Báo cáo khoa học: Amyloid oligomers: formation and toxicity of Ab oligomers ppt

11 516 0
Báo cáo khoa học: Amyloid oligomers: formation and toxicity of Ab oligomers ppt

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

MINIREVIEW Amyloid oligomers: formation and toxicity of Ab oligomers Masafumi Sakono 1,2 and Tamotsu Zako 1 1 Bioengineering Laboratory, RIKEN Institute, Wako, Saitama, Japan 2 PRESTO, Japan Science and Technology Agency, Kawaguchi, Saitama, Japan Introduction Alzheimer’s disease (AD) is an age-related, progressive degenerative disorder characterized by the loss of synapses and neurons from the brain, and by the accu- mulation of extracellular protein-containing deposits (referred to as ‘senile plaques’) and neurofibrillary tangles [1]. Amyloid b-peptide (Ab; 39–43 amino acids in length) is the principal component of plaques. Ab is produced by the proteolytic cleavage of the parental amyloid precursor protein (APP) that localizes to the plasma membrane, trans-Golgi network, endoplasmic reticulum (ER) and endosomal, lysosomal and mito- chondrial membranes. Synthetic Ab spontaneously aggregates into b-sheet-rich fibrils, resembling those in plaques. As insoluble fibrillar aggregates are neuro- toxic in vivo and in vitro, it has long been hypothesized that fibrils cause neurodegeneration in AD [2]. However, debate over this ‘amyloid cascade hypothe- sis’ remains contentious. The number of senile plaques in a particular region of the AD brain correlates poorly with the local extent of neuron death or synaptic loss, or with cognitive impairment [3]. However, recent studies show a robust correlation between the soluble Ab oligomer levels and the extent of synaptic loss and severity of cognitive impairment [4–9]. The term ‘soluble’ refers to any form of Ab that is soluble in aqueous buffer and remains in solution after high-speed centrifugation, indicating that it is not insoluble fibrillar Ab. Assemblies ranging from dimers to 24-mers, or even those of higher molecular weight (MW), have been reported as Ab oligomers [5,10,11]. Soluble Ab oligomers are reportedly more cytotoxic than fibrillar Ab aggregates in general, and Keywords Alzheimer’s disease; amyloid b; formation and toxicity mechanism; intracellular and extracellular oligomers; soluble amyloid oligomers Correspondence T. Zako, Bioengineering Laboratory, RIKEN Institute, 2-1 Hirosawa, Wako, Saitama, 351-0198 Japan Fax: +81 48 462 4658 Tel: +81 48 467 9312 E-mail: zako@riken.jp (Received 4 September 2009, revised 11 December 2009, accepted 6 January 2010) doi:10.1111/j.1742-4658.2010.07568.x Alzheimer’s disease (AD) is an age-related, progressive degenerative dis- order that is characterized by synapse and neuron loss in the brain and the accumulation of protein-containing deposits (referred to as ‘senile plaques’) and neurofibrillary tangles. Insoluble amyloid b-peptide (Ab) fibrillar aggregates found in extracellular plaques have long been thought to cause the neurodegenerative cascades of AD. However, accumulating evidence suggests that prefibrillar soluble Ab oligomers induce AD-related synaptic dysfunction. The size of Ab oligomers is distributed over a wide molecular weight range (from < 10 kDa to > 100 kDa), with structural polymor- phism in Ab oligomers of similar sizes. Recent studies have demonstrated that Ab can accumulate in living cells, as well as in extracellular spaces. This review summarizes current research on Ab oligomers, focusing on their structures and toxicity mechanism. We also discuss possible formation mechanisms of intracellular and extracellular Ab oligomers. Abbreviations AD, Alzheimer’s disease; ADDL, Ab-derived diffusible ligand; APP, amyloid precursor protein; Ab, amyloid-b peptide; ER, endoplasmic reticulum; FCS, fluorescence correlation spectroscopy; HD, Huntington’s disease; LTP, long-term potentiation; MW, molecular weight; NGF, nerve growth factor; NMDAR, N-methyl- D-aspartate (NMDA)-type glutamate receptor; PD, Parkinson’s disease; polyQ, polyglutamine; PrP C, cellular prion protein. 1348 FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS inhibit many critical neuronal activities, including long-term potentiation (LTP), a classic model for syn- aptic plasticity and memory loss in vivo and in culture [12–15]. These studies strongly support the idea that soluble Ab oligomers are the causative agents of AD; however, the biological and structural characteristics of Ab oligomers and their formation mechanism remain unclear. Structure and size of soluble Ab oligomers Many types of natural and synthetic Ab oligomers of different sizes and shapes have been reported, which accounts for their biological and structural diversity and for the complexity of AD pathology (reviewed in [4,5,9–11]). SDS-stable dimers and trimers have been found in the soluble fractions of human brain and amyloid plaque extracts, which suggests that these low-MW Ab oligomers could be the fundamental building blocks of larger oligomers or insoluble amy- loid fibrils [16–18]. Ab oligomers of similar sizes have also been secreted from cultured cells and have been shown to inhibit LTP in vitro [14]. The high toxicity of low-MW Ab oligomers is also supported by in vitro studies showing that Ab dimers are threefold more toxic than monomers, and that Ab tetramers are 13-fold more toxic [19]. Recently, Lesne et al. [13] demonstrated that the level of SDS-stable Ab nonamers and dodecamers (referred to as Ab*56) correlated with memory deficits in an APP transgenic Tg2576 mice model. Purified dodecamers also induced a significant fall-off in the spatial memory performance of wild-type rats. These results suggest that nonamers and dodecamers are associated with deleterious effects on cognition. However, it is unlikely that these oligomers alone cause brain dysfunction. For example, young Tg2576 mice showed decreased dendritic spine density in the dentate gyrus, impaired LTP and impaired contextual fear conditioning, all at an age before the first dodecamer was detected [13,20]. Another recent report also showed that the Ab*56 levels are not correlated with memory deficits in a certain transgenic mice model [21]. These results sug- gest that Ab*56 is not the only key determinant of memory impairment. These oligomers could be classified as low-MW (< 50 kDa) oligomers. However, natural Ab oligomers with a wide-ranging MW distribution (from < 10 kDa to > 100 kDa) have been found in the AD brain [22], suggesting that Ab oligomers of various sizes are associated with the disease. There are also many reports of toxic oligomers from synthetic Ab. Synthetic Ab forms fibrillar aggregates that have properties similar to those found in AD plaques in the brain. In vitro studies using synthetic Ab are useful to complement efforts to determine the disease mechanism. Snyder et al. [23] detected the for- mation of soluble Ab assemblies, rather than fibrils, using an analytical ultracentrifugation technique, and Lambert et al. [12] reported the formation of small Ab globular oligomers (5 nm in diameter) in Hams-F12 medium, which were referred to as Ab-derived diffusible ligands (ADDLs). Importantly, ADDLs strongly bound to the dendritic arbors of cultured neurons, caused neuronal cell death and blocked LTP. The finding of ADDL in soluble brain extracts from the human AD brain using ADDL-specific antibody supports the idea that the existence of ADDLs in the human AD brain causes disease [24]. The formation of annular Ab oligomers, with an outer diameter of 8–12 nm and an inner diameter of 2.0– 2.5 nm (150–250 kDa), has also been reported [25,26]. As these annular Ab oligomers could be preferentially formed from mutant Ab (such as those carrying the Arctic mutation), and because the amyloid ‘pore’ resem- bles bacterial cytolytic b-barrel pore-forming toxins, it has been suggested that these doughnut-like oligomers could be responsible for the Ab-associated cytotoxicity [26]. The largest globular assemblies are amylospheroids [27], which are highly neurotoxic, off-pathway, sphe- roidal structures with diameters of 10–15 nm. Although A b oligomer structures at atomic resolu- tion are unclear, studies using conformation-dependent antibodies suggest that structural variants could exist among even morphologically similar Ab oligomers. A difference in antibody-binding properties indicates a difference in epitope exposure. For example, Glabe et al. used two antibodies – A11 and OC – which are specific for oligomers and fibrils, respectively. They proposed two distinct types of oligomers: prefibrillar oligomers that are A11-positive and OC-negative, and fibrillar oligomers that are A11-negative and OC-posi- tive [10]. As prefibrillar oligomers are not recognized by the fibril-specific antibodies, and are considered to be transient intermediates in the fibril-formation pro- cess, a conformational change is necessary for them to become fibrils. It should be noted that the oligomer- specific A11 antibody also recognizes soluble oligomers from various proteins, such as those from a-synuclein, islet amyloid polypeptide, polyglutamine (PolyQ), lyso- zyme, human insulin and prion peptide (106–126) [28]. These findings suggest that various proteins may form prefibrillar oligomers that share a common structure regardless of their amino acid sequence [8,28]. How- ever, because the fibrillar oligomers are recognized by the fibril-specific antibody, but not by A11, they at M. Sakono and T. Zako Formation of toxic Ab oligomers FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS 1349 least possess the structural characteristics of fibrils. Thus, it is plausible that the fibrillar oligomer might represent fibril nuclei to which the monomers can attach before elongation [10]. Ab oligomers formed at a low pH, but not those formed at a neutral pH, are recognized by the 6E10 antibody [29]. These results strongly suggest the existence of a structural polymor- phism of A b oligomers. There have been several other attempts to examine Ab oligomer structures to elucidate the mechanism of formation of Ab oligomers. Studies using atomic force microscopy and scanning tunneling microscopy showed that the structures of dimers, tetramers and other low- MW Ab oligomers were consistent with the model of the Ab monomers as b-hairpins [30,31]. These low- MW Ab oligomers are relatively compact, being 1–3 nm in height and 5–10 nm in width ⁄ length, and could be the fundamental building blocks of larger oligomers and protofibrils. Bernstein et al. [18] developed a new method, called electrospray-ionization ion-mobility mass spectrometry, to obtain oligomer size distributions and the qualita- tive structure of each oligomer. Electrospray ionization allows a fixed population of different Ab oligomer states in solution to be isolated from one another, and their size and shape could be determined using ion- mobility spectrometry. By analyzing the cross- sectional area of each oligomer obtained by ion mobil- ity, the structure of the Ab 42 tetramer is theoretically assumed to take an open ‘V’ form, which is neither linear nor square. A planar hexagon form was assumed for the Ab 42 hexamer. It is interesting to note that a stacked hexamer paranuclei structure, rather than side-by-side planar hexagons, was suggested for the Ab 42 dodecamer, (Ab*56). These authors also showed that oligomer size distribution was very different between Ab 42 and Ab 40 , consistent with previous studies, indicating that their oligomerization pathways are different [25]. Although the oligomer structure in the gas phase may not be completely identical to that in solution, the information obtained using this novel technique probably reflects the characteristics of Ab oligomers, at least in part, and may be useful for understanding the physical aspects of Ab oligomers. Further attempts to characterize Ab oligomers at a single molecule level have been performed. Dukes et al. [32] and Ding et al. [33] recently reported oligomer size determination with single molecule spectroscopy using fluorescently labeled Ab.By directly counting the photobleaching steps in the fluorescence of each oligomer on a cover-glass sur- face, the number of monomer molecules in individual oligomers could be determined, enabling the determi- nation of more precise oligomer size distributions. For example, an Ab 40 sample incubated at a neutral pH was shown to be a mixture of monomers, dimers, trimers and tetramers, and the presence of zinc ion in the sample buffer increased the number of tetramers [33]. Although application of this method is limited to small oligomers, the single mol- ecule approach overcomes the limitations of resolu- tion and sample heterogeneity. Analyses of the size of the Ab oligomer in solution at the single molecule level have also been performed using fluorescence correlation spectroscopy (FCS), which detects the fluorescence of dye-labeled molecules in a very small confocal volume excited by a sharply focused laser beam [34]; FCS enables estimation of the size distribution of an oligomeric species in solution over a wide range of sizes (from monomers to large soluble particles) with a good time resolution ( 1 min). From the fluorescence intensity fluctuations, one can calculate the number of molecules in the confocal volume and their diffusion times (corresponding to size). For example, the oligomer size distribution of the incubated Ab 40 sample showed a peak ranging from 50 to 120 nm, indicating the formation of large oligomers [34]. It should be noted that a low concen- tration (nm) of dye-labeled protein is required for single molecule detection using FCS. Orte et al. [35] used a two-color single-molecule fluorescence technique (‘two-color coincidence detec- tion’) to characterize oligomer formation of the SH3 domain of phosphatidylinositol 3-kinase, which is known to form small granular toxic aggregates. In this technique, fluorescence bursts from single oligomer particles made from protein monomers, each labeled with one of two fluorescent dyes that emit light at dif- ferent wavelengths, are observed using optics similar to that of FCS. The coincident detection of both emit- ted wavelengths with dual excitation indicates the presence of oligomers consisting of more than one molecule. The size and population of oligomers can be determined from the fluorescence intensity and the frequency of such coincident bursts, respectively. Oligomer stability at low concentrations can be exam- ined from changes in the monomer in solution, which can be evaluated from the frequency of noncoincident monomer bursts. Experimental data suggest that the stability of the SH3 domain of the phosphatidylinosi- tol 3-kinase oligomer changes from unstable oligomer to stable oligomers that show no monomer dissocia- tion [35]. It would be interesting to apply this method to examine the time-course of the stability of Ab oligomers. Formation of toxic Ab oligomers M. Sakono and T. Zako 1350 FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS Although these in vitro studies provide insight into how Ab monomers assemble into oligomeric com- plexes, further characterizations, by such as visualiza- tion of Ab oligomer at the molecular level in living cells and animal models, may be required to elucidate the mechanism of formation of Ab oligomers. Possible mechanism of soluble oligomer formation and toxicity The mechanism of formation of soluble Ab oligomer in vivo remains unclear. Glabe et al. [10] proposed that multiple Ab oligomer conformations were produced via different pathways, indicating the complexity of the oligomer formation mechanism. The mechanisms of formation may also differ for extracellular and intracellular oligomers. In this section, we discuss pos- sible formation mechanisms of extracellular and intra- cellular Ab oligomers, and also discuss how these Ab oligomers can cause cell death or neuronal impairment (Figs 1 and 2). Extracellular soluble Ab oligomer formation and its toxicity A recent study by Yamamoto et al. [36] showed the formation of toxic Ab oligomers in the presence of GM1 ganglioside. This Ab oligomer was spherical, with a diameter of 10–20 nm and a molecular mass of 200–300 kDa, and therefore much larger than ADDL. Furthermore, Ab monomers produced extracellularly can interact with GM1, and an Ab complex with GM1 has been found in AD brain [37]. These observations support the idea that extracellular soluble Ab oligo- mers could be formed by GM1. The Ab oligomer– GM1 complex is not recognized using a seed-specific mAb, suggesting that the GM1-induced Ab oligomer is formed via a pathway distinct from that of fibril Fig. 1. Formation and toxicity mechanisms of extracellular Ab oligomers. Ab is released extracellularly as a product of proteolytically cleaved, plasma membrane-localized amyloid precursor protein (APP). Extracellular Ab oligomers can be formed in the presence of GM1 ganglioside on the cell membrane. GM1 induces Ab oligomer-induced neuronal cell death mediated by nerve growth factor (NGF) receptors. Toxic non- fibrillar Ab is also produced in the presence of aB-crystallin and ApoJ. A cellular prion protein (PrP C ) acts as an Ab oligomer receptor with nanomolar affinity, and mediates synaptic dysfunction. Furthermore, the membrane pore is formed by Ab oligomers. The pores allow abnor- mal flow of ions, such as Ca 2+ , which causes cellular dysfunction. Binding of Ab oligomers to the NMDA-type glutamate receptor (NMDAR) also causes abnormal calcium homeostasis, leading to increased oxidative stress and synapse loss. Binding of Ab oligomers to the Frizzled (Fz) receptor can inhibit Wnt signaling, leading to cell dysfunctions such as tau phosphorylation and neurofibrillary tangles. Moreover, Ab oligomer can induce insulin receptor loss from the neuronal surface and impaired kinase activity related to long-term potentiation. M. Sakono and T. Zako Formation of toxic Ab oligomers FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS 1351 formation [36]. Furthermore, nonfibrillar Ab can be produced in the presence of aB-crystallin [38] and clus- terin (also known as Apo J) [39], suggesting that extra- cellular Ab oligomers could be formed by various bio- components such as proteins and gangliosides (Fig. 1). The GM1-induced Ab oligomer induces neuronal cell death mediated by nerve growth factor (NGF) receptors, suggesting that binding of the Ab oligomer to the NGF receptor is important for the toxicity mechanism [36] (Fig. 1). Potent alternation of NGF- mediated signaling by ADDL supports this concept [40]. Moreover, previous studies suggested that apopto- tic cell death occurs through the interaction of Ab with low-affinity NGF receptor [pan neurotrophin receptor (p75NTR)] and the activation of downstream signaling molecules, such as c-Jun N-terminal kinase (reviewed in ref. [41]). However, it has also been demonstrated that p75NTR promotes neuronal survival and differen- tiation, indicating that p75NTR might have diverse functions in both cell death and cell survival [42]. Consistent with this notion, there are also conflicting reports showing that p75NTR is protective against Ab toxicity [43,44]. These results imply that the NGF-mediated toxicity mechanism is complicated. Other reports on neuronal receptor-mediated toxicity mechanisms (reviewed in ref. [9]) have shown that ADDL binding to an N-methyl-d-aspartate (NMDA)- type glutamate receptor (NMDAR) causes abnormal calcium homeostasis, leading to increased oxidative stress and synapse loss [45,46]. ADDL can also induce the loss of insulin receptors from the neuronal surface [47,48] and impair LTP-associated kinase activity [49]. However, such insulin receptor impairment is inhibited by extracellular insulin, suggesting that insulin plays an important role in oligomer-induced cell death. Magdesian et al. [50] showed that Ab oligomers bind- ing to the Frizzled (Fz) receptor, an acceptor of Wnt protein, inhibited Wnt signaling, leading to cellular dysfunction. Wnt signaling, which promotes progenitor cell proliferation and directs cells into a neuronal Fig. 2. Formation and toxicity mechanisms of intracellular Ab oligomers. Ab can be localized intracellularly by the uptake of extracellular Ab or by the cleavage of APP in endosomes generated from the ER or the Golgi apparatus. Extracellular Ab is internalized through various receptors and transporters, such as formyl peptide receptor-like protein 1 (FPRL1) or scavenger receptor for advanced glycation end-products (RAGE). These receptor–Ab complexes are internalized into early endosomes. Most Ab in the endosome is degraded by the endosome ⁄ lyso- some system. However, Ab in the lysosome can leak into the cytosol by destabilization of the lysosome membrane. Although cytosolic Ab can be degraded by the proteasomal degradation system, inhibition of the proteasome function by Ab oligomers causes cell death. Suppres- sion of protein aggregation by interactions with various cellular proteins, such as prefoldin (PFD) or other molecular chaperones, may cause the formation of Ab oligomers. Formation of toxic Ab oligomers M. Sakono and T. Zako 1352 FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS phenotype during brain development, inactivated glycogen synthase kinase-3b (GSK-3b) and increased b-catenin levels. Inhibition of Wnt signaling by Ab oligomers causes tau phosphorylation and neurofibril- lary tangles, which suggests a Wnt ⁄ b-catenin toxicity pathway [50]. A recent report by Nimmrich et al. [51] showed that Ab oligomers can also impair presynaptic P ⁄ Q-type calcium currents, which are related to neurotransmis- sion and synaptic plasticity in the brain, at both gluta- matergic and gamma-amino butyric acid (GABA)-ergic synapses. This impairment is specific for Ab oligomers, but not for Ab monomer or fibrils. Although the detailed mechanism of this impairment remains unclear, the interaction of Ab oligomers with synaptic proteins or channels may cause modification of the P ⁄ Q current. By contrast, another study showed that the cell membrane could be destabilized by the Ab oli- gomer [52]. The membrane pores formed by the Ab oligomer would allow the abnormal flow of ions, such as Ca 2+ , suggesting another plausible mechanism for Ab oligomer toxicity [53,54]. Recent observations by Lauren et al. [55] indicate that cellular prion protein (PrP C ) can act as an Ab oligomer receptor with a nanomolar affinity, mediating synaptic dysfunction. Although misfolded prion protein (PrP Sc ) is thought to cause prion disease, the interaction between the Ab oli- gomer and the prion does not require the infectious PrP Sc conformation. This interaction may disrupt the interaction between PrP C and a co-receptor, such as NMDAR, impairing the neuron signal-transduction pathways. This discovery by Lauren et al. also suggests that AD is linked with other neurodegenerative diseases. Recently, interactions between Ab and a-synuclein in vivo and in vitro have recently been observed [56,57]. Alpha-synuclein is an aggregation-prone protein that causes Parkinson’s disease (PD), and interactions between a-synuclein and Ab therefore indicate that AD and PD could be related. Ab also promotes a-syn- uclein aggregation and toxicity. These results suggest that the AD and PD pathologies could overlap. Inter- estingly, interactions between Ab and a-synuclein induce the formation of hybrid pore-like oligomers [58]. Ab-treated cells expressing a-synuclein display increased current amplitudes and calcium influx, consistent with the formation of cation channels. It is therefore assumed that the hybrid pore-like oligomers may alter neuronal activity and cause neurodegenera- tion. These observations support the idea that there are various Ab oligomer-formation pathways, and that cell death might occur via multiple pathways (Fig. 1). Intracellular Ab Although Ab was first identified as a component of extracellular amyloid plaques, ample evidence has demonstrated that Ab is also generated intracellularly (reviewed in ref. [6]). Besides the plasma membrane, APP localizes to the trans-Golgi network, to the ER and to the endosomal, lysosomal and mitochondrial membranes. Ab is produced by the sequential cleavage of APP by b-secretase (also known as BACE) and c-secretase in endosomes as well as at the plasma membrane [59]. Ab is also produced intracellularly within the ER and the trans-Golgi network system along the secretory pathway. Identification of the intracellular protein, endoplasmic reticulum associated binding protein (ERAB), which binds to Ab, also strongly suggests the existence of intracellular Ab [60]. In addition to Ab being produced intracellularly, previously secreted Ab that forms the extracellular Ab pool can be taken up by cells and internalized into intracellular pools through various receptors and trans- porters, such as the nicotinic acetylcholine receptor, low-density lipoprotein receptor, formyl peptide recep- tor-like protein 1, NMDAR and the scavenger recep- tor for advanced glycation end-products [6] (Fig. 2). These receptor-associated Ab complexes could be internalized into endosomes. Recent findings also sup- port the idea that Ab is present within the cytosolic compartment. Intracellular accumulation of Ab in the multivesicular body is linked to cytosolic proteasome inhibition [61]. Furthermore, in vivo and in vitro pro- teasome inhibition also leads to higher Ab levels [62,63]. As the proteasome is primarily located within the cytosol, these findings strongly suggest that Ab is also located within the cytosolic compartment. Extra- cellular Ab can enter the cytosolic compartment and inhibit the proteasome activity of cultured neuronal cells [62]. Clifford et al. [64] showed that fluorescently labeled Ab which is injected into the tail of mice with a defective blood–brain barrier (which is common in AD patients) accumulates in the perinuclear cytosol of pyramidal neurons in the cerebral cortex. These obser- vations strongly support the notion that neurons can take up extracellular Ab in the cytosolic compartment. The destabilization of intracellular membranes may also contribute to the presence of cytosolic Ab. A high proportion of autophagy-related vesicular structures, which would suggest impaired maturation of autophagosomes to lysosomes, has been found in the AD brain, but not in the normal brain [65]. Although most Ab formed in endosomes is normally degraded within lysosomes, Ab can accumulate in lysosomes in M. Sakono and T. Zako Formation of toxic Ab oligomers FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS 1353 the AD brain. Ab within the lysosomal compartment destabilizes its membrane [66], which would also lead to the presence of Ab in the cytosolic compartment. Intracellular soluble A b oligomer formation and its toxicity How intracellular Ab monomers assemble and form soluble oligomers remains unclear. One possibility is that the uptake of extracellularly-produced Ab oligo- mers occurs via endocytic pathways or various recep- tors and transporters, as described above (Fig. 2). Another possibility is that the interaction of Ab with intracellular proteins results in oligomer formation. Recent observations by Yuyama et al., [67] showing GM1 accumulation in early endosomes, support the idea that intracellular GM1 could also induce Ab oli- gomer formation. Recently, we found formation of toxic high-MW (50–250 kDa) soluble Ab oligomers by the cytosolic molecular chaperone protein, prefoldin, in vitro [68]. In general, molecular chaperones stabilize and mediate the folding of unfolded proteins. Molecu- lar chaperones play essential roles in many cellular processes, such as protein folding, targeting, transpor- tation, degradation and signal transductions [69]. Prefoldin reportedly captures and delivers denatured protein to another cytosolic chaperone, chaperonin [70–73]. Our results also suggested that the interaction between prefoldin and A b oligomers prevents further aggregation and stabilizes the oligomer structure (Fig. 2). Molecular chaperones are potent suppressors of pro- tein aggregation, leading to neurodegenerative disor- ders such as AD, PD and Huntington’s disease (HD) [74–76]. Various molecular chaperones are upregulated in patients and co-localize with aggregated proteins in plaques ⁄ inclusion bodies. These molecular chaperones prevent aggregation in vivo and in vitro; for example, the cytosolic chaperonin CCT can inhibit aggregation of the polyglutamine (polyQ) expansion protein, which causes HD in vivo and in vitro [77–79]. Reduced CCT levels also enhance the aggregation and toxicity of pol- yQ in neuronal cells, strongly supporting the idea that molecular chaperones can be a defense against the aggregation of misfolded protein. Importantly, how- ever, our findings also suggest the possibility that the suppression of protein aggregation may cause the for- mation of toxic oligomeric species, which is consistent with previous results showing that toxic nonfibrillar Ab was produced in the presence of aB-crystallin [38] and clusterin (also known as Apo J) [39]. These results suggest that intracellular Ab oligomers could be pro- duced by interaction with various cellular proteins, including molecular chaperone proteins (Fig. 2). The toxicity mechanism of intracellular Ab oligo- mers also remains unclear. Microinjection of Ab or a cDNA-expressing cytosolic Ab induces the cell death of primary neurons and the simultaneous formation of low-MW Ab oligomers [80]. Furthermore, intracellular Ab accumulation is closely correlated with apoptotic cell death via the P53-BAX pathway [81]. Recently, Mousnier et al. [82] reported a possible prefoldin-medi- ated proteasomal protein-degradation pathway. It is therefore plausible that Ab oligomer–prefoldin com- plexes could bind to proteasome, causing proteasome dysfunction and subsequent cell death. This idea is supported by interaction studies between Ab oligomers and proteasome, which showed that the proteasomal function was inhibited while interacting with Ab [63]. Impairment of proteasomal function by the Ab oligo- mer also leads to age-related pathological accumula- tion of Ab and tau protein [63]. Recent research has shown that the dysfunction of autophagy, a lysosomal pathway for degrading organelles and proteins, is related to neurodegenerative diseases, including AD and PD [65,76]. These observations support the idea that the toxicity mechanism of intracellular oligomers may be different from that of extracellular oligomers (Fig. 2). However, more studies, particularly those focused especially on the proteolysis system in AD brains, are necessary to understand AD pathology in relation to intracellular soluble Ab oligomers. Concluding remarks It has long been argued that insoluble Ab fibrillar aggregates found in extracellular amyloid plaques initi- ate the neurodegenerative cascades of AD. However, recent emerging results indicate that prefibrillar soluble Ab oligomers are the key intermediates in AD-related synaptic dysfunction. Various amyloidogenic proteins can form toxic soluble oligomers, suggesting that solu- ble oligomers are the general key factors in various diseases such as AD, PD, HD and other amyloidosis [5,28,83]. Although much research effort is being direc- ted towards characterizing oligomer states, their con- formations and formation mechanisms remain unclear. Recent evidence suggests that the size of Ab oligomers is distributed in a wide MW range (from < 10 kDa to > 100 kDa), and that there is structural polymor- phism of Ab oligomers, even for those of a similar size. The biochemical properties of these oligomers in relation to disease pathology also seem to differ depending on their sizes and structures. Formation of toxic Ab oligomers M. Sakono and T. Zako 1354 FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS Ab can form various distinct oligomeric states via various pathways. The formation and toxicity mecha- nisms of extracellular and intracellular Ab oligomers can also be different from one another. Regardless of the complexity of the oligomer-formation mechanism, recent findings suggest that Ab oligomers can be formed through interactions between Ab mono- mers ⁄ oligomers and cellular proteins ⁄ biomolecules, such as molecular chaperones and lipids. Prevention of aggregation may cause the formation ⁄ stabilization of oligomer states. Acknowledgements The authors are grateful for financial support from the Japan Science Technology Agency (PRESTO Program, MS), RIKEN (Nano-scale Science and Technology Research, TZ) and the Japanese Society for the Pro- motion of Science (TZ). We wish to thank Drs Karin So ¨ rgjerd, Naofumi Terada (RIKEN) and Hiroshi Ku- bota (Akita University) for helpful comments. References 1 Selkoe DJ (2001) Alzheimer’s disease: genes, proteins, and therapy. Physiol Rev 81, 741–766. 2 Hardy JA & Higgins GA (1992) Alzheimer’s disease: the amyloid cascade hypothesis. Science 256, 184–185. 3 Terry RD, Maslia E & Hansen LA (1999) The neuropa- thology of Alzheimer disease and the structural basis of its cogninitve alterations. In Alzheimer Disease (Terry RD, Katzman R, Bick KL & Sisodia SS eds), pp 187– 206. Lipponcott Williams and Wilikins, Philadelphia. 4 Caughey B & Lansbury PT (2003) Protofibrils, pores, fibrils, and neurodegeneration: separating the responsi- ble protein aggregates from the innocent bystanders. Annu Rev Neurosci 26, 267–298. 5 Haass C & Selkoe DJ (2007) Soluble protein oligomers in neurodegeneration: lessons from the Alzheimer’s amyloid beta-peptide. Nat Rev Mol Cell Biol 8, 101–112. 6 Laferla FM, Green KN & Oddo S (2007) Intracellular amyloid-beta in Alzheimer’s disease. Nat Rev Neurosci 8, 499–509. 7 Klein WL, Krafft GA & Finch CE (2001) Targeting small Abeta oligomers: the solution to an Alzheimer’s disease conundrum? Trends Neurosci 24, 219–224. 8 Chiti F & Dobson CM (2006) Protein misfolding, func- tional amyloid, and human disease. Annu Rev Biochem 75, 333–366. 9 Ferreira ST, Vieira MN & De Felice FG (2007) Soluble protein oligomers as emerging toxins in Alzheimer’s and other amyloid diseases. IUBMB Life 59, 332–345. 10 Glabe CG (2008) Structural classification of toxic amy- loid oligomers. J Biol Chem 283, 29639–29643. 11 Roychaudhuri R, Yang M, Hoshi MM & Teplow DB (2009) Amyloid beta-protein assembly and Alzheimer disease. J Biol Chem 284, 4749–4753. 12 Lambert MP, Barlow AK, Chromy BA, Edwards C, Freed R, Liosatos M, Morgan TE, Rozovsky I, Trom- mer B, Viola KL et al. (1998) Diffusible, nonfibrillar ligands derived from Abeta1-42 are potent central nervous system neurotoxins. Proc Natl Acad Sci USA 95, 6448–6453. 13 Lesne S, Koh MT, Kotilinek L, Kayed R, Glabe CG, Yang A, Gallagher M & Ashe KH (2006) A specific amyloid-beta protein assembly in the brain impairs memory. Nature 440, 352–357. 14 Walsh DM, Klyubin I, Fadeeva JV, Cullen WK, Anwyl R, Wolfe MS, Rowan MJ & Selkoe DJ (2002) Naturally secreted oligomers of amyloid beta protein potently inhibit hippocampal long-term potentiation in vivo. Nature 416, 535–539. 15 Wang HW, Pasternak JF, Kuo H, Ristic H, Lambert MP, Chromy B, Viola KL, Klein WL, Stine WB, Krafft GA et al. (2002) Soluble oligomers of beta amyloid (1-42) inhibit long-term potentiation but not long-term depression in rat dentate gyrus. Brain Res 924, 133–140. 16 Podlisny MB, Ostaszewski BL, Squazzo SL, Koo EH, Rydell RE, Teplow DB & Selkoe DJ (1995) Aggrega- tion of secreted amyloid beta-protein into sodium dodecyl sulfate-stable oligomers in cell culture. J Biol Chem 270, 9564–9570. 17 Walsh DM, Tseng BP, Rydel RE, Podlisny MB & Selkoe DJ (2000) The oligomerization of amyloid beta-protein begins intracellularly in cells derived from human brain. Biochemistry 39 , 10831–10839. 18 Bernstein SL, Dupuis NF, Lazo ND, Wyttenbach T, Condron MM, Bitan G, Teplow DB, Shea J-E, Ruotolo BT, Robinson CV et al. (2009) Amyloid-b protein oligomerization and the importance of tetramers and dodecamers in the aetiology of Alzheimer’s disease. Nat Chem 1, 326–331. 19 Ono K, Condron MM & Teplow DB (2009) Structure- neurotoxicity relationships of amyloid beta-protein oligomers. Proc Natl Acad Sci USA 106, 14745–14750. 20 Jacobsen JS, Wu CC, Redwine JM, Comery TA, Arias R, Bowlby M, Martone R, Morrison JH, Pangalos MN, Reinhart PH et al. (2006) Early-onset behavioral and synaptic deficits in a mouse model of Alzheimer’s disease. Proc Natl Acad Sci USA 103, 5161–5166. 21 Lefterov I, Fitz NF, Cronican A, Lefterov P, Staufenb- iel M & Koldamova R (2009) Memory deficits in APP23 ⁄ Abca1+ ⁄ - mice correlate with the level of Abeta oligomers. ASN Neuro 1, e00006. 22 Kuo YM, Emmerling MR, Vigo-Pelfrey C, Kasunic TC, Kirkpatrick JB, Murdoch GH, Ball MJ & Roher AE (1996) Water-soluble Abeta (N-40, N-42) oligomers in normal and Alzheimer disease brains. J Biol Chem 271, 4077–4081. M. Sakono and T. Zako Formation of toxic Ab oligomers FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS 1355 23 Snyder SW, Ladror US, Wade WS, Wang GT, Barrett LW, Matayoshi ED, Huffaker HJ, Krafft GA & Holzman TF (1994) Amyloid-beta aggregation: selective inhibition of aggregation in mixtures of amyloid with different chain lengths. Biophys J 67, 1216–1228. 24 Lambert MP, Viola KL, Chromy BA, Chang L, Morgan TE, Yu J, Venton DL, Krafft GA, Finch CE & Klein WL (2001) Vaccination with soluble Abeta oligomers generates toxicity-neutralizing antibodies. J Neurochem 79, 595–605. 25 Bitan G, Kirkitadze MD, Lomakin A, Vollers SS, Ben- edek GB & Teplow DB (2003) Amyloid beta -protein (Abeta) assembly: Abeta 40 and Abeta 42 oligomerize through distinct pathways. Proc Natl Acad Sci U S A 100, 330–335. Epub 2002 Dec 2027. 26 Lashuel HA & Lansbury PT Jr (2006) Are amyloid dis- eases caused by protein aggregates that mimic bacterial pore-forming toxins? Q Rev Biophys 39, 167–201. 27 Hoshi M, Sato M, Matsumoto S, Noguchi A, Yasutake K, Yoshida N & Sato K (2003) Spherical aggregates of beta-amyloid (amylospheroid) show high neurotoxicity and activate tau protein kinase I ⁄ glycogen synthase kinase-3beta. Proc Natl Acad Sci USA 100, 6370–6375. 28 Kayed R, Head E, Thompson JL, McIntire TM, Milton SC, Cotman CW & Glabe CG (2003) Common struc- ture of soluble amyloid oligomers implies common mechanism of pathogenesis. Science 300, 486–489. 29 Necula M, Kayed R, Milton S & Glabe CG (2007) Small molecule inhibitors of aggregation indicate that amyloid beta oligomerization and fibrillization path- ways are independent and distinct. J Biol Chem 282, 10311–10324. 30 Mastrangelo IA, Ahmed M, Sato T, Liu W, Wang C, Hough P & Smith SO (2006) High-resolution atomic force microscopy of soluble Abeta42 oligomers. J Mol Biol 358, 106–119. 31 Losic D, Martin LL, Mechler A, Aguilar MI & Small DH (2006) High resolution scanning tunnelling microscopy of the beta-amyloid protein (Abeta1-40) of Alzheimer’s disease suggests a novel mechanism of oligomer assembly. J Struct Biol 155, 104–110. 32 Ding H, Wong PT, Lee EL, Gafni A & Steel DG (2009) Determination of the oligomer size of amyloido- genic protein beta-amyloid(1-40) by single-molecule spectroscopy. Biophys J 97, 912–921. 33 Dukes KD, Rodenberg CF & Lammi RK (2008) Moni- toring the earliest amyloid-beta oligomers via quantized photobleaching of dye-labeled peptides. Anal Biochem 382, 29–34. 34 Garai K, Sengupta P, Sahoo B & Maiti S (2006) Selec- tive destabilization of soluble amyloid beta oligomers by divalent metal ions. Biochem Biophys Res Commun 345, 210–215. 35 Orte A, Birkett NR, Clarke RW, Devlin GL, Dobson CM & Klenerman D (2008) Direct characterization of amyloidogenic oligomers by single- molecule fluorescence. Proc Natl Acad Sci USA 105, 14424–14429. 36 Yamamoto N, Matsubara E, Maeda S, Minagawa H, Takashima A, Maruyama W, Michikawa M & Yanagis- awa K (2007) A ganglioside-induced toxic soluble Abeta assembly. Its enhanced formation from Abeta bearing the Arctic mutation. J Biol Chem 282, 2646–2655. 37 Yanagisawa K (2007) Role of gangliosides in Alzhei- mer’s disease. Biochim Biophys Acta 1768, 1943–1951. 38 Stege GJ, Renkawek K, Overkamp PS, Verschuure P, van Rijk AF, Reijnen-Aalbers A, Boelens WC, Bosman GJ & de Jong WW (1999) The molecular chaperone alphaB-crystallin enhances amyloid beta neurotoxicity. Biochem Biophys Res Commun 262, 152–156. 39 Oda T, Wals P, Osterburg HH, Johnson SA, Pasinetti GM, Morgan TE, Rozovsky I, Stine WB, Snyder SW, Holzman TF et al. (1995) Clusterin (apoJ) alters the aggregation of amyloid beta-peptide (A beta 1-42) and forms slowly sedimenting A beta complexes that cause oxidative stress. Exp Neurol 136 , 22–31. 40 Chromy BA, Nowak RJ, Lambert MP, Viola KL, Chang L, Velasco PT, Jones BW, Fernandez SJ, Lacor PN, Horowitz P et al. (2003) Self-assembly of Abeta(1-42) into globular neurotoxins. Biochemistry 42, 12749–12760. 41 Coulson EJ (2006) Does the p75 neurotrophin receptor mediate Abeta-induced toxicity in Alzheimer’s disease? J Neurochem 98, 654–660. 42 Dechant G & Barde YA (2002) The neurotrophin receptor p75(NTR): novel functions and implications for diseases of the nervous system. Nat Neurosci 5, 1131–1136. 43 Costantini C, Della-Bianca V, Formaggio E, Chiamul- era C, Montresor A & Rossi F (2005) The expression of p75 neurotrophin receptor protects against the neuro- toxicity of soluble oligomers of beta-amyloid. Exp Cell Res 311, 126–134. 44 Zhang Y, Hong Y, Bounhar Y, Blacker M, Roucou X, Tounekti O, Vereker E, Bowers WJ, Federoff HJ, Goodyer CG et al. (2003) p75 neurotrophin receptor protects primary cultures of human neurons against extracellular amyloid beta peptide cytotoxicity. J Neuro- sci 23, 7385–7394. 45 De Felice FG, Velasco PT, Lambert MP, Viola K, Fernandez SJ, Ferreira ST & Klein WL (2007) Abeta oligomers induce neuronal oxidative stress through an N-methyl-D-aspartate receptor-dependent mechanism that is blocked by the Alzheimer drug memantine. J Biol Chem 282, 11590–11601. 46 Shankar GM, Bloodgood BL, Townsend M, Walsh DM, Selkoe DJ & Sabatini BL (2007) Natural oligo- mers of the Alzheimer amyloid-beta protein induce reversible synapse loss by modulating an NMDA-type Formation of toxic Ab oligomers M. Sakono and T. Zako 1356 FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS glutamate receptor-dependent signaling pathway. J Neurosci 27, 2866–2875. 47 De Felice FG, Vieira MN, Bomfim TR, Decker H, Velasco PT, Lambert MP, Viola KL, Zhao WQ, Ferreira ST & Klein WL (2009) Protection of synapses against Alzheimer’s-linked toxins: insulin signaling prevents the pathogenic binding of Abeta oligomers. Proc Natl Acad Sci USA 106, 1971–1976. 48 Zhao WQ, De Felice FG, Fernandez S, Chen H, Lambert MP, Quon MJ, Krafft GA & Klein WL (2008) Amyloid beta oligomers induce impairment of neuronal insulin receptors. FASEB J 22, 246–260. 49 Townsend M, Mehta T & Selkoe DJ (2007) Soluble Abeta inhibits specific signal transduction cascades common to the insulin receptor pathway. J Biol Chem 282, 33305–33312. 50 Magdesian MH, Carvalho MM, Mendes FA, Saraiva LM, Juliano MA, Juliano L, Garcia-Abreu J & Ferreira ST (2008) Amyloid-beta binds to the extracellular cysteine-rich domain of Frizzled and inhibits Wnt ⁄ beta-catenin signaling. J Biol Chem 283, 9359–9368. 51 Nimmrich V, Grimm C, Draguhn A, Barghorn S, Lehmann A, Schoemaker H, Hillen H, Gross G, Ebert U & Bruehl C (2008) Amyloid beta oligomers (A beta(1-42) globulomer) suppress spontaneous synaptic activity by inhibition of P ⁄ Q-type calcium currents. J Neurosci 28, 788–797. 52 Valincius G, Heinrich F, Budvytyte R, Vanderah DJ, McGillivray DJ, Sokolov Y, Hall JE & Losche M (2008) Soluble amyloid beta-oligomers affect dielectric membrane properties by bilayer insertion and domain formation: implications for cell toxicity. Biophys J 95, 4845–4861. 53 Kawahara M & Kuroda Y (2000) Molecular mecha- nism of neurodegeneration induced by Alzheimer’s beta-amyloid protein: channel formation and disruption of calcium homeostasis. Brain Res Bull 53, 389–397. 54 Soto C (2003) Unfolding the role of protein misfolding in neurodegenerative diseases. Nat Rev Neurosci 4, 49–60. 55 Lauren J, Gimbel DA, Nygaard HB, Gilbert JW & Strittmatter SM (2009) Cellular prion protein mediates impairment of synaptic plasticity by amyloid-beta oligo- mers. Nature 457, 1128–1132. 56 Mandal PK, Pettegrew JW, Masliah E, Hamilton RL & Mandal R (2006) Interaction between Abeta peptide and alpha synuclein: molecular mechanisms in overlap- ping pathology of Alzheimer’s and Parkinson’s in dementia with Lewy body disease. Neurochem Res 31, 1153–1162. 57 Masliah E, Rockenstein E, Veinbergs I, Sagara Y, Mallory M, Hashimoto M & Mucke L (2001) beta- amyloid peptides enhance alpha-synuclein accumulation and neuronal deficits in a transgenic mouse model linking Alzheimer’s disease and Parkinson’s disease. Proc Natl Acad Sci USA 98, 12245–12250. 58 Tsigelny IF, Crews L, Desplats P, Shaked GM, Sharikov Y, Mizuno H, Spencer B, Rockenstein E, Trejo M, Platoshyn O et al. (2008) Mechanisms of hybrid oligomer formation in the pathogenesis of combined Alzheimer’s and Parkinson’s diseases. PLoS ONE 3, e3135. 59 Kinoshita A, Fukumoto H, Shah T, Whelan CM, Iriz- arry MC & Hyman BT (2003) Demonstration by FRET of BACE interaction with the amyloid precursor protein at the cell surface and in early endosomes. J Cell Sci 116, 3339–3346. 60 Yan SD, Fu J, Soto C, Chen X, Zhu H, Al-Mohanna F, Collison K, Zhu A, Stern E, Saido T et al. (1997) An intracellular protein that binds amyloid-beta peptide and mediates neurotoxicity in Alzheimer’s disease. Nature 389, 689–695. 61 Almeida CG, Takahashi RH & Gouras GK (2006) Beta-amyloid accumulation impairs multivesicular body sorting by inhibiting the ubiquitin-proteasome system. J Neurosci 26, 4277–4288. 62 Oh S, Hong HS, Hwang E, Sim HJ, Lee W, Shin SJ & Mook-Jung I (2005) Amyloid peptide attenuates the proteasome activity in neuronal cells. Mech Ageing Dev 126, 1292–1299. 63 Tseng BP, Green KN, Chan JL, Blurton-Jones M & LaFerla FM (2008) Abeta inhibits the proteasome and enhances amyloid and tau accumulation. Neurobiol Aging 29, 1607–1618. 64 Clifford PM, Zarrabi S, Siu G, Kinsler KJ, Kosciuk MC, Venkataraman V, D’Andrea MR, Dinsmore S & Nagele RG (2007) Abeta peptides can enter the brain through a defective blood-brain barrier and bind selectively to neurons. Brain Res 1142, 223– 236. 65 Nixon RA (2006) Autophagy in neurodegenerative disease: friend, foe or turncoat? Trends Neurosci 29, 528–535. 66 Yang AJ, Chandswangbhuvana D, Margol L & Glabe CG (1998) Loss of endosomal ⁄ lysosomal membrane impermeability is an early event in amyloid Abeta1-42 pathogenesis. J Neurosci Res 52, 691–698. 67 Yuyama K, Yamamoto N & Yanagisawa K (2008) Accelerated release of exosome-associated GM1 ganglioside (GM1) by endocytic pathway abnormality: another putative pathway for GM1-induced amyloid fibril formation. J Neurochem 105, 217–224. 68 Sakono M, Zako T, Ueda H, Yohda M & Maeda M (2008) Formation of highly toxic soluble amyloid beta oligomers by the molecular chaperone prefoldin. Febs J 275, 5982–5993. 69 Hartl FU & Hayer-Hartl M (2002) Molecular chaperones in the cytosol: from nascent chain to folded protein. Science 295, 1852–1858. 70 Vainberg IE, Lewis SA, Rommelaere H, Ampe C, Vandekerckhove J, Klein HL & Cowan NJ (1998) M. Sakono and T. Zako Formation of toxic Ab oligomers FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 FEBS 1357 [...]... promotes the assembly of polyQ expansion proteins into nontoxic oligomers Mol Cell 23, 887–897 80 Zhang Y, McLaughlin R, Goodyer C & LeBlanc A (2002) Selective cytotoxicity of intracellular amyloid beta peptide1-42 through p53 and Bax in cultured primary human neurons J Cell Biol 156, 519–529 81 Chui DH, Dobo E, Makifuchi T, Akiyama H, Kawakatsu S, Petit A, Checler F, Araki W, Takahashi K & Tabira T (2001)... Localization of prefoldin interaction sites in the hyperthermophilic group II chaperonin and correlations between binding rate and protein transfer rate J Mol Biol 364, 110–120 Okochi M, Nomura T, Zako T, Arakawa T, Iizuka R, Ueda H, Funatsu T, Leroux M & Yohda M (2004) Kinetics and binding sites for interaction of the prefoldin with a group II chaperonin: contiguous non-native substrate and chaperonin.. .Formation of toxic Ab oligomers 71 72 73 74 75 76 77 M Sakono and T Zako Prefoldin, a chaperone that delivers unfolded proteins to cytosolic chaperonin Cell 93, 863–873 Zako T, Iizuka R, Okochi M, Nomura T, Ueno T, Tadakuma H, Yohda M & Funatsu T (2005) Facilitated release of substrate protein from prefoldin by chaperonin FEBS Lett 579,... Macario AJ & Conway de Macario E (2005) Sick chaperones, cellular stress, and disease N Engl J Med 353, 1489–1501 Muchowski PJ & Wacker JL (2005) Modulation of neurodegeneration by molecular chaperones Nat Rev Neurosci 6, 11–22 Powers ET, Morimoto RI, Dillin A, Kelly JW & Balch WE (2009) Biological and chemical approaches to diseases of proteostasis deficiency Annu Rev Biochem 78, 959–991 Tam S, Geller R,... Spiess C & Frydman J (2006) The chaperonin TRiC controls polyglutamine aggregation and toxicity through subunit-specific interactions Nat Cell Biol 8, 1155–1162 1358 78 Kitamura A, Kubota H, Pack CG, Matsumoto G, Hirayama S, Takahashi Y, Kimura H, Kinjo M, Morimoto RI & Nagata K (2006) Cytosolic chaperonin prevents polyglutamine toxicity with altering the aggregation state Nat Cell Biol 8, 1163–1170 79 Behrends... Takahashi K & Tabira T (2001) Apoptotic neurons in Alzheimer’s disease frequently show intracellular Abeta42 labeling J Alzheimers Dis 3, 231–239 82 Mousnier A, Kubat N, Massias-Simon A, Segeral E, Rain JC, Benarous R, Emiliani S & Dargemont C (2007) von Hippel Lindau binding protein 1-mediated degradation of integrase affects HIV-1 gene expression at a postintegration step Proc Natl Acad Sci USA 104,... affects HIV-1 gene expression at a postintegration step Proc Natl Acad Sci USA 104, 13615–13620 83 Sorgjerd K, Klingstedt T, Lindgren M, Kagedal K & Hammarstrom P (2008) Prefibrillar transthyretin oligomers and cold stored native tetrameric transthyretin are cytotoxic in cell culture Biochem Biophys Res Commun 377, 1072–1078 FEBS Journal 277 (2010) 1348–1358 ª 2010 The Authors Journal compilation ª 2010 . MINIREVIEW Amyloid oligomers: formation and toxicity of Ab oligomers Masafumi Sakono 1,2 and Tamotsu Zako 1 1 Bioengineering Laboratory, RIKEN Institute,. biological and structural characteristics of Ab oligomers and their formation mechanism remain unclear. Structure and size of soluble Ab oligomers Many types of

Ngày đăng: 06/03/2014, 09:22

Từ khóa liên quan

Tài liệu cùng người dùng

Tài liệu liên quan