer AJ, Bia MJ, Lifton RP, Simon DB 2001 Gitelman’s syndrome revisited: an evaluation of symptoms and health-related quality of life. Kidney Int 59:710 –717 Scheinman SJ, Guay-Woodford LM, Thakker RV, Warnock DG 500 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. 133. 134. 135. 136. 137. Endocrine Reviews, June 2005, 26(4):479 –503 1999 Genetic disorder of renal electrolyte transport. N Engl J Med 340:1177–1187 De Jong JC, Van Der Vliet WA, Van Den Heuvel LPWJ, Willems PHGM, Knoers NVAM, Bindels RJM 2002 Functional expression of mutations in the human NaCl cotransporter: evidence for impaired routing mechanisms in Gitelman’s syndrome. J Am Soc Nephrol 13:1442–1448 Gamba G, Miyanoshita A, Lombardi M, Lytton J, Lee WS, Hediger MA, Hebert SC 1994 Molecular cloning, primary structure, and characterization of two members of the mammalian electroneutral sodium-(potassium)-chloride cotransporter family expressed in kidney. J Biol Chem 269:17713–17722 Velazquez H, Náray-Fejes-Tóth A, Silva T, Andújar E, Reilly RF, Desir V, Ellison DH 1998 The distal convoluted tubule of the rabbit coexpresses NaCl cotransporter and 11-hydroxysteroid dehydrogenase. Kidney Int 54:464 – 472 Wyse BD, Ali N, Ellison DH 2002 Interaction with grp58 increases activity of the thiazide-sensitive Na-Cl cotransporter. Am J Physiol 282:F424 –F430 Giannetti AM, Björkman PJ 2004 HFE and transferrin directly compete for transferrin receptor in solution and at the cell surface. J Biol Chem 279:25866 –25875 Waheed A, Parkkila S, Zhou XY, Tomatsu S, Tsuchihashi Z, Feder JN, Schatzman RC, Britton RS, Bacon BR, Sly WS 1997 Hereditary hemochromatosis: effects of C282Y and H63D mutations on association with 2-microglobulin, intracellular processing, and cell surface expression of the HFE protein in COS-7 cells. Proc Natl Acad Sci USA 94:12384 –12389 Feder JN, Gnirke A, Thomas W, Tsuchihashi Z, Ruddy DA, Basava A, Dormishian F, Domingo R, Ellis ML, Fullan A, Hinton LM, Jones NL, Kimmel BE, Kronmal GS, Lauer P, Lee VK, Loeb DB, Mapa FA, McClelland EE, Meyer NC, Mintier GA, Moeller NN, Moore T, Morikang E, Prass LE, Quintana L, Starnes SM, Schatzman RC, Brunke KJ, Drayna DT, Risch NJ, Bacon BR, Wolff RK 1996 A novel MHC class I-like gene is mutated in patients with hereditary haemochromatosis. Nat Genet 13:399 – 408 Bodmer JG, Parham P, Albert ED, Marsh SGE 1997 Putting a hold on “HLA-H’. The WHO Nomenclature Committee for Factors of the HLA System. Nat Genet 15:234 –235 Feder JN, Penny DM, Irrinki A, Lee VK, Lebron JA, Watson N, Tsuchihashi Z, Sigal E, Bjorkman PJ, Schatzman RC 1998 The hemochromatosis gene product complexes with the transferrin receptor and lowers its affinity for ligand binding. Proc Natl Acad Sci USA 95:1472–1477 Björkman PJ, Parham P 1990 Structure, function, and diversity of class I major histocompatibility complex molecules. Annu Rev Biochem 59:253–288 Miyazaki J, Appella E, Ozato K 1986 Intracellular transport blockade caused by disruption of the disulfide bridge in the third external domain of major histocompatibility complex class I antigen. Proc Natl Acad Sci USA 83:757–761 Hansen TH, Myers NB, Lee DR 1988 Studies of two antigenic forms of Ld with disparate 2-microglobulin (2m) associations suggest that 2m facilitate the folding of the ␣1 and ␣2 domains during de novo synthesis. J Immunol 140:3522–3527 Lebron JA, West Jr AP, Björkman PJ 1999 The hemochromatosis protein HFE competes with transferrin for binding to the transferrin receptor. J Mol Biol 294:239 –245 Laham N, Rotem-Yehudar R, Shechter C, Coligan JE, Ehrlich R 2004 Transferrins receptor association with endosomal localization of soluble HFE are not sufficient for regulation of cellular iron homeostasis. J Cell Biochem 91:1130 –1145 Williams DB, Watts TH 1995 Molecular chaperones in antigen presentation. Curr Opin Immunol 7:77– 84 Feder JN, Tsuchihashi Z, Irrinki A, Lee VK, Mapa FA, Morikang E, Prass CE, Starnes SM, Wolff RK, Parkkila S, Sly WS, Schatzman RC 1997 The hemochromatosis founder mutation in HLA-H disrupts 2-microglobulin interaction and cell surface expression. J Biol Chem 272:14025–14028 Hsu VW, Yuan LC, Nuchtern JG, Lippincott-Schwartz J, Hammerling GJ, Klausner RD 1991 A recycling pathway between the endoplasmic reticulum and the Golgi apparatus for retention of unassembled MHC class I molecules. Nature 352:441– 444 Castro-Fernández et al. • Protein Routing in Health and Disease 138. Jørgensen MM, Jensen ON, Holst HU, Hansen JJ, Corydon TJ, Bross P, Bolund L, Gregersen N 2000 Grp78 is involved in retention of mutant low density lipoprotein receptor protein in the endoplasmic reticulum. J Biol Chem 275:33861–33868 139. Li Y, Lu W, Schwartz AL, Bu G 2004 Degradation of the lowdensity lipoprotein receptor class 2 mutants is mediated by a proteasome-dependent pathway. J Lipid Res 45:1084 –1091 140. Hobbs HH, Russell DW, Brown MS, Goldstein JL 1990 The LDL receptor locus in familial hypercholesterolemia: mutational analysis of a membrane protein. Annu Rev Genet 24:133–170 141. Li Y, Lu W, Schwartz AL, Bu G 2002 Receptor-associated protein facilitates proper holding and maturation of the low-density lipoprotein receptor and its class 2 mutants. Biochemistry 41:4921– 4928 142. Solomon B 2002 Anti-aggregating antibodies, a new approach towards treatment of conformation diseases. Curr Med Chem 9:1737–1749 143. Carrell RW, Lomas DA 1997 Conformational disease. Lancet 350: 134 –138 144. Kopito RR, Ron D 2000 Conformational disease. Nat Cell Biol 2:E207–E209 145. Martin JB 1999 Molecular basis of the neurodegenerative diseases. N Engl J Med 340:1970 –1980 146. Soto C 2003 Unfolding the role of protein misfolding in neurodegenerative diseases. Nat Rev 4:49 –59 147. Jiménez JL, Guijarro JI, Orlova E, Zurdo J, Dobson CM, Sunde M, Saibil HR 1999 Cryo-electron microscopy structure of an SH3 amyloid fibril and model of the molecular packing. EMBO J 18: 815– 821 148. Stefani M, Dobson CM 2003 Protein aggregation and aggregate toxicity: new insights into protein folding, misfolding diseases and biological evolution. J Mol Med 81:678 – 699 149. Dobson CM 1999 Protein misfolding, evolution and disease. Trends Biochem Sci 24:329 –332 150. Dobson CM 2001 The structural basis of protein folding and its links with human disease. Philos Trans R Soc Lond B Biol Sci 356:133–145 151. Lorenzo A, Yanker BA 1994 -Amyloid neurotoxicity requires fibril formation and its inhibited by Congo red. Proc Natl Acad Sci USA 91:12243–12247 152. Thomas T, Thomas G, McLendon C, Sutton T, Mullan M 1996 -Amyloid-mediated vasoactivity and vascular endothelial damage. Nature 380:168 –171 153. Dickson DW 1995 Correlation of synaptic and pathological markers with cognition of the elderly. Neurobiol Aging 16:285–298 154. Csermely P 2001 Chaperone overload is a possible contributor to “civilization diseases.” Trends Genet 17:701–704 155. Dobson CM 2002 Getting out of shape. Nature 418:729 –730 156. Evans DA, Funkenstein HH, Albert MS, Scherr PA, Cook NR, Chown MJ, Hebert LE, Hennekens CH, Taylor JO 1989 Prevalence of Alzheimer’s disease in a community population of older persons. Higher than previously reported. JAMA 262:2551–2556 157. Price DL, Borchelt DR, Sisodia SS 1993 Alzheimer disease and the prion disorders amyloid -protein and prion protein amyloidoses. Proc Natl Acad Sci USA 90:6381– 6384 158. Shastry BS 2003 Neurodegenerative disorders of protein aggregation. Neurochem Int 43:1–7 159. Thompson AJ, Barrow CJ 2002 Protein conformational misfolding and amyloid formation: characteristics of a new class of disorders that include Alzheimer’s and prion diseases. Curr Med Chem 9:1751–1762 160. Hardy J, Selkoe DJ 2002 The amyloid hypothesis of Alzheimer’s disease: progress and problems on the road to therapeutics. Science 297:353–356 161. Temussi PA, Masino L, Pastore A 2003 From Alzheimer to Huntington: why is a structural understanding so difficult? EMBO J 22:355–361 162. Lee VM, Goedert M, Trojanowski JQ 2001 Neurodegenerative tautopathies. Annu Rev Neurosci 24:1121–1159 163. McLlory SP, Vahidassr AD, Svage DA, Patterson LC, Lawson JT, Passmore AP 1999 Risk of Alzheimer’s disease is associated with a very low density lipoprotein receptor genotype in Northern Ireland. Am J Med Genet 88:140 –144 Castro-Fernández et al. • Protein Routing in Health and Disease 164. Lichtenthaler SF, Beher D, Grim HS, Wang R, Shearman MS, Masters CL, Beyreuther K 2002 The intramembrane cleavage site of the amyloid precursor protein depends on the length of its transmembrane domain. Proc Natl Acad Sci USA 99:1365–1370 165. Castaño EM, Ghiso J, Prelli F, Gorevic PD, Migheli A, Frangione B 1986 In vitro formation of amyloid fibrils from two synthetic peptides of different lengths homologous to Alzheimer’s disease -protein. Biochem Biophys Res Commun 141:782–789 166. Soto C, Castaño EM, Frangione B, Inestrosa NC 1995 The ␣-helical to -strand transition in the amino-terminal fragment of the amyloid -peptide modulates amyloid formation. J Biol Chem 270: 3063–3067 167. Hardy J, Allsop D 1991 Amyloid deposition as the central event in the aetiology of Alzheimer’s disease. Trends Pharmacol Sci 12:383– 388 168. Selkoe DJ 1996 Amyloid -protein and the genetics of Alzheimer’s disease. J Biol Chem 271:18295–18298 169. Hardy J 1997 Amyloid, the presenilins and Alzheimer’s disease. Trends Neurosci 20:154 –159 170. Goate A, Chartier-Harlin M-C, Mullan M, Brown J, Crawford F, Fidani L, Giuffra L, Haynes A, Irving N, James L, Mant R, Newton P, Rooke K, Roques P, Talbot C, Pericak-Vance M, Roses A, Williamson R, Rossor M, Owen M, Hardy J 1991 Segregation of a missense mutation in the amyloid precursor protein gene with familial Alzheimer’s disease. Nature 349:704 –706 171. Hardy J, Duff K, Hardy KG, Perez-Tur J, Hutton M 1998 Genetic dissection of Alzheimer’s disease and related dementias: amyloid and its relationship to . Nat Neurosci 1:355–358 172. Selkoe D 2001 Presenilin, notch and the genesis and treatment of Alzheimer’s disease. Proc Natl Acad Sci USA 98:11039 –11041 173. Canu N, Dus L, Barbato C, Ciotti MT, Brancolini C, Rinaldi AM, Novak M, Cattaneo A, Bradbury A, Calissano P 1998 Cleavage and dephosphrylation in cerebellar granule neurons undergoing apoptosis. J Neurosci 18:7061–7074 174. Von Bergen M, Friedhoff P, Biernat J, Heberle J, Mandelkow EM, Mandelkow E 2000 Assembly of protein into Alzheimer paired helical filaments depends on a local sequence motif ((306)VQIVYK(311)) forming  structure. Proc Natl Acad Sci USA 97:5129 –5134 175. Forloni G, Terreni L, Bertani I, Fogliarino S, Invernizzi R, Assini A, Ribizzi G, Negro A, Calabrese E, Volonté MA, Mariani C, Franceschi M, Tabaton M, Bertoli A 2002 Protein misfolding in Alzheimer’s and Parkinson’s disease: genetics and molecular mechanisms. Neurobiol Aging 23:957–976 176. Shastry BS 2001 Parkinson’s disease: etiology, pathogenesis and future of gene therapy. Neurosci Res 41:5–12 177. Tang GM, Xie XJ, Xu L, Hao YX, Lin DY, Ren DM 2002 Genetic study of apolipoprotein E gene, ␣-1 antichymotrypsin gene in sporadic Parkinson’s disease. Am J Med Genet 114:446 – 449 178. Checkoway H, Powers D, Smith-Weller T, Franklin GM, Longstreth WT, Swanson PD 2002 Parkinson’s disease risks associated with cigarette smoking, alcohol consumption and caffeine intake. Am J Epidemiol 155:738 –752 179. Clayton DF, George JM 1998 The synucleins: a family of proteins involved in synaptic function, plasticity, neurodegeneration and disease. Trends Neurosci 21:249 –254 180. Kahle PJ, Neumann M, Ozmen L, Muller V, Jacobsen H, Schindzielorz A, Okochi M, Leimer U, van Der Putten H, Probst A, Kremmer E, Kretzschmar HA, Haass C 2000 Subcellular localization of wild-type and Parkinson’s disease-associated mutant ␣-synuclein in human and transgenic mouse brain. J Neurosci 20:6365– 6373 181. Iwai A, Masliah E, Yoshimoto M, Ge N, Flanagan L, de Silva HA, Kittel A, Sayito T 1995 The precursor protein of non-A component of Alzheimer’s disease amyloid is a presynaptic protein of the central nervous system. Neuron 14:467– 475 182. Crowther RA, Daniel SE, Goedert M 2000 Characterization of isolated ␣-synuclein filaments from substantia nigra of Parkinson’s disease brain. Neurosci Lett 292:128 –130 183. Spillantini MG, Schidt ML, Lee VM, Trojanowski JQ, Jakes R, Goedert M 1997 ␣-Synuclein in Lewy bodies. Nature 88:839 – 840 184. Spillantini MG, Crowther RA, Jakes R, Hasegawa M, Goedert M 1998 ␣-Synuclein in filamentous inclusions of Lewy bodies from Endocrine Reviews, June 2005, 26(4):479 –503 501 185. 186. 187. 188. 189. 190. 191. 192. 193. 194. 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. 205. Parkinson’s disease and dementia with Lewy bodies. Proc Natl Acad Sci USA 95:6469 – 6473 Baba M, Nakajo S, Tu PH, Tomita T, Nakaya K, Lee VM, Trojanowski JQ, Iwatsubo T 1998 Aggregation of ␣-synuclein in Lewy bodies of sporadic Parkinson’s disease and dementia with Lewy bodies. Am J Pathol 52:879 – 884 Jenco JM, Rawlingson A, Daniels B, Morris AJ 1998 Regulation of phospholipase D2: selective inhibition of mammalian phospholipase D isoenzymes by ␣- and  synucleins. Biochemistry 37:4901– 4909 Colley WC, Sung TC, Roll R, Jenco J, Hammond SM, Altshuller Y, Bar-Sagi D, Morris AJ, Frohman MA 1997 Phospholipase D2, a distinct phospholipase D isoform with novel regulatory properties that provokes cytoskeletal re-organization. Curr Biol 7:191–201 Jensen PH, Hager H, Nielsen MS, Hojrup P, Gliemann J, Jakes R 1999 ␣-Synuclein binds to and stimulates the protein kinase Acatalyzed phosphorylation of serine residues 262 and 356. J Biol Chem 274:25481–25489 Leroy E, Boyer R, Auburguer G, Lleube B, Ulm G, Mezey E 1998 The ubiquitin pathway in Parkinson’s disease. Nature 395:451– 452 Kitada T, Asakawa S, Hattori N, Matsumine H, Yamamura Y, Minoshima S, Yokochi M, Mizuno Y, Shimizu N 1998 Mutations in the Parkin gene cause autosomal recessive juvenile Parkinsonism. Nature 392:605– 608 Solano SM, Miller DW, Augood SJ, Young AB, Penney JB 2000 Expression of ␣-synuclein, Parkin and ubiquitin carboxy-terminal hydrolase L1 mRNA in human brain: genes associated with familial Parkinson’s disease. Ann Neurol 47:201–210 Margolis RL, Ross CA 2001 Expansion explosion: new clues to the pathogenesis of repeat expansion neurodegenerative diseases. Trends Mol Med 7:479 – 482 Ross CA 2002 Polyglutamine pathogenesis: emergence of unifying mechanisms for Huntington’s disease and related disorders. Neuron 35:819 – 822 Browne SE, Beal MF 2004 The energetics of Huntington’s disease. Neurochem Res 29:531–546 MacDonald ME, Vonsattel J-P, Shrinidhi J, Couropmitree NN, Cupples LA, Bird ED, Gusella JF, Myers RH 1999 Evidence for the GluR6 gene associated with younger onset age of Huntington’s disease. Neurology 53:1330 –1332 Butterworth NJ, Williams L, Bullock JY, Love DR, Faull RL, Dragunow M 1998 Trinucleotide (CAG) repeat length is positively correlated with the degree of DNA fragmentation in Huntington’s disease striatum. Neuroscience 87:49 –53 Chen S, Ferrone FA, Wetzel R 2002 Huntington’s disease age-ofonset linked to polyglutamine aggregation nucleation. Proc Natl Acad Sci USA 99:11884 –11889 Ross CA 1995 When less is more: pathogenesis in glutamine repeat neurodegenerative diseases. Neuron 15:493– 496 Lunkes A, Lindenberg KS, Ben-Haiem L, Weber C, Devys D, Landwehrmeyer GB, Mandel J-L, Trottier Y 2002 Proteases acting on mutant Huntington generate cleaved products that differentially build up cytoplasmic and nuclear inclusions. Mol Cell 10: 259 –269 Meriin AB, Zhang XQ, He HW, Newnam GP, Chernoff YO, Sherman MY 2002 Huntington toxicity in yeast model depends on polyglutamine aggregation mediated by a prion-like protein Rnq1. J Cell Biol 157:997–1004 Nishitoh H, Matsuazawa D, Tobiume K, Saegusa K, Takeda K, Inoue K, Horis S, Kakizuka A, Ichijo H 2002 ASK1 is essential for endoplasmic reticulum stress-induced neuronal cell death triggered by expanded polyglutamine repeats. Genes Dev 16:1345– 1355 Will RG, Ironside JW, Zeidler M, Cousens SN, Estibeiro K, Alperovitch A, Poser S, Pocchiari M, Hofman A, Smith PG 1996 A new variant of Creutzfeldt-Jakob disease in the UK. Lancet 347: 921–925 Weissmann C, Enari M, Klöhn PC, Rossi D, Flechsig E 2002 Transmission of prions. J Infect Dis 186:S157–S165 Weissman C 1991 Spongiform encephalopathies. The prion’s progress. Nature 349:569 –571 Oesch B, Westaway D, Wälchli M, McKinley MP, Kent SB, Aebersold R, Barry RA, Tempst P, Teplow DB, Hood LE, Prusiner 502 206. 207. 208. 209. 210. 211. 212. 213. 214. 215. 216. 217. 218. 219. 220. 221. 222. 223. 224. 225. 226. 227. 228. Endocrine Reviews, June 2005, 26(4):479 –503 SB,Weissmann C 1985 A cellular gene encodes scrapie PrP 27–30 protein. Cell 40:735–746 Basler K, Oesch B, Scott M, Westaway D, Walchli M, Groth DF, McKinley MP, Prusiner SB, Weissmann C 1986 Scrapie and cellular PrP isoforms are encoded by the same chromosomal gene. Cell 46:417– 428 Prusiner SB 1989 Scrapie prions. Annu Rev Microbiol 43:345–347 Stahl N, Baldwin MA, Teplow DB, Hood L, Gibson BW, Burlingame AL, Prusiner SB 1993 Structural studies of the scrapie prion protein using mass spectrometry and amino acid sequencing. Biochemstry 32:1991–2002 Riek R, Hornemann S, Wider G, Glockshuber R, Wüthrich K 1997 NMR characterization of the full-length recombinant murine prion protein, mPrP (23–231). FEBS Lett 413:282–288 Pan KM, Baldwin M, Nguyen J, Gasset M, Serban A, Groth D, Mehlhorn I, Huang Z, Fletterick RJ, Cohen FE, Prusiner SB 1993 Conversion of ␣-helices into -sheets features in the formation of the scrapie prion proteins. Proc Natl Acad Sci USA 90:10962–10966 Caughey BW, Dong A, Bhat KS, Ernst D, Hayes SF, Caughey WS 1991 Secondary structure analysis of the scrapie-associated protein PrP 27–30 in water by infrared spectroscopy. Biochemistry 30:7672– 7680 Jarrett JT, Lansbury PJ 1993 Seeding “one-dimensional crystallization” of amyloid: a pathogenic mechanism in Alzheimer’s disease and scrapie? Cell 73:1055–1058 Prusiner SB 1991 Molecular biology of prion diseases. Science 252:1515–1522 Kovacs GG, Zerbi P, Voigtländer T, Strohschneider M, Trabattoni G, Hainfellner JA, Budka H 2002 The prion protein in human neurodegenerative disorders. Neurosci Lett 329:269 –272 Chesebro B 2003 Introduction to the transmissible spongiform encephalopathies or prion diseases. Br Med Bull 66:1–20 Will RG 1993 Epidemiology of Creutzfeldt-Jacob disease. Br Med Bull 49:960 –970 Fradkin JE, Schonberger LB, Mills JL, Gunn WJ, Piper JM, Wysowski DK, Thomson R, Durako S, Brown P 1991 CreutzfeldtJacob disease in pituitary growth hormone recipients in the United States. JAMA 265:880 – 884 Will RG, Zeidler M, Stewart GE, Macleod MA, Ironside JW, Cousens SN, Mackenzie J, Estibeiro K, Green AJ, Knight RS 2000 Diagnosis of new variant Creutzfeldt-Jakob disease. Ann Neurol 47:575–582 Kapaki E, Kolidireas K, Paraskevas GP, Michalopoulou M, Patsouris E 2001 Highly increased CSF protein and decreased -amyloid (1– 42) in sporadic CJD: a discrimination from Alzheimer’s diasese. J Neurol Neurosurg Psychiatry 71:401– 403 Otto M, Wiltfang J, Cepek L, Neumann M, Mollenhauer B, Steinacker P, Ciesielczyk B, Schult-Schaeffer W, Kretzschmar HA, Poser S 2002 protein and 14 –3-3 protein in the differential diagnosis of Creutzfeldt-Jakob disease. Neurology 58:192–197 Narang H 2002 A critical review of the nature of the spongiform encephalopathy agent: protein theory versus virus theory. Exp Biol Med 227:4 –19 Raymond GJ, Hope J, Kocisko DA, Priola SA, Raymond LD, Bossers A, Ironside J, Will RG, Chen SG, Petersen RB, Gambetti P, Rubenstein R, Smits MA, Lansbury Jr PT, Caughey B 1997 Molecular assessment of the transmissibilities of BSE and scrapie to humans. Nature 388:285–288 Horwitz J, Bova MP, Ding LL, Haley DA, Stewart PL 1999 Lens ␣-crystallin: function and structure. Eye 13:403– 408 Dasgupta S, Hohman TC, Carper D 1992 Hypertonic stress induces ␣ B-crystallin expression. Exp Eye Res 54:461– 470 Aoyama A, Frohli E, Schafer R, Klemenz R 1993 ␣ B-crystallin expression in mouse NIH 3T3 fibroblasts: glucocorticoid responsiveness and involvement in thermal protection Mol Cell Biol 13: 1824 –1835 Mehlen P, Schulze-Osthoff K, Arrigo AP 1996 Small stress proteins as novel regulators of apoptosis. Heat shock protein 27 blocks Fas/APO-1- and staurosporine-induced cell death. J Biol Chem 271:16510 –16514 Horwitz J 1992 ␣-Crystallin can function as a molecular chaperone. Proc Natl Acad Sci USA 89:10449 –10453 Rao PV, Huang Q-L, Horwitz J, Zigler JS 1995 Evidence that Castro-Fernández et al. • Protein Routing in Health and Disease 229. 230. 231. 232. 233. 234. 235. 236. 237. 238. 239. 240. 241. 242. 243. 244. 245. 246. 247. ␣-crystallin prevents non-specific protein aggregation in the intact eye lens. Biochim Biophys Acta 1245:439 – 447 Renkawek K, Voorter CEM, Bosman GJCGM, van Workim FPA, de Jong WW 1994 Expression of ␣ B-crystallin in Alzheimer’s disease. Acta Neuropathol 87:155–160 Lowe J, Landon M, Pike I, Spendlove I, McDermott H, Mayer RJ 1990 Dementia with -amyloid deposition: involvement of ␣ Bcrystallin supports two main diseases. Lancet 336:515–516 Renkawek K, de Jong WW, Merck KB, Frenken CWGM, van Workim FPA, Bosman GJCGM 1992 ␣ B-crystallin is present in reactive glia in Creutzfeldt-Jakob disease. Acta Neuropathol 83: 324 –327 Head MW, Corbin E, Goldman JE 1993 Overexpression and abnormal modification of the stress proteins ␣ B-crystallin and HSP27 in Alexander disease. Am J Pathol 143:1743–1753 Vicart P, Caron A, Guicheney P, Li A, Prevost M-C, Faure A, Chateau D, Chapon F, Tome F, Dupret JM, Paulin D, Fardeau M 1998 A missense mutation in the ␣B-crystallin chaperone gene causes a desmin-related myopathy. Nat Genet 20:92–95 Bova MP, Yaron O, Huang Q, Ding L, Haley DA, Stewart PL, Horwitz J 1999 Mutation R120G in ␣B-crystallin, which is linked to a desmin-related myopathy, results in an irregular structure and defective chaperone-like function. Proc Natl Acad Sci USA 96: 6137– 6142 Burrows JAJ, Willis LK, Perlmutter DH 2000 Chemical chaperones mediate increased secretion of mutant ␣1-antitrypsin (␣1-AT) Z: a potential pharmacological strategy for prevention of liver injury and emphysema in ␣1-AT deficiency. Proc Natl Acad Sci USA 97:1796 –1801 Teckman J, Qu D, Perlmutter DH 1996 Molecular pathogenesis of liver disease in ␣1-antitrypsin deficiency. Hepatology 24:1504 – 1516 Marcus NY, Perlmutter DH 2000 Glucosidase and mannosidase inhibitors mediate increased secretion of mutant ␣1 antitrypsin Z. J Biol Chem 275:1987–1992 Morello JP, Salahpour A, Laperrièr A, Bernier V, Arthus MF, Lonergan M, Petäjä-Repo U, Angers S, Morin D, Bichet DG, Bouvier M 2000 Pharmacological chaperones rescue cell-surface expression and function of misfolded V2 vasopressin receptor mutants. J Clin Invest 105:887– 895 Friedler A, DeDecker BS, Freund SMV, Rüdiger CBS, Fersht AR 2003 Structural distortion of p53 by the mutation R249S and its rescue by a designed peptide: implications for “mutant conformation.” J Mol Biol 336:187–196 Soto C, Kascsak RJ, Saborio GP, Aucouturier P, Wisniewski T, Prelli R, Kascsak R, Mendez E, Harris DA, Ironside J, Tagliavini F, Carp RI, Frangione B 2000 Reversion of prion protein conformational changes by synthetic -sheet breaker peptides. Lancet 355:192–197 Permanne B, Adessi C, Saborio GP, Fraga S, Frossard MJ, Van Dorpe J, Dewachter I, Banks WA, Van Leuven F, Soto C 2002 Reduction of amyloid load and cerebral damage in a transgenic animal model of Alzheimer’s disease by treatment with a b-sheet breaker peptide. FASEB J 16:860 – 862 Merlini G, Ascari E, Amboldi N, Bellotti V, Arbustini E, Perfetti V, Ferrari M, Zorzoli I, Marinone MG, Garini P, Diegoli M, Trizio D, Ballinari D 1995 Interaction of the anthracycline 4⬘-yodo-4⬘deoxydoxorubicin with amyloid fibrils: inhibition of fibrilogenesis. Proc Natl Acad Sci USA 92:2959 –2963 Forloni G, Colombo L, Girola L, Tagliavini F, Salmona M 2001 Anti-amyloidogenic activity of tetracyclines: studies in vitro. FEBS Lett 487:404 – 407 Findels MA 2002 Peptide inhibitors of  amyloid aggregation. Curr Top Med Chem 2:417– 423 Kelly JW, Balch WE 2003 Amyloid as a natural product. J Cell Biol 161:461– 462 Jarrett JT, Berger EP, Lansbury Jr PT 1993 The carboxy terminus of the  amyloid protein is critical for the seeding of amyloid formation: implications for the pathogenesis of Alzheimer’s disease. Biochemistry 32:4693– 4697 Haass C, De Strooper B 1999 The presenilins in Alzheimer’s disease-proteolysis holds the key. Science 286:916 –919 Castro-Fernández et al. • Protein Routing in Health and Disease 248. Cohen FE, Kelly JW 2003 Therapeutic approaches to proteinmisfolding diseases. Nature 426:905–909 249. Spooner ET, Desal RV, Mori C, Leverone JF, Lemere CA 2002 The generation and characterization of potentially therapeutic A antibodies in mice: differences according to strain and immunization protocol. Vaccine 21:290 –297 250. Nicoll JA, Wilkinson D, Holmes C, Steart P, Markham H, Weller RO 2003 Neuropathology of human Alzheimer disease after immunization with amyloid- peptide: a case report. Nat Med 9:448 – 452 251. Heiser V, Scherzinger E, Boeddrich A, Nordhoff E, Lutz R, Schugardt N, Lehrach H, Wanker EE 2000 Inhibition of Huntington fibrillogenesis by specific antibodies and small molecules: implications for Huntington’s disease therapy. Proc Natl Acad Sci USA 97:6739 – 6744 252. Wellington CL, Hayden MR 1997 Of molecular interactions, mice and mechanisms: new insights into Huntington’s disease. Curr Opin Neurol 10:291–298 253. Wellington CL, Singaraja R, Ellerby L, Savill J, Roy S, Leavitt B, Cattaneo E, Hackam A, Sharp A, Thornberry N, Nicholson DW, Bredesen DE, Hayden MR 2000 Inhibiting caspase cleavage of Huntington reduces toxicity and aggregate formation in neuronal and nonneuronal cells. J Biol Chem 275:19831–19838 254. Chen M, Ona VO, Li M, Ferrante RJ, Fink KB, Zhu S, Bian J, Guo L, Farrell LA, Hersch SM, Hobbs W, Vonsattel JP, Cha JH, Friedlander RM 2000 Minocycline inhibits caspase-1 and caspase-3 expression and delays mortality in a transgenic mouse model of Huntington disease. Nat Med 6:797– 801 255. Korth C, May BCH, Cohen FE, Prusiner SB 2001 Acridine and phenothiazine derivatives as pharmacotherapeutics for prion diseases. Proc Natl Acad Sci USA 98:9836 –9841 Endocrine Reviews, June 2005, 26(4):479 –503 503 256. May BCH, Fafarman AT, Hong SB, Rogers M, Deady LW, Prusiner SB, Cohen FE 2003 Potent inhibition of scrapie prion replication in cultured cells by bis-acridines. Proc Natl Acad Sci USA 100:3416 –3421 257. Vogtherr M, Grimme S, Elshorst B, Jacobs DM, Fiebig K, Griesinger C, Zahn R 2003 Antimalarial drug quinacrine binds to C-terminal helix of cellular prion protein. J Med Chem 46:3563– 3564 258. Peretz D, Williamson RA, Kaneko K, Vergara J, Leclerc E, Schmitt-Ulms G, Mehlhorn IR, Legname G, Wormald MR, Rudd PM, Dwek RA, Burton DR, Prusiner SB 2001 Antibodies inhibit prion propagation and clear cell culture of prion infectivity. Nature 412:739 –743 259. Blass JP 2001 Brain metabolism and brain disease: is metabolic deficiency the proximate cause of Alzheimer dementia? J Neurosci Res 66:851– 856 260. Nourhashemi F, Gillette-Guyonnet S, Andrieu S, Ghisolfi A, Ousset PJ, Grandjean H, Grand A, Pous J, Vellas B, Albarede JL 2000 Alzheimer disease: protective factors. Am J Clin Nutr 71: 643S– 649S 261. Grundman M 2000 Vitamin E and Alzheimer disease: the basis for additional clinical trials. Am J Clin Nutr 71:630S– 636S 262. Hughes RE, Olson JM 2001 Therapeutic opportunities in polyglutamine disease. Nat Med 7:419 – 423 263. Perlmutter DH 2000 Liver injury in ␣1-antitrypsin deficiency. Clin Liver Dis 4:387– 408 264. Ulloa_Aguirre A, Brothers S, Janovick J, Conn PM 2004 Pharmacological rescue of conformationally-defective proteins: implications for the treatment of human disease. Traffic 5:821– 837 Endocrine Reviews is published bimonthly by The Endocrine Society (http://www.endo-society.org), the foremost professional society serving the endocrine community.
© Copyright 2025 Paperzz