Abstract
The aggregation of proteins into amyloid fibrils and their deposition into plaques and intracellular inclusions is the hallmark of amyloid disease. The accumulation and deposition of amyloid fibrils, collectively known as amyloidosis, is associated with many pathological conditions that can be associated with ageing, such as Alzheimer disease, Parkinson disease, type II diabetes and dialysis-related amyloidosis. However, elucidation of the atomic structure of amyloid fibrils formed from their intact protein precursors and how fibril formation relates to disease has remained elusive. Recent advances in structural biology techniques, including cryo-electron microscopy and solid-state NMR spectroscopy, have finally broken this impasse. The first near-atomic-resolution structures of amyloid fibrils formed in vitro, seeded from plaque material and analysed directly ex vivo are now available. The results reveal cross-β structures that are far more intricate than anticipated. Here, we describe these structures, highlighting their similarities and differences, and the basis for their toxicity. We discuss how amyloid structure may affect the ability of fibrils to spread to different sites in the cell and between organisms in a prion-like manner, along with their roles in disease. These molecular insights will aid in understanding the development and spread of amyloid diseases and are inspiring new strategies for therapeutic intervention.
Despite the first observation of amyloid deposits nearly four centuries ago1 (FIG. 1), it has been a long wait to see the structures of amyloid fibrils associated with devastating conditions such as Alzheimer disease (AD) and Parkinson disease (PD) in atomic detail. However, in the past 2 years fibril structures have been determined in near-atomic detail, thanks to developments in methods to create fibrils in vitro, or to purify them from ex vivo material, as well as breakthroughs in cryo-electron microscopy (cryo-EM) and solid-state NMR spectroscopy (ssNMR). Together, these methods have revealed the structures of amyloid fibrils implicated in some of the gravest of human diseases — amyloid-β (Aβ)40/42 and tau associated with AD2,3 and α-synuclein associated with PD4. Although, as expected, the fibrils adopt the canonical cross-β structure of amyloid (see below), these fibrils are more complex and elaborate than suspected previously, suggesting that a wide variety of amyloid structures exists, reminiscent perhaps of the different organization of α-helices and β-strands in globular proteins (1,375 distinct folds in globular proteins have been identified to date5). Such diversity may explain why amyloid diseases are so difficult to understand and to treat, with different clinical presentations, even when aggregation of the same protein is the culprit. Understanding the molecular architecture of amyloid fibrils may be an important step towards development of therapeutic interventions based on targeting the fibrils themselves or the processes that generate them.
Fig. 1. Progression of amyloid structure research over close to 400 years that has culminated in the first atomic structures of amyloid fibrils.
The timeline displays the history of key discoveries in the amyloid field from the initial identification of amyloid to discoveries that led to the first structures of amyloid fibrils associated with disease in all-atom detail288–296. Aβ, amyloid-β; β2m, β2-microglobulin; cryo-EM, cryo-electron microscopy; cryo-ET, cryo-electron tomography; EM, electron microscopy; micro-ED, micro-electron diffraction; polyA, poly-alanine; polyQ, poly-glutamine; ssNMR, solid-state NMR spectroscopy. Images reproduced with permission from REF.288, Cold Spring Harbor Laboratory Press; REF.289,290, Elsevier; REF.291, American Society for Clinical Investigation; reFs292,294, John Wiley and Sons; REF.293, American Chemical Society; REFS295,296, Springer Nature Limited.
Amyloid.
Fibrils formed from proteins, marked by a characteristic cross-β organization with an ~4.7–4.8 Å repeat running down the fibril axis.
Cross-β.
The structural motif consisting of β-strands organized perpendicular to the axis of a fibril and stabilized by inter-strand hydrogen bonds and dry steric zipper interfaces between adjacent β-sheets.
Approximately 50 different proteins or peptides are currently known to assemble into amyloid fibrils associated with human disease6 (TABLE 1). These precursors, each with a different primary sequence, self-assemble into amyloid fibrils that accumulate into extracellular plaques and intracellular inclusions associated with disease, which are especially prevalent during ageing7,8. Emerging evidence has shown that intracellular inclusions of amyloid can interfere with cellular physiology, for example, by disrupting transport of proteins and RNA9 and by sequestering chaperones and protea-somes10. Plaques from different phenotypic forms of the same disease can also accumulate different proteins11 and non-protein components12,13 to varying extents, potentially providing clues to the different phenotypes observed within an amyloid disease.
Table 1. Protein precursors associated with amyloid disease.
Disease | Aggregating protein or peptide | Number of amino acids | Native structure of protein or peptide |
---|---|---|---|
Neurodegenerative diseases | |||
Alzheimer disease | Amyloid-β peptide | 40 or 42 | Natively unfolded |
Familial encephalopathy with neuroserpin inclusion bodies | Neuroserpin | 410 | α + β |
Various neurodegenerative disorders | Actin | ~400 | Mostly α, some β |
Neuroferritinopathy | Ferritin | 175 or 183 | All α |
Spongiform encephalopathies | Prion protein or fragments thereof | 253 | Natively unfolded (residues 1–120) and α-helical (residues 121–230) |
Parkinson disease | α-Synuclein | 140 | Natively unfolded |
Dementia with Lewy bodies | α-Synuclein | 140 | Natively unfolded |
Frontotemporal dementia with Parkinsonism | Tau | 352–441 | Natively unfolded |
Amyotrophic lateral sclerosis | Superoxide dismutase | 153 | All β, immunoglobulin-like |
Huntington disease | Huntingtin with polyQ expansion | 3,144 | The polyQ-containing region is largely unstructured |
Spinocerebellar ataxias | Ataxins with polyQ expansion | 816 | All β, AXH domain (residues 562–694); the rest are unknown |
Spinocerebellar ataxia 17 | TATA box-binding protein with polyQ expansion | 339 | α + β, TBP-like (residues 159–339); unknown (residues 1–158) |
Spinal and bulbar muscular atrophy | Androgen receptor with polyQ expansion | 919 | All α, nuclear receptor ligand-binding domain (residues 669–919); the rest are unknown |
Hereditary dentatorubral- pallidoluysian atrophy | Atrophin 1 with polyQ expansion | 1,185 | Unknown |
Familial British dementia | ABri | 23 | Natively unfolded |
Familial Danish dementia | ADan | 23 | Natively unfolded |
Non-neuropathic systemic amyloidoses | |||
AL amyloidosis | Immunoglobulin light chains or fragments | ~90 | All β, immunoglobulin-like |
AH amyloidosis | Immunoglobulin heavy chains or fragments | ~220 | All β, immunoglobulin-like |
AA amyloidosis | Fragments of serum amyloid A protein | 76–104 | All α, unknown fold |
Familial Mediterranean fever | Fragments of serum amyloid A protein | 76–104 | All α, unknown fold |
Senile systemic amyloidosis | Wild-type transthyretin | 127 | All β, prealbumin-like |
Familial amyloidotic polyneuropathy | Mutants of transthyretin | 127 | All β, prealbumin-like |
Haemodialysis-related amyloidosis | β2-Microglobulin | 99 | All β, immunoglobulin-like |
ApoAI amyloidosis | N-terminal fragments of ApoAI | 80–93 | Natively unfolded |
ApoAII amyloidosis | N-terminal fragment of ApoAII | 98 | Unknown |
ApoAIV amyloidosis | N-terminal fragment of ApoAIV | ~70 | Unknown |
ApoCII amyloidosis | ApoCII | 79 | α + unstructured |
ApoCIII amyloidosis | ApoCIII | 79 | α + unstructured |
Finnish hereditary amyloidosis | Fragments of gelsolin mutants | 71 | Natively unfolded |
Lysozyme amyloidosis | Mutants of lysozyme | 130 | α + β, lysozyme fold |
Fibrinogen amyloidosis | Variants of fibrinogen a-chain | 27–81 | Unknown |
Icelandic hereditary cerebral amyloid angiopathy | Mutant of cystatin C | 120 | α + β, cystatin-like |
Non-neuropathic localized diseases | |||
Type II diabetes | Islet amyloid polypeptide (also known as amylin) | 37 | Natively unfolded |
Aortic media amyloidosis | Lactadherin C2-like domain | 50 | Unfolded |
LECT2 amyloidosis | Leukocyte cell-derived chemotaxin 2 | 151 | Unknown |
Localized cutaneous amyloidosis | Gelactin 7 | 136 | All β |
Hypotrichosis simplex of the scalp | Corneodesmosin | 529 (truncations cause amyloid) | Unknown |
Calcifying epithelial odontogenic tumours | Odontogenic ameloblast-associated protein | 153 | Unknown |
Senile seminal vesicle amyloidosis | Semenogelin 1 | 462 | Unknown |
Medullary carcinoma of the thyroid | Calcitonin | 32 | Natively unfolded |
Atrial amyloidosis | Atrial natriuretic factor | 28 | Natively unfolded |
Hereditary cerebral haemorrhage with amyloidosis | Mutants of amyloid-β peptide | 40 or 42 | Natively unfolded |
Pituitary prolactinoma | Prolactin | 199 | All α, four-helical cytokines |
Injection-localized amyloidosis | Insulin | 21 + 30 | All α, insulin-like |
Injection-localized amyloidosis | Enfuvirtide | 36 | Unstructured |
Aortic medial amyloidosis | Medin | 50 | Unknown |
Hereditary lattice corneal dystrophy | Mainly C-terminal fragments of kerato-epithelin | 50–200 | Unknown |
Corneal amyloidosis associated with trichiasis | Lactoferrin | 692 | α + β, periplasmic-binding protein like II |
Cataract | γ-Crystallins | Variable | All β, γ-crystallin like |
Calcifying epithelial odontogenic tumours | Unknown | ~46 | Unknown |
Pulmonary alveolar proteinosis | Lung surfactant protein C | 35 | Unknown |
Inclusion-body myositis | Amyloid-β peptide | 40 or 42 | Natively unfolded |
Cutaneous lichen amyloidosis | Keratins | Variable | Unknown |
Chaperones.
Proteins that assist in the folding, unfolding, assembly or disassembly of other macromolecular structures.
Amyloid fibrils share a common underlying architecture, in which the β-strands within each protofilament align perpendicular to the long axis of the fibril, termed a cross-β amyloid fold (FIG. 2) marked by a characteristic ~4.7–4.8 Å repeat running down the fibril axis. This structure has the strength of steel14,15 and, on the basis of its simplicity and ease of formation, has also been proposed as a potential primordial structure of life16. Amyloid can be either functional (in bacteria, fungi and higher eukaryotes)17–26 or disease-associated (TABLE 1), with both types of fibril sharing the canonical cross-β architecture unique to the amyloid fold.
Fig. 2. Schematic of amyloid formation.
Native proteins are in dynamic equilibrium with their less-structured, partially folded and/or unfolded states. One (or possibly several) of these states initiates amyloid fibril formation by assembling into oligomeric species. The precursor of aggregation (native, partially folded or unfolded) may differ for different protein sequences. Oligomeric species can then assemble further to form higher-order oligomers, one or more of which can form a fibril nucleus, which, by rapidly recruiting other monomers, can nucleate assembly into amyloid fibrils. This process occurs in the lag time (nucleation phase) of assembly. As fibrils grow, they can fragment, yielding more fibril ends that are capable of elongation by the addition of new aggregation-prone species140,144,145. This elongation results in an exponential growth of fibrillar material (blue line) until nearly all free monomer is converted into a fibrillar form. Fibrils are dynamic and can release oligomers that may or may not be toxic177. Fibrils can also associate further with each other, with other proteins and with non-proteinaceous factors188 (not shown here) to form the amyloid plaques and intracellular inclusions characteristic of amyloid disease. Note that any and/or all of these steps are potential points for drug intervention.
Protofilament.
A structural component of an amyloid fibril with a cross-β structure that twists together with one or more additional protofibrils to form a mature amyloid fibril.
Here, we first review the history of our understanding of the amyloid fold, reporting a timeline of discovery from the 17th century to the present day (FIG. 1). We then describe our current understanding of amyloid fibril structure, inspired by recent reports of fibrils formed from proteins associated with human disease27,28. We also discuss how amyloid forms and the relationship between protein sequence, fibril morphology and plaque formation, including how these affect disease presentation and progression. Finally, despite recent high-profile setbacks in developing drugs to treat amyloid diseases29–32, we speculate on how this new, atomprecise vision of amyloid may transform our efforts to treat disease.
Fibril formation and disease
Our understanding of how amyloid fibrils relate to their associated diseases has expanded rapidly in recent years, but from the initial identification of amyloid to the atomic models we have today has been a journey over nearly 400 years.
A timeline of amyloid discovery
First described in 1639 as lardaceous liver and ‘white stone’-containing spleen1, the term amyloid (derived from amylum and amylon, the Latin and Greek words for starch, respectively), was coined ~200 years later by Virchow on the basis of his discovery that these deposits stained positively with iodine1,33. Just 5 years after its misidentification as a polysaccharide, Friedrich and Kekulé showed that amyloid is predominantly proteinaceous1, with carbohydrates, specifically glycosaminoglycans, being ubiquitously associated with these deposits34. Developments in light microscopy, combined with the finding that amyloid deposits show the unique tinctorial property of red–green birefringence in the presence of Congo red35, revealed that amyloid is formed of highly organized protein subunits. The identification of amyloid A36 (in 1971), antibody light chains37 (in 1971) and transthyretin (TTR)38 (in 1978) as the major protein component of fibrils in plaques showed that an individual protein precursor can be responsible for an amyloid disease. In the meantime, Bill Astbury’s pioneering X-ray fibre diffraction studies had shown that amyloid-like fibrils could be created from normally globular, soluble proteins by denaturing them in vitro39, opening the door to synthetic materials that Astbury hoped would replace wool40. In 1968, Sandy Geddes, Bill Aiken and colleagues, working in Astbury’s Department of Biophysics at the University of Leeds, UK, used X-ray fibre diffraction to reveal that the egg stalk of the lacewing fly also has a distinct ~4.7 Å repeating feature down the fibril axis, which they named cross-β41. This ground-breaking work established a structural fingerprint for amyloid and showed that amyloid is not simply associated with disease (TABLE 1) but can also perform physiological functions in an organism42–47.
Subunits.
The smallest units that make up an amyloid fibril, generally single copies of the precursor protein.
In the 50 years since Geddes et al. coined the term cross-β41, information about the structure of amyloid fibrils has been drip-fed to the scientific community, with tantalizing insights coming from studies of fibrils formed from short peptides using X-ray fibre diffraction48, X-ray crystallography49–51, micro-electron diffraction (micro-ED)52 of microcrystals53–56, ssNMR57–66, cryo-EM27,28,67–72 and other methods (BOX 1). The ability to assemble fibrils in vitro from synthetic peptides73–75 and naturally occurring amyloidogenic proteins and peptides76–78, or from proteins not associated with functional amyloid or disease68,79, has fuelled developments from each structural technique. However, some of the most recent and exciting breakthroughs have been enabled by the step change in the resolving power of cryo-EM that is currently revolutionizing structural biology80 (BOX 2). Combining these techniques has finally cracked the amyloid fold, revealing structures that are as beautiful and intricate as those of their globular counterparts27,28. These structures give us the first glimpse of how individual protein subunits form cross-β structure, and by combining these data with insights from super-resolution microscopy81,82, cryo-electron tomography (cryo-ET)13,83,84 and ssNMR85, we are entering a new era of integrative structural biology that allows us to ‘see’ the structure of amyloid fibrils on multiple scales, from individual subunits and how they form fibrils to the cellular consequences of fibril deposition into plaques and tangles.
Box 1. Techniques that inform on amyloid structure.
Electron microscopy (em) is by no means the only technique that has driven the understanding of amyloid fibril structure in recent years. A variety of biophysical techniques have been utilized to make advances in understanding fibril structure.
Atomic force microscopy
AFm has been used to examine fibrils adsorbed onto a surface and can report on length distribution, fibril height, width and crossover length. AFm has also been used to track the dynamics of fibrils275 and oligomers276 over time.
Solid-state NMR spectroscopy
ssNmr (also referred to as magic angle spinning (mAS) Nmr) can be used to determine dihedral angles and inter-atom distances in the subunit of fibrils. These restraints have been used to build atomic models of fibrils65,66,216,251,277,278 as well as models of the subunit structure, which were then fit into lower-resolution em models64,67.
Fluorescent dyes
Fluorescent dyes such as thioflavin T and congo red have a long history of use identifying amyloid fibrils35. more recently, polythiophene dyes have also been shown to have distinct emission wavelengths when bound to different fibril morphologies and can be used to differentiate fibril types279.
Super-resolution microscopy
Super-resolution microscopy with labelled proteins has been used for high-resolution light microscopy imaging280 and to observe the addition of monomers to fibril ends, yielding information about fibril polarity and growth81.
Single-molecule Förster resonance energy transfer
FreT experiments have also been used to track fibril growth and characterize amyloid oligomers172.
X-ray fibre diffraction
one of the earliest methods used to identify the cross-β motif41, it is still used, especially as an orthogonal validation technique, as simulated fibre diffraction patterns from a model can be compared with experimental data281.
X-ray crystallography and micro-electron diffraction
Two methods that have been especially useful to determine the structures of small amyloidogenic peptides; although these peptides usually do not crystallize as fibrils, they have been used to predict the interactions present in the fibrillar form51,55, to link structure to toxicity54 and to design inhibitors of fibril formation272.
Spectroscopy
methods such as circular dichroism (cd) and Fourier transform infrared (FTIr) can be used to characterize the secondary structure of fibril-forming proteins over time, to track the formation of oligomers and/or fibrils and, in the case of FTIr, to differentiate amyloid from non-amyloid β-sheets201.
Mass spectrometry
A variety of mass spectrometry (mS) methods have been used to characterize fibrils, including interrogating exposed surfaces via N-ethylmaleimide labelling coupled with electrospray ionization mS (eSI-mS)282 and hydrogen-deuterium exchange (HdX)283.
Electron paramagnetic resonance
ePr spectroscopy has been used to identify residues involved in the cross-β core of amyloid284 and to characterize amyloid precursors, oligomers and membrane interactions285.
Box 2. Advances in electron microscopy of amyloid fibrils.
Although the use of electron microscopy to examine amyloid fibril structure has a long history209,286,287, determination of atomic resolution structures became a possibility only with the advent of cryo-electron microscopy (cryo-em). cryo-em led to observation of several fibril structures at intermediate resolutions68,71,72,210,211,229. Although this gave information about gross fibril morphology, it could not be used to build atomic models. only in the past year have cryo-em structures reached the resolution at which atomic detail can be resolved27,28. This advance in resolution has been driven by technological advances in microscope and electron detection equipment and improvements in image processing software80. The increase in achievable resolution of cryo-em reconstructions of fibrils is illustrated with structures of fibrils formed from SH3 domains68, prion protein (PrP)210, amyloid-β (Aβ) 1–40 formed in vitro71 and ex vivo tau28 (see the figure, parts a–d).
Part a adapted with permission from REF.68, elsevier. Part b adapted with permission from REF.210, elsevier. Part c adapted with permission from REF.71 copyright (2008) National Academy of Sciences uSA. Part d adapted from REF.28, Springer Nature limited.
Diseases associated with amyloid fibril deposition
Alois Alzheimer reported the first documented case of AD in his patient, Auguste Deter, whom he first encountered in 1901 (REFS86,87) (FIG. 1). This work, which included the post-mortem visualization of Congo-red-positive plaques in the brain, added AD to the growing list of amyloidoses. Today, more than 50 disease-causing amyloidogenic proteins have been identified, which give rise to an even greater number of diseases depending on the sequence of the precursor and the site of amyloid deposition88–90 (TABLE 1). These diseases include neurodegenerative disorders such as AD (involving aggregation of Aβ91 and/or tau2), Creutzfeldt−Jakob disease (CJD; prion protein (PrP))92, Huntington disease (huntingtin93), PD (α-synuclein94) and amyotrophic lateral sclerosis (ALS; superoxide dismutase (SOD), TAR DNA-binding protein 43 (TDP43) and others95). Amyloid disorders affect other tissues: type II diabetes involves the aggregation of amylin (otherwise known as islet amyloid polypeptide (IAPP)96) in the islets of Langerhans; in AL amyloidosis, antibody light chains are deposited in the kidney and heart97; and in dialysis-related amyloidosis, β2-microglobulin (β2m) forms amyloid plaques in the osteoarticular tissues98. There is also crosstalk between amyloid diseases. For example, patients with type II diabetes have a higher risk of AD99, and the non-amyloid component (NAC) of α-synuclein has been found in the plaques of patients with AD100.
Amyloidoses.
A class of diseases associated with the formation of amyloid fibrils, tangles and plaques, although the causative agents of disease have yet to be determined definitively.
Prion.
A class of infectious amyloid fibrils.
What initiates the onset of amyloid disease remains unclear. Many of the diseases are associated with ageing and involve the aggregation of wild-type proteins, including most cases of PD and AD8. Owing to the ageing of the human population, the estimated economic burden in Europe from AD and PD alone is estimated to rise to€357 billion by 2050 (REF.101), comparable to the gross domestic product (GDP) of Austria in 2016. Mutations in amyloidogenic precursors, such as the α-synuclein variants A30P or A53T in PD102, the Aβ variants E22Δ or E22K in AD103 or D76N in β2m-associated amyloidosis104, can cause diseases to present earlier (TABLE 1). Other disorders are caused by the expansion of an amyloidogenic sequence, such as the trinucleotide repeat diseases, which result in poly-glutamine (polyQ)-associated ataxias such as Huntington disease93,105, poly-alanine (polyA) expansions in polyadenylate-binding protein 2 (PABPN1) (REF.106) associated with oculopharyngeal muscular dystrophy107 and dipeptide expansions such as poly-GlyAla in C9orf72, the most common genetic form of ALS and frontotemporal dementia108,109. The age of onset for people with these diseases is variable, as a critical threshold number of repeats determines their pathogenicity, and the rate of disease onset correlates with the number of repeats105,110.
Age of onset.
The age at which a patient first presents symptoms. For amyloid-associated disorders, this is not necessarily directly correlated with fibril load: high fibril loads may be asymptomatic, whereas low fibril loads may lead to severe symptoms.
Modification of a precursor’s primary sequence (for example, truncations by the action of proteases or hyperphosphorylation) can enhance or suppress amyloidogenicity. Trimming the precursor protein can reduce amyloidogenicity, evidenced by the lower risk of AD in patients with an increased ratio of Aβ40 to Aβ42 (REF.111). By contrast, proteolysis of apolipoprotein A1 (APOA1) increases the risk of amyloidosis in systemic amyloidosis112 (TABLE 1), and truncation of the amino-terminal six amino acids of β2m is observed in dialysis-related amyloidosis113,114 (TABLE 1). Simply increasing the concentration of the monomeric precursor by gene duplication can also cause disease, with gene duplication, triplication or even quadruplication associated with the early onset of PD94,115, and patients with trisomy of chromosome 21, which contains the gene encoding amyloid precursor protein (APP; from which Aβ40/42 are derived), have a higher risk of AD116. The inability of haemodialysis to remove wild-type β2m from the serum of patients with renal failure causes an ~50-fold increase in β2m concentration in the serum117. Liquid phase separation118 may also lead to local increases in concentration leading to fibril formation119. Intrinsically disordered proteins have a higher propensity towards coacervation, which may explain why many are amyloidogenic, and cross-β polymerization has been suggested as a driving force in the formation of liquid droplets in vivo120, which can result in dialysis-related amyloidosis in patients undergoing long-term (>10 years) renal replacement therapy117. Finally, amyloidogenicity can be modulated by small molecules such as metabolites or metal ions, by membranes and glycosaminoglycans and by the status of the cell itself (chaperone levels, rate of protein synthesis, etc.), which affect the rate of protein aggregation and the ability of the cell to respond to the formation of potentially toxic species9,121.
Haemodialysis.
A dialysis-based filtration treatment that acts to replace kidney function in patients experiencing kidney failure.
Phase separation.
A process driven by liquid– liquid demixing, leading to a liquid mixture separating into individual components. in cells, this can lead to localized increased concentration and supersaturation of biological molecules.
Intrinsically disordered proteins.
Proteins that lack a fixed or ordered 3D structure.
Prion aggregation and disease
The ability of amyloid to seed its own assembly, causing disease to spread between cells and, in some cases, between organisms, is a hallmark of prion diseases122,123. In mammals, PrP causes a family of diseases known as the transmissible spongiform encephalopathies92. The soluble α-helical protein, PrPC, can conformationally rearrange to an infectious isoform, PrPSc, which has β-sheet structure (a process that is not yet understood in atomic detail). As an infectious agent, PrPSc is believed to be devoid of nucleic acid124, although this is contested by some studies125. There are many prion diseases in humans, such as CJD, familial fatal insomnia, Gerstmann–Sträussler–Scheinker disease and Kuru, which result in different diseases characterized by different age of onset126. Bovine spongiform encephalopathy (BSE) in cattle, chronic wasting disease in deer and elk and scrapie in sheep and goats127 have also been identified as transmissible spongiform encephalopathies. BSE can cross the species barrier to humans, leading to variant CJD (vCJD)123,124.
Although prion diseases are predominantly neurological, in vCJD, PrPSc can be found in different tissue types128. Prion-like spreading of disease within the brain is not unique to PrP; it is also observed in both PD129 and AD130, suggesting that common mechanisms exist by which protein aggregation spreads and causes cell death. In yeast, prion-like spreading of protein aggregates, such as those formed from the proteins Ure2p and Sup35, can be beneficial, endowing metabolic advantages to daughter cells under certain selective pressures131. Indeed, the large number of prion-like domains identified in yeast132 suggest that such selective advantages are commonly adopted, giving enhanced fitness. Despite the prion-like seeding ability of amyloid fibrils, other than in the prionopathies, there is no direct evidence for the spread of amyloid diseases between mammals. There is, however, some evidence for transmissibility of amyloid disease in an experimental setting133,134. Thus, it would be wise to treat amyloidogenic proteins and peptides as potential infectious agents.
Mechanisms of fibril and plaque formation
Amyloid is formed by the aggregation of monomeric protein precursors into fibrils by a common nucleation growth mechanism135,136 (FIG. 2). Monomeric precursors may be unfolded (intrinsically disordered) or partially folded (formed by transient unfolding of a native protein or transient folding of an unfolded protein) (TABLE 1). In rare cases, aggregation may be initiated by the native protein itself137. The first step in fibril assembly involves the formation of oligomers, which are dynamic, transient, heterogeneous and of unknown and possibly varied structure136,138,139. Oligomers can then further associate to produce higher-order species, which can be either essential precursors of amyloid fibrils (on-pathway) or dead-end assemblies that do not produce fibrils (off-pathway)136. Although off-pathway oligomers do not go on to form fibrils, they may still be cytotoxic and relevant to disease. At some stage during oligomerization, a critical nucleus is formed, which is defined kinetically as the most unstable (highest energy) species formed before rapid polymerization into amyloid fibrils. The probability of nucleus formation determines (in part) the length of the lag time of amyloid assembly and possibly the age of disease onset (FIG. 2).
Native protein.
The properly assembled form of a protein required for functionality.
At some point during self-assembly, each precursor undergoes a structural transformation to form β-strand-rich secondary structure, irrespective of its initial fold (TABLE 1). Once fibrils with a cross-β structure form, they can fragment, producing new fibril ends that can recruit monomers, reducing the length of the lag time and resulting in exponential fibril growth140 (the elongation phase shown in FIG. 2). Other processes, such as secondary nucleation, in which oligomer formation is catalysed on the surface of a pre-existing fibril, also enhance the rate of fibril formation136,141. Understanding each process, and how they combine to determine the rate of fibril assembly (and hence fibril load), is vital for elucidation of the mechanism of fibril formation in both in vitro and in vivo environments.
Fibril load.
A measure of the total amount of amyloid fibril within a sample or patient.
Various models, inspired by work on sickle cell disease by Eaton, Oosawa and colleagues142,143, including both numerical140,144 and analytical approaches145, allow the individual steps of amyloid formation to be kinetically defined by employing simple-to-use algorithms now available online140. The kinetics of assembly are commonly measured in vitro using the dye thioflavin T (ThT), which binds to amyloid fibrils, generating an enhanced fluorescence signal145,146. Using this approach, the effect of synthetic membranes147, solution conditions148, molecular chaperones149, small molecules150 and the primary sequence141 on the rate of aggregation can be elucidated. Most importantly, the ability to determine the role of different additives on assembly offers the opportunity to control aggregation by the synergistic application of reagents that target different steps in amyloid formation.
Cytotoxicity and implications for disease
Identifying the toxic species formed during amyloid formation, and how they cause cellular dysfunction and death, remains both a challenge and a priority (FIG. 3). Questions abound, including how aggregation is initiated, how aggregates are recognized by chaperones and other cellular components and how and why aggregates saturate the cellular chaperone and degradation networks151–153. Models of amyloid-associated cytotoxicity include inhibition of proteasomal degradation10,151, impairment of autophagy154, perturbation of mitochondrial function155, production of reactive oxygen species (ROS)156, sequestration of other proteins121 and disruption of membranes, including mitochondria, the endoplasmic reticulum (ER), lysosomes and plasma membrane84,121,157,158 (FIG. 3).
Fig. 3. Amyloid aggregates can cause cell disruption by a variety of mechanisms.
Amyloid aggregates can deposit extracellularly or intracellularly, and both can give rise to cellular dysfunction and disease. The aggregates that form from different protein precursors may have different cellular effects, but deconvoluting the toxic mechanism of an individual protein and its ensemble of misfolded or aggregated states (misfolded monomers, oligomers, fibrils or plaques and intracellular inclusions) remains a challenge. Plaques and inclusions sequester a range of other molecules that include glycosaminoglycans12,188, lipids158,297 and metal ions237, which stabilize their assembly. Plaques are physically large and can disrupt organ function by their sheer size. Small fibrils can also be taken up into a cell via endocytosis, but this can be perturbed by preventing binding to certain cell surface receptors such as lymphocyte activation gene 3 protein267. Within endosomes and lysosomes, fibrils can release toxic oligomers and can disrupt the endosomal and lysosomal function and dynamics because fibrils are highly resilient to degradation190,298. Fibrils can also access the intracellular space following release from cells, thus spreading disease by uptake into adjacent cells. Other effects of aggregates within cells include disruption of endoplasmic reticulum (ER) dynamics84, release of reactive oxygen species (ROS)156 from mitochondria and the induction of stress responses121 (not shown here).
Crucially, the severity of cognitive decline in patients with AD does not correlate with plaque formation159, suggesting that pre-amyloid aggregates are the cause of disease138,139,160–164. Consistent with this view, numerous experiments in vitro have demonstrated the cytotoxicity of oligomeric species, including their ability to disrupt membranes165–167. Oligomers formed from proteins not normally associated with disease can also disrupt membranes and can be cytotoxic, adding weight to the view that oligomers are the causative agents of amyloid-associated cellular dysfunction162,166–168. Oligomers formed from disease-related precursors have also been shown to impair memory and long-term potentiation, again supporting their role in disease169,170. However, not all oligomers are toxic171,172. What is known is that toxic oligomers expose hydrophobic surfaces not found in innocuous precursors or non-toxic counterparts173, consistent with their ability to perturb membranes and expose the cytoplasm to the extracellular space, causing calcium flux and ultimately cell death164,174. The molecular basis for an oligomer’s cytotoxicity will remain unclear until a high-resolution structure of a toxic oligomer is solved. However, because oligomers are unstable, dynamic and heterogeneous in mass and structure175,176, determining their structure–function relationships is challenging.
Long-term potentiation.
A persistent increase in synaptic strength after stimulation of the synapse.
Recent experiments have re-energized the debate about how, or whether, amyloid fibrils contribute to disease177. Using cryo-ET, amyloid fibrils and, in particular, their ends have been shown to perturb artificial lipid membranes (inducing breaks, blebbing and formation of pinched sharp points with high membrane curvature)178 and to cause pinching of cellular membranes84, suggesting a role for fibrils in disease. Cryo-ET also showed that intracellular inclusions, formed from exon-1-encoded huntingtin, which contains 97 glutamines, localize to the rough ER, perturbing ER function and dynamics84. Interestingly, extracellular assemblies of Aβ42 fibrils in cell culture take distinct forms, including meshworks, semi-parallel bundles and ‘stars’ radiating out from a central point13,83. These extracellular fibrils also interact with membranes, sequestering lipids and forming tubular inclusions at the cell surface83. Intracellular inclusions formed by ATG-independent translation of a non-coding region of the ALS-related gene C9orf72 sequester proteasomes and impair proteasome activity10, demonstrating that fibrillar assemblies from different proteins have different cellular effects. In support of this view, NMR metabolomics showed that monomeric, oligomeric and fibrillar α-synuclein and Aβ40/42 have different meta-bolic effects in neuroblastoma cells, suggesting that cells attempt to counter the toxicity imposed by pre-fibrillar species, whereas fibrils led to cellular shutdown179. Recent experiments have also shown that different fibril morphologies, even when formed from the same protein, can cause different cellular effects, presumably by binding to different molecules and/or by depositing in different locations60,180–186. This is illustrated by Aβ40 fibrils produced in vitro under different conditions. So-called 2A fibrils have two-fold symmetry and were made by seeding from Aβ40 fibrils produced under agitation, whereas fibrils formed from the same protein with three-fold symmetry (named 3Q) were seeded from fibrils produced under quiescent conditions187. These Aβ40 fibrils bind the glycosaminoglycan, heparin, with different affinities12,188, whereas other fibrils bind to specific RNA molecules or other proteins121. Finally, the role of fibril structure in disease is being studied. For example, fibrils from brain extracts of patients with AD show distinct size, concentration and conformational characteristics, which correlate with disease duration and severity189. This difference in fibril characteristics may in part explain why amyloid fibril load does not always correlate with the severity of disease, as the biological effects and clinical symptoms may depend on which fibril polymorphs are present within amyloid deposits60.
Although it is fascinating to dissect the effects of different oligomers and fibrils in cellular dysfunction, it is likely that a combination of species will correlate with disease. Amyloid formation is a dynamic process, with monomers and oligomers in rapid exchange with each other148. Oligomers can also be generated directly, by loss of monomers and/or oligomers from fibril ends190,191 (FIG. 2). This phenomenon may be enhanced by the cellular environment, such as the low pH of endosomes and lysosomes190. The protein concentration in cells is also finely tuned, with tight coupling of the rates of protein synthesis and degradation to keep proteins within their solubility limit192,193. Given this balance, it is perhaps not surprising that overproduction of a protein can trigger wholesale aggregation of susceptible proteins that are at the cusp of their solubility in the cell194–196. Precursor mutation, changes in post-translational modifications, stress or ageing can also result in a breakdown of these usually highly protective networks, leading to aggregation.
What emerges from a consideration of this complex network of interactions is the need to elucidate the structures of protein aggregates at atomic resolution. This knowledge would help us to understand how these aggregates perturb cells in molecular detail and to identify methods and molecules that can help control protein aggregation, with beneficial effects on cellular dysfunction and disease. In addition, we also need to explore why assembly of functional amyloid does not result in cell death. Although a number of mechanisms may be employed to prevent toxicity, such as sequestering functional amyloids within membrane-bound compartments42, the comparison of the structures of functional and disease-associated amyloid fibrils and their assembly mechanisms will be required to reveal how amyloid toxicity can be avoided when fibril assembly is beneficial to the organism.
Diversity of amyloid structures at high resolution
Amyloid fibrils are held together by a large variety of inter-monomer and inter-fibril interactions. The particular modes of interaction vary in different fibril structures formed from the same protein, as well as fibrils formed from different proteins. These interactions affect the physical properties of the fibril assemblies, which may contribute to the phenotypic effects observed for different fibril polymorphs.
Amyloid fibril structure at the subunit level
The defining structural feature of amyloid fibrils is the cross-β fold that all fibrils share. This ‘amyloid fold’ involves a ladder of stacked β-strands oriented perpendicular to the fibril axis, with each ‘rung’ of the cross-β ladder separated by a 4.7–4.8 Å spacing that arises from the regular hydrogen bond distance between paired carbonyl and amide groups in adjacent β-strands (FIG. 4). This spacing was first demonstrated in 1968–1969 using X-ray fibre diffraction41,197 (FIG. 1) and is found in all amyloid fibrils irrespective of the sequence of the protein precursor. The presence of the cross-β conformation has now been verified as ubiquitous in amyloid fibrils, whether functional or disease-related, using X-ray fibre diffraction198, X-ray crystallography63, cryo-EM199 and Fourier transform infrared spectroscopy (FTIR)200. In FTIR, the strong hydrogen bonds between adjacent β-strands in the cross-β fold absorb at a characteristic frequency of ~1,618 cm−1, whereas the more twisted, less stable β-sheets in globular proteins absorb at longer wavelengths201,202.
Fig. 4. Structural motifs that stabilize amyloid fibrils.
a | A ten-residue peptide from transthyretin (TTR), showing β-sheet stacking in which each β-strand ‘rung’ is stabilized by hydrogen bonds (denoted by fine black dotted lines) between the polypeptide backbones of precursors, which are separated by the canonical 4.7–4.8 Å repeat of the cross-β amyloid fold (PDB accession number 2nm5 (REF.67)). Further stabilization is provided by a steric zipper between the β-sheets, which stabilizes the fibril core. b | The β-helix of HET-S illustrating its steric zippers (PDB accession number 2rnm (REF.66)). c | A structure of amyloid-β (Aβ)42 fibrils (PDB accession number 5oqv (REF.27)) illustrating the variety of interactions that stabilize the fibril, including β-strand stacking (top left), formation of inter-protofilament salt bridges (top right), intra-protofilament steric zippers (bottom left) and inter-protofilament steric zippers (bottom right).
It is important to note that there are several fibrillar structures that are not considered to be amyloid despite sharing some traits. For example, small aromatic molecules (such as diphenylalanine203) assemble into fibril-like structures stacked with a 3.4 Å spacing and stabilized by pi-stacking interactions204. Because these structures are not composed of protein, they are generally not considered amyloid, despite their structural similarity and the fact that they are also deleterious to cells204, although there are no reports of disease caused by such ‘chemi-fibrils’. Other proteinaceous assemblies have features reminiscent of amyloid but violate some of the defining characteristics of the amyloid fold. For example, short helix–turn–helix peptides assemble into twisted fibrils, stabilized by stacked α-helices205. Fibrils in which α-helices, rather than β-strands, orient perpendicular to the fibril axis (so-called cross-α) are also found in vivo, where they are important for biofilm formation in Gram-positive organisms206. For example, a toxic oligomer of SOD1, associated with ALS (TABLE 1), shares many structural and pathological characteristics with amyloid but is not considered as a canonical amyloid as its β-sheets are arranged in a corkscrew stack at an ~45° angle to the fibril axis, akin to an extended β-barrel207. A similar structure has been observed in an oligomer called cylindrin, assembled from a β-hairpin peptide derived from αβ-crystallin208. Interestingly, this cylindrin structure is toxic to HeLa and HEK293 cells208.
Although the 4.7–4.8 Å stacked β-strand motif is the signature of amyloid, some amyloid fibrils also share features at a larger length scale. Some of the earliest EM observations of fibrils isolated from patient spleens described ‘beads’ with an ~100 Å periodicity down the fibril axis209. Later cryo-EM studies of fibrils formed in vitro from PrP and β2m reported repeats of ~60 Å (REF.210) or 52.5 Å (REF.211), respectively, in addition to the canonical ~4.7 Å cross-β repeat, suggestive of higher-order repeating structures. These longer repeats have not been observed in the high-resolution structures of tau or Aβ42 fibrils reported recently27,28, leaving the structural basis of these larger repeats open to debate.
Pi-stacking interactions.
Attractive, noncovalent interactions between aromatic rings (phenylalanine, Tyr and Trp in proteins).
Biofilm.
A group of microorganisms that have adhered to each other and/or a surface.
Gram-positive organisms.
Bacteria that possess a peptidoglycan-containing cell wall, which can be positively stained with crystal violet dye, known as gram stain.
Other common features are found in the subunit structures of amyloid fibrils regardless of the sequence and structure of their precursor (TABLE 1). In all cases, the subunit structure adopted within the fibril is dramatically different from that of the native protein (TABLE 1). Thus, a major structural conversion must occur as amyloid forms: from unfolded to β-strand, α-helical to β-strand or reorganization of pre-existing β-sheet structures (TABLE 1). Despite the commonality of their cross-β fold, fibril structures are not identical, with recent studies revealing a remarkable diversity of architectures that all conform to the amyloid fold (FIG. 5). A ten-residue peptide from TTR67 forms antiparallel β-strand pairs, with each peptide forming one rung of the cross-β ladder, whereas Aβ40 forms distinct fibril structures that have an in-register parallel organization of their β-strands but differ in the precise location and/or organization of their β-loop β-motif 212,213(FIG. 5c). In larger proteins, multiple sets of antiparallel β-sheets, termed super-pleated β-sheets, were predicted for a Ure2p prion filament214 and IAPP fibrils215, whereas a more complex arrangement in which the β-strands form so-called Leu-Ser motifs that then stack in a parallel in-resister arrangement on the fibril long axis were observed in a structure of a fibril formed by Aβ42 (FIG. 5b, left)27,65. Other fibril structures with complex organization of the β-strands within each rung of the cross-β ladder have been observed, including those of α-synuclein55,216 (FIG. 5e) and tau28 (FIG. 5d). What is clear is that very different organizations of polypeptide chains can stack into a fibril that conforms to the cross-β structure of amyloid. As more atomic structures of amyloid fibrils are solved, it will be interesting to see just how many different organizations of β-strands will fall under the umbrella of the cross-β fold.
Fig. 5. Subunit packing in amyloid fibrils.
Space-filling representations of near-atomic-resolution models of different amyloid fibrils, each filtered to 4 Å. Individual subunits are coloured in red to highlight different inter-protofilament packing in different fibril types. a | The β-helix of HET-S that forms a single filament (PDB accession number 2lbu (REF.66)). b | Two polymorphs of amyloid-β (Aβ)42 fibrils formed under different growth conditions (PDB accession number 5oqv27 (left) and PDB accession number 5kk3 (REF.278) (right)). c | Two polymorphs of Aβ40. Fibrils formed under the same solution conditions but propagated from seeds with different morphologies (2A, PDB accession number 2lmn (REF.212) (left) and 3Q, PDB accession number 2lmp (REF.212) (right)). d | Two polymorphs of tau fibrils: paired helical (PHF) (left) (PDB accession number 5o3l (REF.28) and straight (SF) (right) (PDB accession number 5o3t (REF.28)). e | The single filament of α-synuclein fibrils (PDB accession number 2n0a216). The main chain of the top layer of polypeptide chain in each fibril is shown in red. ‘2A’ indicates fibrils with two-fold symmetry; ‘3Q’ indicates fibrils with three-fold symmetry.
In all of the amyloid fibril structures determined to date with atomic precision, the β-strands are stabilized by dry ‘steric zippers’51 (tight interfaces of interdigitated hydrophobic side chains that exclude water (FIGS 4,5)) and almost perfect packing of their amino acid side chains. These zippers appear to be unique to amyloid fibrils, as they have not yet been observed in a globular protein or in other natural fibrous proteins51. Eisenberg’s original description of the steric zipper interface was based on structures from microcrystals of short peptides derived from amyloid-forming proteins, including fragments of Aβ, tau, PrP, insulin, IAPP, lysozyme, β2m and α-synuclein51. This work described eight classes of zipper — four formed from parallel β-strands and four from antiparallel β-strands51. Five of the eight proposed structures were observed experimentally51. Microcrystals have similarities with fibrils in that they can grow under similar conditions, have an ~4.7 Å repeat in their unit cells217 and, in some cases, small changes in conditions drive interconversion of fibrils into crystals and vice versa218. Peptide microcrystals can also seed fibrils of similar morphology from the full-length protein54, suggesting that steric zippers are the ‘core’ of amyloid fibrils formed by their intact protein counterparts54,55. However, the steric zippers identified in fragments of tau53 did not form similar zippers in the cryo-EM structure of fibrils formed from the full-length protein28, although this does not preclude formation of zippers by these regions in tau fibrils with different morphologies.
Proteins with a cross-β structure are also found in β-helices219,220. These structures are stabilized by a specific pattern of hydrophilic side chains on the outside of the β-helix and hydrophobic side chains on the inside, along with a terminal glycine220 (FIG. 4b). β-Helices have been formed from designed peptides221, are found in soluble proteins (including a variety of bacterial lyases222, antifreeze proteins223 and viral tail spikes224) and have been implicated in the formation and prionic nature of fibrils in fungi and other organisms28,198,225. High-resolution structures of tau fibrils determined using cryo-EM28 and HET-S from ssNMR66,226 show β-helix motifs with very similar backbones (root-mean-square deviation (r.m.s.d.) 1.3 Å) despite low sequence similarity in their β-helical regions. Together, the results highlight the array of possible structures that conform to the general amyloid fold but that differ in the details of how their β-strands are arranged.
Filament architecture of the amyloid fibril
The morphology of an amyloid fibril is determined by the number and arrangement of protofilaments that form the fully assembled fibrils28,64,70–72,211,227 as well as the structure of the subunit itself. Variation in the arrangement and type of interactions between protofilaments adds to the diversity of the amyloid fold. Some amyloid fibrils comprise a single protofilament (for example, the β-helix of HET-S66 (FIGS 4b,5a)), but the majority contain multiple protofilaments27,28,64,199,227 that twist together68,228–230. Different twists are observed between fibrils formed in the same growth mixture70,211,229 and even within a single fibril. Other arrangements of protofilaments have been found, including cylinders231, flat ribbons64,227,230 and pseudo-crystalline sheets218. Protofilaments in twisted fibrils have been found with a variety of symmetries viewed down the fibril axis as well as pairing in an asymmetrical manner28. These symmetries include two or three monomeric units in the same plane (in-register two-fold and three-fold)212,213 and 21 screw symmetries27,28 (a rotation of 180° followed by a translation of half of the β-sheet spacing down the fibril axis). In the recent high-resolution cryo-EM structures of tau and Aβ42, the protofilaments pack in a parallel manner, giving the fibril polarity27,28. By contrast, lower-resolution cryo-EM reconstructions of β2m fibrils suggest that fibrils constructed from both nonpolar (antiparallel) and polar (parallel) arrangements of protofilaments are formed within the same preparation211. However, it should be noted that no high-resolution structure of a nonpolar fibril with an antiparallel arrangement of its protofilaments has yet been determined. Mass per unit length (MPL) measurements using scanning transmission EM (STEM)65,211,227 have been used to determine the number of protofilaments in β2m211, IAPP228 and α-synuclein216 fibrils and to differentiate polymorphs of Aβ40 with two-fold and three-fold symmetry structures58,212,213,232 (FIG. 5c). MPL can also be determined by tilt-beam transmission EM (TB-TEM) imaging233 or by mass spectrometry of whole fibrils234, although STEM remains the most commonly used approach.
The arrangements of the protofilaments in recently determined amyloid fibril structures are shown in FIG. 5. These include the 2A and 3Q fibrils of Aβ40 described above (FIG. 5c), in which different fibril morphologies correspond to different organization of similar (but not identical) β-loop–β subunit motifs212. In other cases, including the paired helical and straight filaments of tau28, different fibril morphologies are formed by identical subunits held together by different interactions (FIG. 5d). Although the cryo-EM structures of the tau polymorphs provide the first direct evidence for different fibril morphologies arising from alternative packing of identical subunits, this phenomenon has been suggested for α-synuclein64,230, Aβ40 (REFS71,72), TTR67 and β2m211 fibrils. Protofilaments in paired helical filaments of tau28 are stabilized by backbone hydrogen-bonding within its 332PGGGQ336 sequence, which forms an antiparallel poly-glycine (polyG) II β-spiral235, as well as by hydrogen bonding between the side chain of Gln336 and the backbone of Pro332 in the opposing protofilament. In the recent cryo-EM structure of Aβ42 (REF.27), the protofilaments were shown to be held together by a hydrophobic steric zipper involving the Val39 and Ile41 side chains from each protofilament and stabilized further by a salt bridge between the amino-terminal Asp and Lys28 side chains of opposing protofilaments (FIG. 4c). By contrast, in the asymmetrical straight filament (SF) tau polymorph28, neither backbone nor side chain interactions between protofilaments appear to play a large role in stabilizing its structure. Instead, six side chains (Lys317, Lys321 and Thr319 from each protofilament) coordinate an unknown density that the authors predict is the amino-terminal 7EFE9 of each tau subunit. This density could, however, involve another poly-anionic molecule such as a glycosaminoglycan or nucleic acid in this crude brain extract28. Analyses of amyloid fibrils of α-synuclein with stripes of high electron density running down the fibril axis230,236 suggest a role for metal ions in stabilizing fibrils, consistent with many studies that have proposed metal ion binding in the initiation of amyloid formation237.
Data from X-ray diffraction and micro-ED52 of small, amyloidogenic peptides have been used to predict the formation of steric zippers53–55,76. As the crystal packing interactions in these steric zippers are necessarily formed between identical peptides, they can predict only regions that self-associate in symmetrical, paired protofilaments. A structure of the NAC core domain of α-synuclein by micro-ED55 was interpreted as showing possible candidates for inter-protofilament steric zippers in fibrils formed from the intact 140-residue protein. Despite identification of several polymorphs of α-synuclein fibrils composed of twisted protofilaments227,230, as yet no high-resolution structure of α-synuclein fibrils containing multiple protofilaments is available to validate this hypothesis.
The recent structures of amyloid fibrils show that the same primary sequence can assemble into different structures, even under the same growth conditions76,187,238,239. This finding is in stark contrast with the folding of the vast majority of globular proteins, in which a given sequence forms the same fold every time it emerges from the ribosome or is folded in vitro240. Moreover, amyloid formation is slow, despite being thermodynamically favoured; therefore, it can take an extraordinarily long time, possibly years in vitro241, for fibril growth to reach equilibrium. Thus, the morphology of fibrils can change over time241 and in response to changing environmental conditions77. In energetic terms, the energy landscape of fibril formation is far more rugged and complex than that for globular proteins, potentially involving multiple intermediates, parallel assembly pathways and resulting in diverse amyloid end products with different cross-β structures242.
Amyloid polymorphism in disease
How the observed structural polymorphism of amyloid fibrils affects disease onset, progression and presentation is not well understood. What is clear is that variations in the precursor protein’s gene sequence can be directly tied to variation in both the age of onset and disease duration in numerous amyloid diseases (FIG. 6).
Fig. 6. How changes in primary sequence affect amyloid disease.
Top left panel: various diseases are caused by poly-glutamine (polyQ) expansion disorders. Depending on the specific disease (shown in the figure), polyQ repeat lengths exceeding a critical threshold can cause disease, whereas fewer repeats are innocuous. Data were taken from reFs93,105. Lower left panel: the age of onset of patients with Parkinson disease (PD) is influenced by the copy number of the α-synuclein gene (duplication (2SNCA), triplication (3SNCA) or quadruplication (4SNCA)), with increased expression correlating with earlier onset. Age of onset and disease duration are also influenced by single point mutations, which may result in different aggregation pathways and/or kinetics or different fibril architectures resulting in different disease phenotypes. Data were taken from reFs299,300. Top right panel: the pathology of Alzheimer disease (AD) can be influenced by fibril morphology. In particular, typical-AD (t-AD) and a rapidly progressive form of AD (r-AD) show similar fibril architecture monitored by solid-state NMR spectroscopy (ssNMR) but have varied ages of onset and disease duration. However, in posterior cortical atrophy AD (PCA-AD), fibrils with a different structure are formed. The age of onset and disease duration for PCA-AD are similar to t-AD and r-AD, but the disease primarily affects the cerebellum rather than the temporal lobe. Centre panel: a diagram of the brain highlighting the regions primarily affected by each of the diseases shown. CACNA1A, voltage-dependent P/Q-type calcium channel subunit α1A; HD, Huntington disease; TBP, TATA-box-binding protein; WT, wild type. The top right panel was adapted from REF.60.
Polymorphism.
The same protein can assemble into amyloid fibrils that have different arrangements of subunits in the fibril, numbers of protofilaments, widths and/or crossover distances.
Although the development of disease can take place over different timescales in individuals affected by polyQ expansion diseases, AD and PD (FIG. 6), in prion diseases death is observed over an extraordinarily narrow window of a few days243. Humans also show extended incubation periods in CJD244 but have short disease duration (typically 6 months), which is thus far unique to these diseases. In contrast to prion disorders, mouse models of AD result in death over a much wider time span245, as is observed in humans60, although the ability for mouse models to accurately recapitulate disease processes in humans is debated245. In Huntington disease, individuals containing huntingtin with fewer than 35 Gln residues appear healthy over their lifetime. However, individuals with longer polyQ repeat lengths show an age of onset and severity of disease that correlates with the number of Gln residues above this critical threshold105 (FIG. 6, top left). In Machado−Joseph disease, another polyQ disorder in which ataxin 3 contains a glutamine expansion, the non-pathological polyQ repeat length is 12–40 residues, with pathology not presenting until 62 repeats105. The wide range in the number of repeats, coupled with the fact that increasing the number of repeats increases the severity of disease symptoms, suggests that polymorphic effects due to primary sequence expansion are linked with disease105.
PD can be caused by mutations within the SNCA gene, which encodes α-synuclein, with variants such as A30P, E46K, H50Q, G51D and A53T leading to early-onset disease and shorter disease duration102,246–249 (FIG. 6, lower left). For example, the G51D variant is as aggressive as quadruplication of the SNCA gene in terms of age of disease onset but has also been shown to induce other symptoms, including epilepsy250. Although gene multiplication decreases the age of onset of PD, as expected from an increased concentration of the precursor protein, the additional symptoms in patients with SNCA gene multiplications cannot be accounted for simply by amyloid load. This finding adds weight to the view that specific amyloid polymorphs may explain the observed phenotypes of PD250. Specific point mutations can give rise to diverse effects on cognitive impairment, psychiatric disturbances, hallucinations, autonomic dysfunction and the previously mentioned symptoms250.
Primary sequence variations including sequence expansions (such as polyQ, polyA and poly-GlyAla expansions93) or single point mutations (such as those described for α-synuclein above) may result in a different subunit fold and subunit packing in the fibril. Consistent with this notion, ssNMR studies of fibrils formed from the ‘Osaka’ (ΔE22) and ‘Iowa’ (D23N) mutants of Aβ40 revealed structures that are different from those formed by wild-type Aβ40/42 (REFS61,251). Links between polymorphism and pathology in fibrils with identical primary sequences are less clear but still evident. The age of disease onset in patients with Huntington disease has been correlated with the number of CAG repeats252, and different polymorphs of Aβ40 (REF.253) and α-synuclein227 have both different cytotoxicity in vitro and different rates of plaque deposition in vivo on the basis of the source of the fibrillar seed material254,255. Different cytotoxicity was also observed in fibrils of full-length IAPP seeded from toxic and non-toxic short peptide fibrils54. Isolation and characterization of fibrils from patients also suggest that fibril morphology varies from patient to patient. As discussed above, seeding fibril formation of Aβ40/42 using plaques extracted from patients who displayed different disease phenotype and progression resulted in different ssNMR spectra60,213, suggesting that a different fibril polymorph was predominant in each patient. This is not to suggest that all fibrils formed in vivo will be monomorphic. Indeed, multiple polymorphs of tau were present in a single patient-derived sample28, and various Aβ polymorphs have been identified within the same patients using X-ray diffraction256. Further evidence comes from analysis of fibrils seeded from ex vivo material from patients who presented either the posterior cortical atrophy form of AD (PCA-AD), rapid-AD (r-AD) or typical AD (t-AD)60. A single Aβ40 polymorph was most abundant in PCA-AD and t-AD, whereas the r-AD samples contained a higher proportion of additional structures. Again, the ex vivo samples were structurally distinct from synthetic strains created in vitro, highlighting remarkable strain polymorphism even within an individual patient60.
The diversity of amyloid polymorphs could explain the different disease phenotypes observed in patients in which the same APP aggregates. Alternatively, given that fibril formation can take years to reach equilibrium241, fibrils may slowly change in structure. Indeed, fibril plaques of Aβ in transgenic mouse models have been shown to undergo structural rearrangement257, and different α-synuclein fibril polymorphs evolve over time in vitro241. Fibril structures in patients may thus vary during disease progression. Such findings may explain why amyloid plaque load does not correlate with disease symptoms258,259; thus, confirmation of fibril structure post-mortem is needed to link fibril structure with disease type.
Potential for therapeutic interventions
Studies into which amyloid structure is associated with a particular presentation of disease are in their infancy, but development of better diagnostics may help to identify patients in most need of urgent therapeutic intervention. Disease is currently diagnosed using cerebral spinal fluid diagnostics, positron emission tomography (PET) imaging and detection of oligomers using immunoassays, including conformation-dependent approaches260. However, personalized amyloid-blocking medicines tailored to the specific amyloid polymorphs present in individual patients remain an aspiration. At present, with one notable exception261, no therapeutics have been shown to reduce amyloid formation in humans.
Another important consideration in targeting these devastating diseases is when in the amyloid aggregation cascade or disease process to intervene. This consideration is exacerbated by the dynamic nature of the precursors of amyloid formation that are unfolded, non-native or partially folded, precluding structure-based drug design. Decreasing the concentration of monomer would reduce the total amount of protein available to form amyloid, but such strategies could be deleterious if the monomer has a vital functional role. Interestingly, however, the function of several amyloid proteins remains unresolved262,263. Perturbing oligomer formation (Fig. 2) may offer a route to treatment, but the dynamic nature of these species makes them difficult to target264, and no high-resolution amyloid oligomer structures are currently available to guide such efforts. Decreasing the population of any one oligomer may drive the equilibrium towards a more toxic assembly. Small molecules could be sought to prevent oligomer formation or, conversely, to promote fibrillation, which, in turn, would decrease the lifetime and hence toxic potential of the oligomeric species. Designing interventions that promote the assembly of less toxic fibril polymorphs might also reduce disease severity.
Another approach would be to promote the sequestration of amyloidogenic monomers into non-amyloid amorphous aggregates265, although this could itself have deleterious effects. Finally, preventing the spread of toxic species may offer a new form of intervention. Oligomers have been shown to interact with many cellular surface receptors in both presynaptic and post-synaptic membranes, such as ephrin-type B receptor 2 (EphB2), PrPC (REF.266), renal tumour antigen (RAGE; also known as MOK) and scavenger receptor class A (SCARA) and class B (SCARB), mediating both their toxic effects and internalization. Other cell surface receptors such as lymphocyte activation gene 3 protein (LAG3) (REF.267) have been shown to be important for endocytic import of fibrillar species. Containing toxic species in individually affected cells and/or preventing uptake by antibody blocking of receptors267 may prevent the prion-like spread and hence may halt the progression of disease. Knowing where to intervene relies on identifying the toxic species in amyloid aggregation, which may depend on the protein sequence and/or the cell types involved. Culprits could include any or all of on-pathway and off-pathway oligomers as well as the fibrils themselves268.
Although effective therapy still currently remains out of reach, both drug-dependent and drug-resistant strains of prion diseases have been reported269–271, suggesting that an opportunity for therapeutic intervention does exist. Moreover, small molecules, peptides and peptidomimetics able to target different, but closely related, steric zippers have been successfully designed272–274, giving hope that we will soon be able to identify which amyloid fibrils form in different individuals and hence to relate amyloid structure and assembly to disease type and aetiology. What is needed at the current time is the ability of reagents able to stabilize oligomers with specific conformations or to prevent their formation so that the link between disease and the population of a specific conformeric species can be defined for different amyloid diseases. In a similar vein, reagents able to guide amyloid formation to a specific, well-defined structural state will also enable the link between fibril formation and disease to be made. What is clear is that the answer may be different for different proteins and protein sequences; hence, an understanding of amyloid structure and of oligomer structure and conformation is going to be required before the key question of what causes amyloid disease can be answered at last.
Conclusions and future perspectives
Amyloid fibrils have highly organized hierarchical structures that are built from protofilaments in which individual monomeric subunits form (most often) parallel in-register β-strands. These protofilaments then pack against each other, forming fibrils that are stabilized by dry steric interfaces. Further inter-fibril interaction leads to the formation of the plaques and inclusions characteristic of amyloid disease. Recent advances in cryo-EM and ssNMR have given insights into the structure of amyloid fibrils in unprecedented detail, revealing a variety of structures that conform to the cross-β amyloid fold. Coupling these techniques with orthogonal data (for example, with other biophysical measurements) and data on disease type, presentation and progression will help to elucidate which amyloid formation pathways, amyloid intermediates and/or fibril structures are responsible for disease. As the capabilities of cryo-ET improve, we will be able to study amyloid fibrils and plaques and intracellular inclusions in situ with increased resolution. This increased resolution will allow us to discern how polymorphism relates to disease phenotype and how fibril structure affects and is affected by the cellular environment. The time has never been better to finally understand amyloid structure, protein aggregation mechanisms and how they relate to disease. In turn, these breakthroughs bring new hope and renewed vigour in our quest to develop agents that can diagnose, delay or even halt the progression of disease.
Crossover.
The distance it takes a fibril to achieve 180° of rotation. Crossover appears as the distance between the two narrowest points on a 2D eM or atomic force microscopy image of a twisted fibril.
Acknowledgements
The authors thank members of their laboratories and their colleagues for many helpful discussions while preparing this Review. M.G.I., M.P.J., E.W.H., N.A.R and S.E.R. acknowledge funding from the European Research Council (ERC) under the European Union’s Seventh Framework Programme (FP7/2007-2013) ERC grant agreement no. 322408 and from the Wellcome Trust (092896MA and 204963).
Footnotes
Author contributions
All authors wrote the manuscript.
Competing interests
The authors declare no competing interests.
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
- 1.Sipe JD, Cohen AS. Review: History of the amyloid fibril. J Struct Biol. 2000;130:88–98. doi: 10.1006/jsbi.2000.4221. [DOI] [PubMed] [Google Scholar]
- 2.Goedert M, Wischik CM, Crowther RA, Walker JE, Klug A. Cloning and sequencing of the cDNA encoding a core protein of the paired helical filament of Alzheimer disease: identification as the microtubule-associated protein tau. Proc Natl Acad Sci USA. 1988;85:4051–4055. doi: 10.1073/pnas.85.11.4051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Murphy MP, LeVine IIIH. Alzheimer’s disease and the β-amyloid peptide. J Alzheimers Dis. 2010;19:311–323. doi: 10.3233/JAD-2010-1221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Stefanis L. α-Synuclein in Parkinson’s disease. Cold Spring Harb Perspect Med. 2012;2:a009399. doi: 10.1101/cshperspect.a009399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Protein Data Bank. Yearly growth of total structures. rcbs.org. 2018. http://www.rcsb.org/pdb/statistics/contentGrowthChart.do?content=total .
- 6.Sipe JD, et al. Amyloid fibril proteins and amyloidosis: chemical identification and clinical classification international society of amyloidosis 2016 nomenclature guidelines. Amyloid. 2016;23:209–213. doi: 10.1080/13506129.2016.1257986. [DOI] [PubMed] [Google Scholar]
- 7.Sloane PD, et al. The public health impact of Alzheimer’s disease, 2000–2050: potential implication of treatment advances. Annu Rev Publ Health. 2002;23:213–231. doi: 10.1146/annurev.publhealth.23.100901.140525. [DOI] [PubMed] [Google Scholar]
- 8.Geula C, et al. Aging renders the brain vulnerable to amyloid beta-protein neurotoxicity. Nat Med. 1998;7:827–831. doi: 10.1038/nm0798-827. [DOI] [PubMed] [Google Scholar]
- 9.Woerner AC, et al. Cytoplasmic protein aggregates interfere with nucleocytoplasmic transport of protein and RNA. Science. 2016;350:173–176. doi: 10.1126/science.aad2033. [DOI] [PubMed] [Google Scholar]
- 10.Guo Q, et al. In situ structure of neuronal C9orf72 poly-GA aggregates reveals proteasome recruitment. Cell. 2018;172:696–705. doi: 10.1016/j.cell.2017.12.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Drummond E, et al. Proteomic differences in amyloid plaques in rapidly progressive and sporadic Alzheimer’s disease. Acta Neuropathol. 2017;133:933–954. doi: 10.1007/s00401-017-1691-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Stewart KL, et al. Atomic details of the interactions of glycosaminoglycans with amyloid-β fibrils. J Am Chem Soc. 2016;138:8328–8331. doi: 10.1021/jacs.6b02816. [DOI] [PubMed] [Google Scholar]
- 13.Kollmer M, et al. Electron tomography reveals the fibril structure and lipid interactions in amyloid deposits. Proc Natl Acad Sci USA. 2016;113:5604–5609. doi: 10.1073/pnas.1523496113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Knowles TPJ, et al. Role of intermolecular forces in defining material properties of protein nanofibrils. Science. 2007;318:1900–1903. doi: 10.1126/science.1150057. [DOI] [PubMed] [Google Scholar]
- 15.Smith JF, Knowles TPJ, Dobson CM, MacPhee CE, Welland ME. Characterization of the nanoscale properties of individual amyloid fibrils. Proc Natl Acad Sci USA. 2006;103:15806–15811. doi: 10.1073/pnas.0604035103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Greenwald J, Friedmann MP, Riek R. Amyloid aggregates arise from amino acid condensations under prebiotic conditions. Angew Chem Int Ed Engl. 2016;55:11609–11613. doi: 10.1002/anie.201605321. [DOI] [PubMed] [Google Scholar]
- 17.Romero D, Aguilar C, Losick R, Kolter R. Amyloid fibers provide structural integrity to Bacillus subtilis biofilms. Proc Natl Acad Sci USA. 2010;107:2230–2234. doi: 10.1073/pnas.0910560107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Taglialegna A, et al. Staphylococcal Bap proteins build amyloid scaffold biofilm matrices in response to environmental signals. PLOS Pathog. 2016;12:e1005711. doi: 10.1371/journal.ppat.1005711. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Lipke PN, Klotz SA, Dufrene YF, Jackson DN, Garcia-Sherman MC. Amyloid-like β-aggregates as force-sensitive switches in fungal biofilms and infections. Microbiol Mol Biol Rev. 2018;82:e00035–17. doi: 10.1128/MMBR.00035-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Garvey M, Ecroyd H, Ray NJ, Gerrard JA, Carver JA. Functional amyloid protection in the eye lens: retention of alpha-crystallin molecular chaperone activity after modification into amyloid fibrils. Biomolecules. 2017;7:E67. doi: 10.3390/biom7030067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Biesecker SG, Nicastro LK, Wilson RP, Tukel C. The functional amyloid curli protects Escherichia coli against complement-mediated bactericidal activity. Biomolecules. 2018;8:E5. doi: 10.3390/biom8010005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Bajakian TH, et al. Metal binding properties of the N-terminus of the functional amyloid Orb2. Biomolecules. 2017;7:E57. doi: 10.3390/biom7030057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Audas TE, et al. Adaptation to stressors by systemic protein amyloidogenesis. Dev Cell. 2016;39:155–168. doi: 10.1016/j.devcel.2016.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Guyonnet B, Egge N, Cornwall GA. Functional amyloids in the mouse sperm acrosome. Mol Cell Biol. 2014;34:2624–2634. doi: 10.1128/MCB.00073-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Fowler DM, et al. Functional amyloid formation within mammalian tissue. PLOS Biol. 2006;4:e6. doi: 10.1371/journal.pbio.0040006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Roan NR, et al. Peptides released by physiological cleavage of semen coagulum proteins form amyloids that enhance HIV infection. Cell Host Microbe. 2011;10:541–550. doi: 10.1016/j.chom.2011.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Gremer L, et al. Fibril structure of amyloid-β(1–42) by cryo-electron microscopy. Science. 2017;358:116–119. doi: 10.1126/science.aao2825. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Fitzpatrick AWP, et al. Cryo-EM structures of tau filaments from Alzheimer’s disease. Nature. 2017;547:185–190. doi: 10.1038/nature23002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.US National Library of Medicine. ClinicalTrials.gov. 2018. https://www.clinicaltrials.gov/ct2/show/NCT02760602 .
- 30.US National Library of Medicine. ClinicalTrials.gov. 2018. https://www.clinicaltrials.gov/ct2/show/NCT00606476 .
- 31.US National Library of Medicine. ClinicalTrials.gov. 2018. https://www.clinicaltrials.gov/ct2/show/NCT01739348 .
- 32.US National Library of Medicine. ClinicalTrialsgov. 2018. https://www.clinicaltrials.gov/ct2/show/NCT00531804 .
- 33.Virchow R. Zur cellulose. Archiv Pathol Anat. 1854;6:416–426. [Google Scholar]
- 34.Friedreich N, Kekulé A. Zur amyloidfrage. Virchows Arch. 1859;16:50–65. [Google Scholar]
- 35.Puchtler H, Sweat F, Levine M. On the binding of Congo red by amyloid. J Histochem Cytochem. 1962;10:355–364. [Google Scholar]
- 36.Benditt EP, Eriksen N, Hermodson MA, Ericsson LH. The major proteins of human and monkey amyloid substance: common properties including unusual N-terminal amino acid sequences. FEBS Lett. 1971;19:169–173. doi: 10.1016/0014-5793(71)80506-9. [DOI] [PubMed] [Google Scholar]
- 37.Glenner GG, Eanes ED, Bladen HA, Terry W, Page DL. Creation of “amyloid” fibrils from Bence Jones proteins in vitro. Science. 1971;174:712–714. doi: 10.1126/science.174.4010.712. [DOI] [PubMed] [Google Scholar]
- 38.Costa PP, Figueria AS, Bravo FR. Amyloid fibril protein related to prealbumin in familial amyloidotic polyneuropathy. Proc Natl Acad Sci USA. 1978;75:4499–4503. doi: 10.1073/pnas.75.9.4499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Astbury WT, Street A. X-ray studies of the structure of hair, wool and related fibres. I. General. Phil Trans R Soc. 1932;230:75–101. [Google Scholar]
- 40.Hall KT. The Man in the Monkeynut Coat. Oxford Univ. Press; Oxford: 2014. [Google Scholar]
- 41.Geddes AJ, Parker KD, Atkins EDT, Beighton E. “Cross-β” conformation in proteins. J Mol Biol. 1968;32:343–358. doi: 10.1016/0022-2836(68)90014-4. [DOI] [PubMed] [Google Scholar]
- 42.Jackson M, Hewitt E. Why are functional amyloids non-toxic in humans? Biomolecules. 2017;7:E71–E73. doi: 10.3390/biom7040071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Hewetson A, et al. Functional amyloids in reproduction. Biomolecules. 2017;7:E46. doi: 10.3390/biom7030046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Ramsook CB, et al. Yeast cell adhesion molecules have functional amyloid-forming sequences. Eukaryot Cell. 2010;9:393–404. doi: 10.1128/EC.00068-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Romero D, Kolter R. Functional amyloids in bacteria. Int Microbiol. 2014;17:65–73. doi: 10.2436/20.1501.01.208. [DOI] [PubMed] [Google Scholar]
- 46.Pham CL, Kwan AH, Sunde M. Functional amyloid: widespread in Nature, diverse in purpose. Essays Biochem. 2014;56:207–219. doi: 10.1042/bse0560207. [DOI] [PubMed] [Google Scholar]
- 47.Fowler DM, Koulov AV, Balch WE, Kelly JW. Functional amyloid — from bacteria to humans. Trends Biochem Sci. 2007;32:217–224. doi: 10.1016/j.tibs.2007.03.003. [DOI] [PubMed] [Google Scholar]
- 48.Morris KL, Serpell LC. X-Ray Fibre Diffraction Studies of Amyloid Fibrils. Humana Press; 2012. [DOI] [PubMed] [Google Scholar]
- 49.Eanes ED, Glenner GG. X-ray diffraction studies on amyloid filaments. J Histochem Cytochem. 1968;10:673–677. doi: 10.1177/16.11.673. [DOI] [PubMed] [Google Scholar]
- 50.Kirschner DA, Abraham C, Selkoe DJ. X-ray diffraction from intraneuronal paired helical filaments and extraneuronal amyloid fibers in Alzheimer disease indicates cross-beta conformation. Proc Natl Acad Sci USA. 1986;83:503–507. doi: 10.1073/pnas.83.2.503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Sawaya MR, et al. Atomic structures of amyloid cross-β spines reveal varied steric zippers. Nature. 2007;447:453–457. doi: 10.1038/nature05695. [DOI] [PubMed] [Google Scholar]
- 52.Shi D, Nannenga BL, Iadanza MG, Gonen T. Three-dimensional electron crystallography of protein microcrystals. eLife. 2013;2:e01345. doi: 10.7554/eLife.01345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.de la Cruz MJ, et al. Atomic-resolution structures from fragmented protein crystals with the cryoEM method MicroED. Nat Med. 2017;14:399–402. doi: 10.1038/nmeth.4178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Krotee P, et al. Atomic structures of fibrillar segments of hIAPP suggest tightly mated beta-sheets are important for cytotoxicity. eLife. 2017;6:e19273. doi: 10.7554/eLife.19273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Rodriguez JA, et al. Structure of the toxic core of alpha-synuclein from invisible crystals. Nature. 2015;525:486–490. doi: 10.1038/nature15368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Sawaya MR, et al. Ab initio structure determination from prion nanocrystals at atomic resolution by MicroED. Proc Natl Acad Sci USA. 2016;113:11232–11236. doi: 10.1073/pnas.1606287113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Colvin MT, et al. High resolution structural characterization of Aβ42 amyloid fibrils by magic angle spinning NMR. J Am Chem Soc. 2015;137:7509–7518. doi: 10.1021/jacs.5b03997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Antzutkin ON, Leapman RD, Balbach JJ, Tycko R. Supramolecular structural constraints on Alzheimer’s β-amyloid fibrils from electron microscopy and solid-state nuclear magnetic resonance. Biochemistry. 2002;41:15436–15450. doi: 10.1021/bi0204185. [DOI] [PubMed] [Google Scholar]
- 59.Chan JCC, Oyler NA, Yau W, Tycko R. Parallel β-sheets and polar zippers in amyloid fibrils formed by residues 10–39 of the yeast prion protein Ure2p. Biochemistry. 2005;44:10669–10680. doi: 10.1021/bi050724t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Qiang W, Yau WM, Lu JX, Collinge J, Tycko R. Structural variation in amyloid-beta fibrils from Alzheimer’s disease clinical subtypes. Nature. 2017;541:217–221. doi: 10.1038/nature20814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Qiang W, Yau WM, Luo Y, Mattson MP, Tycko R. Antiparallel beta-sheet architecture in Iowa-mutant beta-amyloid fibrils. Proc Natl Acad Sci USA. 2012;109:4443–4448. doi: 10.1073/pnas.1111305109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Lührs T, et al. 3D structure of Alzheimer’s amyloid-B(1–42) fibrils. Proc Natl Acad Sci USA. 2005;102:17342–17347. doi: 10.1073/pnas.0506723102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Nelson R, et al. Structure of the cross-β spine of amyloid-like fibrils. Nature. 2005;435:773–778. doi: 10.1038/nature03680. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Vilar M, et al. The fold of α-synuclein fibrils. Proc Natl Acad Sci USA. 2008;105:8637–8642. doi: 10.1073/pnas.0712179105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Walti MA, et al. Atomic-resolution structure of a disease-relevant Abeta(1–42) amyloid fibril. Proc Natl Acad Sci USA. 2016;113:E4976–E4984. doi: 10.1073/pnas.1600749113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Wasmer C, et al. Amyloid fibrils of the HET-s(218–289) prion from a β solenoid with a triangular hydrophobic core. Science. 2008;319:1523–1526. doi: 10.1126/science.1151839. [DOI] [PubMed] [Google Scholar]
- 67.Fitzpatrick AWP, et al. Atomic structure and hierarchical assembly of a cross-β amyloid fibril. Proc Natl Acad Sci USA. 2013;110:5468–5473. doi: 10.1073/pnas.1219476110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Jiménez J, et al. Cryo-electron microscopy structure of an SH3 amyloid fibril and model of the molecular packing. EMBO J. 1999;18:815–821. doi: 10.1093/emboj/18.4.815. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Saibil HR, et al. Heritable yeast prions have a highly organized three-dimensional architecture with interfiber structures. Proc Natl Acad Sci USA. 2012;109:14906–14911. doi: 10.1073/pnas.1211976109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Fändrich M, Meinhardt J, Grigorieff N. Structural polymorphism of Alzheimer Aβ and other amyloid fibrils. Prion. 2009;3:89–93. doi: 10.4161/pri.3.2.8859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Sachse C, Fändrich M, Grigorieff N. Paired β-sheet structure of an Aβ(1–40) amyloid fibril revealed by electron microscopy. Proc Natl Acad Sci USA. 2008;105:7462–7466. doi: 10.1073/pnas.0712290105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Sachse C, et al. Quaternary structure of a mature amyloid fibril from Alzheimer’s Abeta(1–40) peptide. J Mol Biol. 2006;362:347–354. doi: 10.1016/j.jmb.2006.07.011. [DOI] [PubMed] [Google Scholar]
- 73.Kirschner DA, et al. In vitro amyloid fibril formation by synthetic peptides corresponding to the amino terminus of apoSAA isoforms from amyloid-susceptible and amyloid-resistant mice. J Struct Biol. 1998;124:88–98. doi: 10.1006/jsbi.1998.4047. [DOI] [PubMed] [Google Scholar]
- 74.Castaño EM, et al. In vitro formation of amyloid fibrils from two synthetic peptides of different lengths homologous to Alzheimer’s disease β-protein. Biochem Biophys Res Commun. 1986;141:782–789. doi: 10.1016/s0006-291x(86)80241-8. [DOI] [PubMed] [Google Scholar]
- 75.Aggeli A, et al. Responsive gels formed by the spontanious self-assembly of peptides into polymeric β-sheet tapes. Nature. 1997;386:259–262. doi: 10.1038/386259a0. [DOI] [PubMed] [Google Scholar]
- 76.Colletier JP, et al. Molecular basis for amyloid-beta polymorphism. Proc Natl Acad Sci USA. 2011;108:16938–16943. doi: 10.1073/pnas.1112600108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Gosal WS, et al. Competing pathways determine fibril morphology in the self-assembly of beta2-microglobulin into amyloid. J Mol Biol. 2005;351:850–864. doi: 10.1016/j.jmb.2005.06.040. [DOI] [PubMed] [Google Scholar]
- 78.Booth D, et al. Instability, unfolding and aggregation of human lysozyme variants underlying amyloid fibrillogenesis. Nature. 1997;385:797–793. doi: 10.1038/385787a0. [DOI] [PubMed] [Google Scholar]
- 79.Hecht MH, Das A, Go A, Bradley LH, Wei Y. De novo proteins from designed combinatorial libraries. Protein Sci. 2004;13:1711–1723. doi: 10.1110/ps.04690804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Kühlbrandt W. The resolution revolution. Science. 2014;343:1443–1444. doi: 10.1126/science.1251652. [DOI] [PubMed] [Google Scholar]
- 81.Pinotsi D, et al. Direct observation of heterogeneous amyloid fibril growth kinetics via two-color super-resolution microscopy. Nano Lett. 2014;14:339–345. doi: 10.1021/nl4041093. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Kaminski Schierle GS, et al. In situ measurements of the formation and morphology of intracellular beta-amyloid fibrils by super-resolution fluorescence imaging. J Am Chem Soc. 2011;133:12902–12905. doi: 10.1021/ja201651w. [DOI] [PubMed] [Google Scholar]
- 83.Han S, et al. Amyloid plaque structure and cell surface interactions of beta-amyloid fibrils revealed by electron tomography. Sci Rep. 2017;7:43577. doi: 10.1038/srep43577. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Bauerlein FJB, et al. In situ architecture and cellular interactions of polyQ inclusions. Cell. 2017;171:179–187. doi: 10.1016/j.cell.2017.08.009. [DOI] [PubMed] [Google Scholar]
- 85.Tang M, Comellas G, Rienstra CM. Advanced solid-state NMR approaches for structure determination of membrane proteins and amyloid fibrils. Acc Chem Res. 2013;46:2080–2088. doi: 10.1021/ar4000168. [DOI] [PubMed] [Google Scholar]
- 86.Alzheimer A. Über einen eigenartigen schweren Erkrankungsprozeβ der Hirnrincle. Neurol Central. 1906;25:1134 [Google Scholar]
- 87.Alzheimer A. Über eine eigenartige Erkrankung der Hirnrinde. Allg Z Psychiatr Psych-Gerichtl Med. 1907;641:46–48. [Google Scholar]
- 88.Chiti F, Dobson CM. Protein misfolding, functional amyloid, and human disease: a summary of progress over the last decade. Annu Rev Biochem. 2006;86:27–68. doi: 10.1146/annurev-biochem-061516-045115. [DOI] [PubMed] [Google Scholar]
- 89.Eisenberg D, Jucker M. The amyloid state of proteins in human diseases. Cell. 2012;148:1188–1203. doi: 10.1016/j.cell.2012.02.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Westermark P, et al. Amyloid: toward terminology clarification. Report from the Nomenclature Committee of the International Society of Amyloidosis. Amyloid. 2005;12:1–4. doi: 10.1080/13506120500032196. [DOI] [PubMed] [Google Scholar]
- 91.Vassar R, Bennett BD, Babu-Khan S, Kahn S. β-Secretase cleavage of Alzheimer’s amyloid precursor protein by the transmembrane aspartic protease BACE. Science. 1999;286:735–741. doi: 10.1126/science.286.5440.735. [DOI] [PubMed] [Google Scholar]
- 92.Prusiner SB, Bowman KA, Bendheim PE, Glenner GG. Scrapie prions aggregate to form amyloid-like birefringent rods. Cell. 1983;35:349–358. doi: 10.1016/0092-8674(83)90168-x. [DOI] [PubMed] [Google Scholar]
- 93.Warby SC, et al. CAG expansion in the Huntington disease gene is associated with a specific and targetable predisposing haplogroup. Am J Hum Genet. 2009;84:351–366. doi: 10.1016/j.ajhg.2009.02.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Chartier-Harlin MC, et al. Alpha-synuclein locus duplication as a cause of familial Parkinson’s disease. Lancet. 2004;364:1167–1169. doi: 10.1016/S0140-6736(04)17103-1. [DOI] [PubMed] [Google Scholar]
- 95.Valentine JS, Doucette PA, Zittin Potter S. Copper-zinc superoxide dismutase and amyotrophic lateral sclerosis. Annu Rev Biochem. 2005;74:563–593. doi: 10.1146/annurev.biochem.72.121801.161647. [DOI] [PubMed] [Google Scholar]
- 96.Westermark P, Andersson A, Westermark GT. Islet amyloid polypeptide, islet amyloid, and diabetes mellitus. Physiol Rev. 2011;91:795–826. doi: 10.1152/physrev.00042.2009. [DOI] [PubMed] [Google Scholar]
- 97.Sanchorawala V. Light-chain (AL) amyloidosis: diagnosis and treatment. Clin J Am Soc Nephrol. 2006;1:1331–1335. doi: 10.2215/CJN.02740806. [DOI] [PubMed] [Google Scholar]
- 98.Koch KM. Dialysis-related amyloidosis. Kidney Int. 1992;41:1416–1429. doi: 10.1038/ki.1992.207. [DOI] [PubMed] [Google Scholar]
- 99.Li X, Song D, Leng SX. Link between type 2 diabetes and Alzheimer’s disease: from epidemiology to mechanism and treatment. Clin Interv Aging. 2015;10:549–560. doi: 10.2147/CIA.S74042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Uéda K, et al. Molecular cloning of cDNA encoding an unrecognized component of amyloid in Alzheimer disease. Proc Natl Acad Sci USA. 1993;90:11282–11286. doi: 10.1073/pnas.90.23.11282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Maresova P, Klimova B, Novotny M, Kuca K. Alzheimer’s and Parkinson’s diseases: Expected economic Impact on Europe-A call for a uniform European strategy. J Alzheimers Dis. 2016;54:1123–1133. doi: 10.3233/JAD-160484. [DOI] [PubMed] [Google Scholar]
- 102.Li J, Uversky VN, Fink AL. Effect of familial Parkinson’s disease point mutations A30P and A53T on the structural properties, aggregation, and fibrillation of human alpha-synuclein. Biochemistry. 2001;40:11604–11613. doi: 10.1021/bi010616g. [DOI] [PubMed] [Google Scholar]
- 103.Krone MG, et al. Effects of familial Alzheimer’s disease mutations on the folding nucleation of the amyloid beta-protein. J Mol Biol. 2008;381:221–228. doi: 10.1016/j.jmb.2008.05.069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Mangione PP, et al. Structure, folding dynamics, and amyloidogenesis of D76N beta2-microglobulin: roles of shear flow, hydrophobic surfaces, and alpha-crystallin. J Biol Chem. 2013;288:30917–30930. doi: 10.1074/jbc.M113.498857. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Fan HC, et al. Polyglutamine (PolyQ) diseases: genetics to treatments. Cell Transplant. 2014;23:441–458. doi: 10.3727/096368914X678454. [DOI] [PubMed] [Google Scholar]
- 106.Scheuermann T, et al. Trinucleotide expansions leading to an extended poly-L-alanine segment in the poly (A) binding protein PABPN1 cause fibril formation. Protein Sci. 2003;12:2685–2692. doi: 10.1110/ps.03214703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Brais B, et al. Short GCG expansions in the PABP2 gene cause oculopharyngeal muscular dystrophy. Nat Genet. 1998;18:164–167. doi: 10.1038/ng0298-164. [DOI] [PubMed] [Google Scholar]
- 108.Renton AE, et al. A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron. 2011;72:257–268. doi: 10.1016/j.neuron.2011.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.DeJesus-Hernandez M, et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron. 2011;72:245–256. doi: 10.1016/j.neuron.2011.09.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Budworth H, McMurray CT. A brief history of triplet repeat diseases. Methods Mol Biol. 2013;1010:3–17. doi: 10.1007/978-1-62703-411-1_1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Wiltfang J, et al. Amyloid beta peptide ratio 42/40 but not Aβ 42 correlates with phospho-Tau in patients with low- and high-CSF Aβ 40 load. J Neurochem. 2007;101:1053–1059. doi: 10.1111/j.1471-4159.2006.04404.x. [DOI] [PubMed] [Google Scholar]
- 112.Ramella NA, et al. Human apolipoprotein A-I-derived amyloid: its association with atherosclerosis. PLOS One. 2011;6:e22532. doi: 10.1371/journal.pone.0022532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Chiti F, et al. A partially structured species of beta 2-microglobulin is significantly populated under physiological conditions and involved in fibrillogenesis. J Biol Chem. 2001;276:46714–46721. doi: 10.1074/jbc.M107040200. [DOI] [PubMed] [Google Scholar]
- 114.Eichner T, Radford SE. A generic mechanism of β2-microglobulin amyloid assembly at neutral pH involving a specific proline switch. J Mol Biol. 2009;386:1312–1326. doi: 10.1016/j.jmb.2009.01.013. [DOI] [PubMed] [Google Scholar]
- 115.Byers B, et al. SNCA triplication Parkinson’s patient’s iPSC-derived DA neurons accumulate α-synuclein and are susceptible to oxidative stress. PLOS One. 2011;6:e26159. doi: 10.1371/journal.pone.0026159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Lott IT, Head E. Alzheimer disease and Down syndrome: factors in pathogenesis. Neurobiol Aging. 2005;26:383–389. doi: 10.1016/j.neurobiolaging.2004.08.005. [DOI] [PubMed] [Google Scholar]
- 117.Scarpioni R, et al. Dialysis-related amyloidosis: challenges and solutions. Int J Nephrol Renovasc Dis. 2016;9:319–328. doi: 10.2147/IJNRD.S84784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Banani SF, Lee HO, Hyman AA, Rosen MK. Biomolecular condensates: organizers of cellular biochemistry. Nat Rev Mol Cell Biol. 2017;18:285–298. doi: 10.1038/nrm.2017.7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Wegmann S, et al. Tau protein liquid-liquid phase separation can initiate tau aggregation. EMBO J. 2018;37:e98049. doi: 10.15252/embj.201798049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Xiang S, et al. The LC domain of hnRNPA2 adopts similar conformations in hydrogel polymers, liquid-like droplets, and nuclei. Cell. 2015;163:829–839. doi: 10.1016/j.cell.2015.10.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Olzscha H, et al. Amyloid-like aggregates sequester numerous metastable proteins with essential cellular functions. Cell. 2011;144:67–78. doi: 10.1016/j.cell.2010.11.050. [DOI] [PubMed] [Google Scholar]
- 122.Kovacs GG, Budka H. Prion diseases: from protein to cell pathology. Am J Pathol. 2008;172:555–565. doi: 10.2353/ajpath.2008.070442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Aguzzi A, Calella AM. Prions: protein aggregation and infectious diseases. Physiol Rev. 2009;89:1105–1150. doi: 10.1152/physrev.00006.2009. [DOI] [PubMed] [Google Scholar]
- 124.Aguzzi A, Baumann F, Bremer J. The prion’s elusive reason for being. Annu Rev Neurosci. 2008;31:439–477. doi: 10.1146/annurev.neuro.31.060407.125620. [DOI] [PubMed] [Google Scholar]
- 125.Botsios S, Manuelidis L. CJD and scrapie require agent-associated nucleic acids for infection. J Cell Biochem. 2016;117:1947–1958. doi: 10.1002/jcb.25495. [DOI] [PubMed] [Google Scholar]
- 126.Wadsworth JDF, et al. Kuru prions and sporadic Creutzfeldt–Jakob disease prions have equivalent transmission properties in transgenic and wild-type mice. Proc Natl Acad Sci USA. 2008;105:3885–3890. doi: 10.1073/pnas.0800190105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Cobb NJ, Surewicz WK. Prion diseases and their biochemical mechanisms. Biochemistry. 2009;48:2574–2585. doi: 10.1021/bi900108v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Wadsworth JDF, et al. Tissue distribution of protease resistant prion protein in variant Creutzfeldt-Jakob disease using a highly sensitive immunoblotting assay. Lancet. 2001;358:171–180. doi: 10.1016/s0140-6736(01)05403-4. [DOI] [PubMed] [Google Scholar]
- 129.Masuda-Suzukake M, et al. Prion-like spreading of pathological alpha-synuclein in brain. Brain. 2013;136:1128–1138. doi: 10.1093/brain/awt037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Walker LC, Schelle J, Jucker M. The prion-like properties of amyloid-β assemblies: implications for Alzheimer’s disease. Cold Spring Harb Perspect Med. 2016;6:a024398. doi: 10.1101/cshperspect.a024398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.An L, Fitzpatrick D, Harrison PM. Emergence and evolution of yeast prion and prion-like proteins. BMC Evol Biol. 2016;16:24. doi: 10.1186/s12862-016-0594-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Alberti S, Halfmann R, King O, Kapila A, Lindquist S. A systematic survey identifies prions and illuminates sequence features of prionogenic proteins. Cell. 2009;137:146–158. doi: 10.1016/j.cell.2009.02.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Sponarova J, Nystrom SN, Westermark GT. AA-amyloidosis can be transferred by peripheral blood monocytes. PLOS One. 2008;3:e3308. doi: 10.1371/journal.pone.0003308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Solomon A, et al. Amyloidogenic potential of foie gras. Proc Nat Acad Sci USA. 2007;104:10998–11001. doi: 10.1073/pnas.0700848104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Knowles TP, Vendruscolo M, Dobson CM. The amyloid state and its association with protein misfolding diseases. Nat Rev Mol Cell Biol. 2014;15:384–396. doi: 10.1038/nrm3810. [DOI] [PubMed] [Google Scholar]
- 136.Ferrone F. Analysis of protein aggregation kinetics. Methods Enzymol. 1999;309:256–274. doi: 10.1016/s0076-6879(99)09019-9. [DOI] [PubMed] [Google Scholar]
- 137.Sicorello A, et al. Agitation and high ionic strength induce amyloidogenesis of a folded PDZ domain in native conditions. Biophys J. 2009;96:2289–2298. doi: 10.1016/j.bpj.2008.11.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Glabe CG, Kayed R. Common structure and toxic function of amyloid oligomers implies a common mechanism of pathogenesis. Neurology. 2006;66:S74–S78. doi: 10.1212/01.wnl.0000192103.24796.42. [DOI] [PubMed] [Google Scholar]
- 139.Glabe CG. Common mechanisms of amyloid oligomer pathogenesis in degenerative disease. Neurobiol Aging. 2006;27:570–575. doi: 10.1016/j.neurobiolaging.2005.04.017. [DOI] [PubMed] [Google Scholar]
- 140.Meisl G, et al. Molecular mechanisms of protein aggregation from global fitting of kinetic models. Nat Protoc. 2016;11:252–272. doi: 10.1038/nprot.2016.010. [DOI] [PubMed] [Google Scholar]
- 141.Linse S. Monomer-dependent secondary nucleation in amyloid formation. Biophys Rev. 2017;9:329–338. doi: 10.1007/s12551-017-0289-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Oosawa F, Asakura S. Thermodynamics of the Polymerization of Protein. Academic Press; 1975. [Google Scholar]
- 143.Eaton WA, Hofrichter J. Hemoglobin S gelation and sickle cell disease. Blood. 1987;70:1245–1266. [PubMed] [Google Scholar]
- 144.Xue WF, Homans SW, Radford SE. Systematic analysis of nucleation-dependent polymerization reveals new insights into the mechanism of amyloid self-assembly. Proc Natl Acad Sci USA. 2008;105:8926–8931. doi: 10.1073/pnas.0711664105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Cohen SIA, et al. Proliferation of amyloid-β42 aggregates occurs through a secondary nucleation mechanism. Proc Natl Acad Sci USA. 2013;110:9758–9763. doi: 10.1073/pnas.1218402110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.LeVine Hr. Thioflavine T interaction with synthetic Alzheimer’s disease beta-amyloid peptides: detection of amyloid aggregation in solution. Protein Sci. 1993;2:404–410. doi: 10.1002/pro.5560020312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Galvagnion C, et al. Lipid vesicles trigger α-synuclein aggregation by stimulating primary nucleation. Nat Chem Biol. 2015;11:229–234. doi: 10.1038/nchembio.1750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Buell AK, et al. Solution conditions determine the relative importance of nucleation and growth processes in α-synuclein aggregation. Proc Natl Acad Sci USA. 2014;27:7671–7676. doi: 10.1073/pnas.1315346111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Arosio P, et al. Kinetic analysis reveals the diversity of microscopic mechanisms through which molecular chaperones suppress amyloid formation. Nat Commun. 2016;7:10948. doi: 10.1038/ncomms10948. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Habchi J, et al. Systematic development of small molecules to inhibit specific microscopic steps of Abeta42 aggregation in Alzheimer’s disease. Proc Natl Acad Sci USA. 2017;114:E200–E208. doi: 10.1073/pnas.1615613114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Jackson MP, Hewitt EW. Cellular proteostasis: degradation of misfolded proteins by lysosomes. Essays Biochem. 2016;60:173–180. doi: 10.1042/EBC20160005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Balchin D, Hayer-Hartl M, Hartl FU. In vivo aspects of protein folding and quality control. Science. 2016;353:aac4354. doi: 10.1126/science.aac4354. [DOI] [PubMed] [Google Scholar]
- 153.Kim YE, Hipp MS, Bracher A, Hayer-Hartl M, Hartl FU. Molecular chaperone functions in protein folding and proteostasis. Annu Rev Biochem. 2013;82:323–355. doi: 10.1146/annurev-biochem-060208-092442. [DOI] [PubMed] [Google Scholar]
- 154.Salminen A, et al. Impaired autophagy and APP processing in Alzheimer’s disease: the potential role of Beclin 1 interactome. Prog Neurobiol. 2013;106–107:33–54. doi: 10.1016/j.pneurobio.2013.06.002. [DOI] [PubMed] [Google Scholar]
- 155.Winklhofer KF, Haass C. Mitochondrial dysfunction in Parkinson’s disease. Biochim Biophys Acta. 2010;1802:29–44. doi: 10.1016/j.bbadis.2009.08.013. [DOI] [PubMed] [Google Scholar]
- 156.Uttara B, Singh AV, Zamboni P, Mahajan RT. Oxidative stress and neurodegenerative diseases: a review of upstream and downstream antioxidant therapeutic options. Curr Neuropharmacol. 2009;7:65–74. doi: 10.2174/157015909787602823. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.McLaurin J, Chakrabartty A. Membrane disruption by Alzheimer β-amyloid peptides mediated through specific binding to either phospholipids or gangliosides. Implications for neurotoxicity. J Biol Chem. 1996;25:26482–26489. doi: 10.1074/jbc.271.43.26482. [DOI] [PubMed] [Google Scholar]
- 158.Goodchild SC, et al. β2-Microglobulin amyloid fibril-induced membrane disruption is enhanced by endosomal lipids and acidic pH. PLOS One. 2014;9:e104492. doi: 10.1371/journal.pone.0104492. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Nelson PT, et al. Correlation of Alzheimer disease neuropathologic changes with cognitive status: a review of the literature. J Neuropathol Exp Neurol. 2012;71:362–381. doi: 10.1097/NEN.0b013e31825018f7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Reixach N, Deechongkit S, Jiang X, Kelly JW, Buxbaum JN. Tissue damage in the amyloidoses: transthyretin monomers and non-native oligomers are the major cytotoxic species in tissue culture. Proc Natl Acad Sci USA. 2004;101:2817–2822. doi: 10.1073/pnas.0400062101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Baglioni S, et al. Prefibrillar amyloid aggregates could be generic toxins in higher organisms. J Neurosci. 2006;26:8160–8167. doi: 10.1523/JNEUROSCI.4809-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Bucciantini M, et al. Inherent toxicity of aggregates implies a common mechanism for protein misfolding diseases. Nature. 2002;416:507–511. doi: 10.1038/416507a. [DOI] [PubMed] [Google Scholar]
- 163.Simoneau S, et al. In vitro and in vivo neurotoxicity of prion protein oligomers. PLOS Pathog. 2007;3:e125. doi: 10.1371/journal.ppat.0030125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Winner B, et al. In vivo demonstration that α-synuclein oligomers are toxic. Proc Natl Acad Sci USA. 2011;108:4194–4199. doi: 10.1073/pnas.1100976108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Serra-Batiste M, et al. Aβ42 assembles into specific beta-barrel pore-forming oligomers in membrane-mimicking environments. Proc Natl Acad Sci USA. 2016;113:10866–10871. doi: 10.1073/pnas.1605104113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Evangelisti E, et al. Binding affinity of amyloid oligomers to cellular membranes is a generic indicator of cellular dysfunction in protein misfolding diseases. Sci Rep. 2016;6:32721. doi: 10.1038/srep32721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Pfefferkorn CM, Jiang Z, Lee JC. Biophysics of α-synuclein membrane interactions. Biochim Biophys Acta. 2012;1818:162–171. doi: 10.1016/j.bbamem.2011.07.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Lashuel HA, et al. α-Synuclein, especially the Parkinson’s disease-associated mutants, forms pore-like annular and tubular protofibrils. J Mol Biol. 2002;322:1089–1102. doi: 10.1016/s0022-2836(02)00735-0. [DOI] [PubMed] [Google Scholar]
- 169.Lesné S, et al. A specific amyloid-β protein assembly in the brain impairs memory. Nature. 2006;440:352–357. doi: 10.1038/nature04533. [DOI] [PubMed] [Google Scholar]
- 170.Shankar GM, et al. Amyloid-beta protein dimers isolated directly from Alzheimer’s brains impair synaptic plasticity and memory. Nat Med. 2008;14:837–842. doi: 10.1038/nm1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.Fusco G, et al. Structural basis of membrane disruption and cellular toxicity by α-synuclein oligomers. Science. 2017;358:1440–1443. doi: 10.1126/science.aan6160. [DOI] [PubMed] [Google Scholar]
- 172.Tosatto L, et al. Single-molecule FRET studies on α-synuclein oligomerization of Parkinson’s disease genetically related mutants. Sci Rep. 2015;5:16696. doi: 10.1038/srep16696. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Chiti F, Dobson CM. Amyloid formation by globular proteins under native conditions. Nat Chem Biol. 2009;5:15–22. doi: 10.1038/nchembio.131. [DOI] [PubMed] [Google Scholar]
- 174.Tsigelny IF, et al. Role of α-synuclein penetration into the membrane in the mechanisms of oligomer pore formation. FEBS J. 2012;279:1000–1013. doi: 10.1111/j.1742-4658.2012.08489.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Young LM, Cao P, Raleigh DP, Ashcroft AE, Radford SE. Ion mobility spectrometry-mass spectrometry defines the oligomeric intermediates in amylin amyloid formation and the mode of action of inhibitors. J Am Chem Soc. 2014;136:660–670. doi: 10.1021/ja406831n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Young LM, Tu LH, Raleigh DP, Ashcroft AE, Radford SE. Understanding co-polymerization in amyloid formation by direct observation of mixed oligomers. Chem Sci. 2017;8:5030–5040. doi: 10.1039/c7sc00620a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Tipping KW, van Oosten-Hawle P, Hewitt EW, Radford SE. Amyloid fibres: inert end-stage aggregates or key players in disease? Trends Biochem Sci. 2015;40:719–727. doi: 10.1016/j.tibs.2015.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Milanesi L, et al. Direct three-dimensional visualization of membrane disruption by amyloid fibrils. Proc Natl Acad Sci USA. 2012;109:20455–20460. doi: 10.1073/pnas.1206325109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Phelan MM, Caamaño-Gutiérrez E, Gant MS, Grosman RX, Madine J. Using an NMR metabolomics approach to investigate the pathogenicity of amyloid-beta and alpha-synuclein. Metabolomics. 2017;13:151. doi: 10.1007/s11306-017-1289-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Gharibyan AL, et al. Lysozyme amyloid oligomers and fibrils induce cellular death via different apoptotic/necrotic pathways. J Mol Biol. 2007;365:1337–1349. doi: 10.1016/j.jmb.2006.10.101. [DOI] [PubMed] [Google Scholar]
- 181.Grudzielanek S, et al. Cytotoxicity of insulin within its self-assembly and amyloidogenic pathways. J Mol Biol. 2007;370:372–384. doi: 10.1016/j.jmb.2007.04.053. [DOI] [PubMed] [Google Scholar]
- 182.Novitskaya V, Bocharova OV, Bronstein I, Baskakov IV. Amyloid fibrils of mammalian prion protein are highly toxic to cultured cells and primary neurons. J Biol Chem. 2006;281:13828–13836. doi: 10.1074/jbc.M511174200. [DOI] [PubMed] [Google Scholar]
- 183.Berthelot K, Ta HP, Géan J, Lecomte S, Cullin C. In vivo and in vitro analyses of toxic mutants of HET-S: FTIR antiparallel signature correlates with amyloid toxicity. J Mol Biol. 2011;412:137–152. doi: 10.1016/j.jmb.2011.07.009. [DOI] [PubMed] [Google Scholar]
- 184.Lee YJ, Savtchenko R, Ostapchenko VG, Makarava N, Baskakov IV. Molecular structure of amyloid fibrils controls the relationship between fibrillar size and toxicity. PLOS One. 2011;6:e20244. doi: 10.1371/journal.pone.0020244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Mossuto MF, et al. Disulfide bonds reduce the toxicity of the amyloid fibrils formed by an extracellular protein. Angew Chem Int Ed Engl. 2011;50:7048–7051. doi: 10.1002/anie.201100986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Makarava N, et al. Recombinant prion protein induces a new transmissible prion disease in wild-type animals. Acta Neuropathol. 2010;119:177–187. doi: 10.1007/s00401-009-0633-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Qiang W, Kelley K, Tycko R. Polymorph-specific kinetics and thermodynamics of beta-amyloid fibril growth. J Am Chem Soc. 2013;135:6860–6871. doi: 10.1021/ja311963f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Stewart KL, Hughes E, Yates EA, Middleton DA, Radford SE. Molecular origins of the compatibility between glycosaminoglycans and Aβ40 amyloid fibrils. J Mol Biol. 2017;429:2449–2462. doi: 10.1016/j.jmb.2017.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Cohen ML, et al. Rapidly progressive Alzheimer’s disease features distinct structures of amyloid-beta. Brain. 2015;138:1009–1022. doi: 10.1093/brain/awv006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Tipping KW, et al. pH-induced molecular shedding drives the formation of amyloid fibril-derived oligomers. Proc Natl Acad Sci USA. 2015;112:5691–5696. doi: 10.1073/pnas.1423174112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Serra-Vidal B, et al. Hydrogen/deuterium exchange-protected oligomers populated during Abeta fibril formation correlate with neuronal cell death. ACS Chem Biol. 2014;9:2678–2685. doi: 10.1021/cb500621x. [DOI] [PubMed] [Google Scholar]
- 192.Pilla E, Schneider K, Bertolotti A. Coping with protein quality control failure. Annu Rev Cell Dev Biol. 2017;33:439–465. doi: 10.1146/annurev-cellbio-111315-125334. [DOI] [PubMed] [Google Scholar]
- 193.Schneider K, Bertolotti A. Surviving protein quality control catastrophes—from cells to organisms. J Cell Sci. 2015;128:3861–3869. doi: 10.1242/jcs.173047. [DOI] [PubMed] [Google Scholar]
- 194.Walther DM, et al. Widespread proteome remodeling and aggregation in aging C. elegans. Cell. 2015;161:919–932. doi: 10.1016/j.cell.2015.03.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Ciryam P, Kundra R, Morimoto RI, Dobson CM, Vendruscolo M. Supersaturation is a major driving force for protein aggregation in neurodegenerative diseases. Trends Pharmacol Sci. 2015;36:72–77. doi: 10.1016/j.tips.2014.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Kundra R, Ciryam P, Morimoto RI, Dobson CM, Vendruscolo M. Protein homeostasis of a metastable subproteome associated with Alzheimer’s disease. Proc Natl Acad Sci USA. 2017;114:E5703–E5711. doi: 10.1073/pnas.1618417114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Bonar L, Cohen AS, Skinner MM. Characterization of the amyloid fibril as a cross-beta protein. Proc Soc Exp Biol Med. 1969;131:1373–1375. doi: 10.3181/00379727-131-34110. [DOI] [PubMed] [Google Scholar]
- 198.Blake C, Serpell LC. Synchrotron X-ray studies suggest that the core of the transthyretin amyloid fibril is a continuous β-sheet helix. Structure. 1996;4:989–998. doi: 10.1016/s0969-2126(96)00104-9. [DOI] [PubMed] [Google Scholar]
- 199.Serpell LC, Smith JM. Direct visualisation of the β-sheet structure of synthetic Alzheimer’s amyloid. J Mol Biol. 2000;299:225–231. doi: 10.1006/jmbi.2000.3650. [DOI] [PubMed] [Google Scholar]
- 200.Jahn TR, Tennent GA, Radford SE. A common beta-sheet architecture underlies in vitro and in vivo β-2-microglobulin amyloid fibrils. J Biol Chem. 2008;283:17279–17286. doi: 10.1074/jbc.M710351200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Zandomeneghi G, Krebs MR, McCammon MG, Fandrich M. FTIR reveals structural differences between native β-sheet proteins and amyloid fibrils. Protein Sci. 2004;13:3314–3321. doi: 10.1110/ps.041024904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Sarroukh R, Goormaghtigh E, Ruysschaert JM, Raussens V. ATR-FTIR: a “rejuvenated” tool to investigate amyloid proteins. Biochim Biophys Acta. 2013;1828:2328–2338. doi: 10.1016/j.bbamem.2013.04.012. [DOI] [PubMed] [Google Scholar]
- 203.Adler-Abramovich L, et al. Phenylalanine assembly into toxic fibrils suggests amyloid etiology in phenylketonuria. Nat Chem Biol. 2012;8:701–706. doi: 10.1038/nchembio.1002. [DOI] [PubMed] [Google Scholar]
- 204.Julien O, et al. Unraveling the mechanism of cell death induced by chemical fibrils. Nat Chem Biol. 2014;10:969–976. doi: 10.1038/nchembio.1639. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Lazar KL, Miller-Auer H, Getz GS, Orgel JPRO, Meredith SC. Helix-turn-helix peptides that form alpha-helical fibrils: turn sequences drive fibril structure. Biochemistry. 2005;44:12681–12689. doi: 10.1021/bi0509705. [DOI] [PubMed] [Google Scholar]
- 206.Tayeb-Fligelman E, et al. The cytotoxic Staphylococcus aureus PSMa3 reveals a cross-α amyloid-like fibril. Science. 2017;355:831–833. doi: 10.1126/science.aaf4901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Sangwan S, et al. Atomic structure of a toxic, oligomeric segment of SOD1 linked to amyotrophic lateral sclerosis (ALS) Proc Natl Acad Sci USA. 2017;114:8770–8775. doi: 10.1073/pnas.1705091114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Laganowsky A, et al. Atomic view of a toxic amyloid small oligomer. Science. 2012;335:1228–1231. doi: 10.1126/science.1213151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Shirahama T, Cohen AS. Structure of amyloid fibrils after negative staining and high-resolution electron microscopy. Nature. 1965;206:737–738. doi: 10.1038/206737a0. [DOI] [PubMed] [Google Scholar]
- 210.Tattum MH, et al. Elongated oligomers assemble into mammalian PrP amyloid fibrils. J Mol Biol. 2006;357:975–985. doi: 10.1016/j.jmb.2006.01.052. [DOI] [PubMed] [Google Scholar]
- 211.White HE, et al. Globular tetramers of β(2)-microglobulin assemble into elaborate amyloid fibrils. J Mol Biol. 2009;389:48–57. doi: 10.1016/j.jmb.2009.03.066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Paravastua AK, Leapman RD, Yau W, Tycko R. Molecular structural basis for polymorphism in Alzheimer’s β-amyloid fibrils. Proc Natl Acad Sci USA. 2008;105:18349–18354. doi: 10.1073/pnas.0806270105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Lu JX, et al. Molecular structure of beta-amyloid fibrils in Alzheimer’s disease brain tissue. Cell. 2013;154:1257–1268. doi: 10.1016/j.cell.2013.08.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Kajava AV, Baxa U, Wickner RB, Steven AC. A model for Ure2p prion filaments and other amyloids: the parallel superpleated beta-structure. Proc Natl Acad Sci USA. 2004;101:7885–7890. doi: 10.1073/pnas.0402427101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Kajava AV, Aebi U, Steven AC. The parallel superpleated β-structure as a model for amyloid fibrils of human amylin. J Mol Biol. 2005;348:247–252. doi: 10.1016/j.jmb.2005.02.029. [DOI] [PubMed] [Google Scholar]
- 216.Tuttle MD, et al. Solid-state NMR structure of a pathogenic fibril of full-length human alpha-synuclein. Nat Struct Mol Biol. 2016;23:409–415. doi: 10.1038/nsmb.3194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Marshall KE, et al. Characterizing the assembly of the Sup35 yeast prion fragment, GNNQQNY: structural changes accompany a fiber-to-crystal switch. Biophys J. 2010;98:330–338. doi: 10.1016/j.bpj.2009.10.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Reynolds NP, et al. Competition between crystal and fibril formation in molecular mutations of amyloidogenic peptides. Nat Commun. 2017;8:1338. doi: 10.1038/s41467-017-01424-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Saracino GA, Villa A, Moro G, Cosentino U, Salmona M. Spontaneous beta-helical fold in prion protein: the case of PrP(82–146) Proteins. 2009;75:964–976. doi: 10.1002/prot.22306. [DOI] [PubMed] [Google Scholar]
- 220.Kajava AV, Steven AC. β-rolls, β-helices, and other β-solenoid proteins. Adv Protein Chem. 2006;73:55–96. doi: 10.1016/S0065-3233(06)73003-0. [DOI] [PubMed] [Google Scholar]
- 221.Peng Z, Peralta MDR, Toney MD. Extraordinarily stable amyloid fibrils engineered from structurally defined beta-solenoid proteins. Biochemistry. 2017;56:6041–6050. doi: 10.1021/acs.biochem.7b00364. [DOI] [PubMed] [Google Scholar]
- 222.Wolfram F, et al. Catalytic mechanism and mode of action of the periplasmic alginate epimerase AlgG. J Biol Chem. 2014;289:6006–6019. doi: 10.1074/jbc.M113.533158. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Leinala EK, Davies PL, Jia Z. Crystal structure of beta-helical antifreeze protein points to a general ice binding model. Structure. 2002;10:619–627. doi: 10.1016/s0969-2126(02)00745-1. [DOI] [PubMed] [Google Scholar]
- 224.Muller JJ, et al. An intersubunit active site between supercoiled parallel beta helices in the trimeric tailspike endorhamnosidase of Shigella flexneri Phage Sf6. Structure. 2008;16:766–775. doi: 10.1016/j.str.2008.01.019. [DOI] [PubMed] [Google Scholar]
- 225.Kajava AV, Baxa U, Steven AC. Beta arcades: recurring motifs in naturally occurring and disease-related amyloid fibrils. FASEB J. 2010;24:1311–1319. doi: 10.1096/fj.09-145979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Ritter C, et al. Correlation of structural elements and infectivity of the HET-s prion. Nature. 2005;435:844–848. doi: 10.1038/nature03793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Bousset L, et al. Structural and functional characterization of two alpha-synuclein strains. Nat Commun. 2013;4:2575. doi: 10.1038/ncomms3575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Goldsbury CS, et al. Polymorphic fibrillar assembly of human amylin. J Struct Biol. 1997;119:17–21. doi: 10.1006/jsbi.1997.3858. [DOI] [PubMed] [Google Scholar]
- 229.Jiménez J, et al. The protofilament structure of insulin amyloid fibrils. Proc Natl Acad Sci USA. 2002;99:9196–9201. doi: 10.1073/pnas.142459399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Dearborn AD, et al. Alpha-synuclein amyloid fibrils with two entwined, asymmetrically associated protofibrils. J Biol Chem. 2016;291:2310–2318. doi: 10.1074/jbc.M115.698787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231.Andersen CB, et al. Glucagon fibril polymorphism reflects differences in protofilament backbone structure. J Mol Biol. 2010;397:932–946. doi: 10.1016/j.jmb.2010.02.012. [DOI] [PubMed] [Google Scholar]
- 232.Xiao Y, et al. Aβ(1–42) fibril structure illuminates self-recognition and replication of amyloid in Alzheimer’s disease. Nat Struct Mol Biol. 2015;22:499–505. doi: 10.1038/nsmb.2991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Chen B, Thurber KR, Shewmaker F, Wickner RB, Tycko R. Measurement of amyloid fibril mass-per-length by tilted-beam transmission electron microscopy. Proc Natl Acad Sci USA. 2009;106:14339–14344. doi: 10.1073/pnas.0907821106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Doussineau T, et al. Mass determination of entire amyloid fibrils by using mass spectrometry. Angew Chem Int Ed Engl. 2016;55:2340–2344. doi: 10.1002/anie.201508995. [DOI] [PubMed] [Google Scholar]
- 235.Crick FHC, Rich A. Structure of polyglyciene II. Nature. 1955;176:780–781. doi: 10.1038/176780a0. [DOI] [PubMed] [Google Scholar]
- 236.Lee M, et al. Zinc-binding structure of a catalytic amyloid from solid-state NMR. Proc Natl Acad Sci USA. 2017;114:6191–6196. doi: 10.1073/pnas.1706179114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.Viles JH. Metal ions and amyloid fiber formation in neurodegenerative diseases. Copper, zinc and iron in Alzheimer’s, Parkinson’s and prion diseases. Coords Chem Rev. 2012;256:2271–2284. [Google Scholar]
- 238.Gath J, et al. Yet another polymorph of α-synuclein: solid-state sequential assignments. Biomol NMR Assign. 2014;8:395–404. doi: 10.1007/s12104-013-9526-y. [DOI] [PubMed] [Google Scholar]
- 239.Gath J, et al. Unlike twins: an NMR comparison of two α-synuclein polymorphs featuring different toxicity. PLOS One. 2014;9:e90659. doi: 10.1371/journal.pone.0090659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Anfinsen C. Principals that govern the folding of protein chains. Science. 1973;181:223–230. doi: 10.1126/science.181.4096.223. [DOI] [PubMed] [Google Scholar]
- 241.Sidhu A, Segers-Nolten I, Raussens V, Claessens MM, Subramaniam V. Distinct mechanisms determine α-synuclein fibril morphology during growth and maturation. ACS Chem Neurosci. 2017;8:538–547. doi: 10.1021/acschemneuro.6b00287. [DOI] [PubMed] [Google Scholar]
- 242.Eichner T, Radford SE. A diversity of assembly mechanisms of a generic amyloid fold. Mol Cell. 2011;43:8–18. doi: 10.1016/j.molcel.2011.05.012. [DOI] [PubMed] [Google Scholar]
- 243.Weissmann C, Flechsig E. PrP knock-out and PrP transgenic mice in prion research. Br Med Bull. 2003;66:43–60. doi: 10.1093/bmb/66.1.43. [DOI] [PubMed] [Google Scholar]
- 244.Geschwind MD. Prion diseases. Continuum (Minneap Minn) 2015;21:1612–1638. doi: 10.1212/CON.0000000000000251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245.Kitazawa M, Medeiros R, LaFerla FM. Transgenic mouse models of Alzheimer disease: developing a better model as a tool for therapeutic interventions. Curr Pharm Des. 2012;18:1131–1147. doi: 10.2174/138161212799315786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Khalaf O, et al. The H50Q mutation enhances α-synuclein aggregation, secretion, and toxicity. J Biol Chem. 2014;289:21856–21876. doi: 10.1074/jbc.M114.553297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Appel-Cresswell S, et al. Alpha-synuclein p. H50Q, a novel pathogenic mutation for Parkinson’s disease. Mov Disord. 2013;28:811–813. doi: 10.1002/mds.25421. [DOI] [PubMed] [Google Scholar]
- 248.Lesage S, et al. G51D α-synuclein mutation causes a novel parkinsonian-pyramidal syndrome. Ann Neurol. 2013;73:459–471. doi: 10.1002/ana.23894. [DOI] [PubMed] [Google Scholar]
- 249.Nielsen SB, et al. Wildtype and A30P mutant α-synuclein form different fibril structures. PLOS One. 2013;8:e67713. doi: 10.1371/journal.pone.0067713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250.Petrucci S, Ginevrino M, Valente EM. Phenotypic spectrum of α-synuclein mutations: New insights from patients and cellular models. Parkinsonism Relat Disord. 2016;22(Suppl. 1):S16–S20. doi: 10.1016/j.parkreldis.2015.08.015. [DOI] [PubMed] [Google Scholar]
- 251.Schutz AK, et al. Atomic-resolution three-dimensional structure of amyloid beta fibrils bearing the Osaka mutation. Angew Chem Int Ed Engl. 2015;54:331–335. doi: 10.1002/anie.201408598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Andresen JM, et al. The relationship between CAG repeat length and age of onset differs for Huntington’s disease patients with juvenile onset or adult onset. Ann Hum Genet. 2007;71:295–301. doi: 10.1111/j.1469-1809.2006.00335.x. [DOI] [PubMed] [Google Scholar]
- 253.Petkova AT, et al. Self-propagating, molecular-level polymorphism in Alzheimer’s β-amyloid fibrils. Science. 2005;307:262–265. doi: 10.1126/science.1105850. [DOI] [PubMed] [Google Scholar]
- 254.Stöhr j, et al. Purified and synthetic Alzheimer’s amyloid beta (Aβ) prions. Proc Natl Acad Sci USA. 2012;109:11025–11030. doi: 10.1073/pnas.1206555109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Meyer-Luehmann M, et al. Exogenous induction of cerebral β-amyloidogenesis is governed by agent and host. Science. 2006;313:1781–1784. doi: 10.1126/science.1131864. [DOI] [PubMed] [Google Scholar]
- 256.Liu J, et al. Amyloid structure exhibits polymorphism on multiple length scales in human brain tissue. Sci Rep. 2016;6:33079. doi: 10.1038/srep33079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257.Nyström S, et al. Evidence for age-dependent in vivo conformational rearrangement within Aβ amyloid deposits. ACS Chem Biol. 2013;8:1128–1133. doi: 10.1021/cb4000376. [DOI] [PubMed] [Google Scholar]
- 258.Perez-Nievas BG, et al. Dissecting phenotypic traits linked to human resilience to Alzheimer’s pathology. Brain. 2013;136:2510–2526. doi: 10.1093/brain/awt171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Elman JA, et al. Neural compensation in older people with brain amyloid-beta deposition. Nat Neurosci. 2014;17:1316–1318. doi: 10.1038/nn.3806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Strømland Ø, Jakubec M, Furse S, Halskau Ø. Detection of misfolded protein aggregates from a clinical perspective. J Clin Transl Res. 2016;2:11–26. [PMC free article] [PubMed] [Google Scholar]
- 261.Bulawa CE, et al. Tafamidis, a potent and selective transthyretin kinetic stabilizer that inhibits the amyloid cascade. Proc Natl Acad Sci USA. 2012;109:9629–9634. doi: 10.1073/pnas.1121005109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Ludtmann MH, et al. Monomeric alpha-synuclein exerts a physiological role on brain ATP synthase. J Neurosci. 2016;36:10510–10521. doi: 10.1523/JNEUROSCI.1659-16.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Pearson HA, Peers C. Physiological roles for amyloid beta peptides. J Physiol. 2006;575:5–10. doi: 10.1113/jphysiol.2006.111203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264.Barucker C, et al. Abeta42-oligomer interacting peptide (AIP) neutralizes toxic amyloid-beta42 species and protects synaptic structure and function. Sci Rep. 2015;5:15410. doi: 10.1038/srep15410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 265.Du WJ, et al. Brazilin inhibits amyloid beta-protein fibrillogenesis, remodels amyloid fibrils and reduces amyloid cytotoxicity. Sci Rep. 2015;5:7992. doi: 10.1038/srep07992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Jarosz-Griffiths HH, Noble E, Rushworth JV, Hooper NM. Amyloid-beta receptors: the good, the bad, and the prion protein. J Biol Chem. 2016;291:3174–3183. doi: 10.1074/jbc.R115.702704. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267.Mao X, et al. Pathological α-synuclein transmission initiated by binding lymphocyte-activation gene 3. Science. 2016;353:aah3374. doi: 10.1126/science.aah3374. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268.Verma M, Vats A, Taneja V. Toxic species in amyloid disorders: Oligomers or mature fibrils. Ann Indian Acad Neurol. 2015;18:138–145. doi: 10.4103/0972-2327.144284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269.Berry DB, et al. Drug resistance confounding prion therapeutics. Proc Natl Acad Sci USA Proc Natl Acad Sci USA. 2013;110:E4160–E4169. doi: 10.1073/pnas.1317164110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270.Li J, Browning S, Mahal SP, Oelschlegel AM, Weissmann C. Darwinian evolution of prions in cell culture. Science. 2010;327:869–872. doi: 10.1126/science.1183218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271.Oelschlegel AM, Weissmann C. Acquisition of drug resistance and dependence by prions. PLOS Pathog. 2013;9:e1003158. doi: 10.1371/journal.ppat.1003158. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.Seidler PM, et al. Structure-based inhibitors of tau aggregation. Nat Chem. 2018;10:170–176. doi: 10.1038/nchem.2889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273.Sievers SA, et al. Structure-based design of non-natural amino-acid inhibitors of amyloid fibril formation. Nature. 2011;475:96–100. doi: 10.1038/nature10154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 274.Jiang L, et al. Structure-based discovery of fiber-binding compounds that reduce the cytotoxicity of amyloid beta. eLife. 2013;2:e00857. doi: 10.7554/eLife.00857. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 275.Kad NM, et al. Hierarchical assembly of β2-microglobulin amyloid in vitro revealed by atomic force microscopy. J Mol Biol. 2003;330:785–797. doi: 10.1016/s0022-2836(03)00583-7. [DOI] [PubMed] [Google Scholar]
- 276.Watanabe-Nakayama T, et al. High-speed atomic force microscopy reveals structural dynamics of amyloid beta1-42 aggregates. Proc Natl Acad Sci USA. 2016;113:5835–5840. doi: 10.1073/pnas.1524807113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277.Silvers R, et al. Aggregation and fibril structure of AβMO1-42 and Aβ1-42. Biochemistry. 2017;56:4850–4859. doi: 10.1021/acs.biochem.7b00729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 278.Colvin MT, et al. Atomic resolution structure of monomorphic Aβ42 amyloid fibrils. J Am Chem Soc. 2016;138:9663–9674. doi: 10.1021/jacs.6b05129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279.Schütz AK, et al. Binding of polythiophenes to amyloids: structural mapping of the pharmacophore. ACS Chem Neurosci. 2017;9:475–481. doi: 10.1021/acschemneuro.7b00397. [DOI] [PubMed] [Google Scholar]
- 280.Ries J, et al. Superresolution imaging of amyloid fibrils with binding-activated probes. ACS Chem Neurosci. 2013;4:1057–1061. doi: 10.1021/cn400091m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 281.Choi JH, May BC, Wille H, Cohen FE. Molecular modeling of the misfolded insulin subunit and amyloid fibril. Biophys J. 2009;97:3187–3195. doi: 10.1016/j.bpj.2009.09.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282.Ladner CL, et al. Stacked sets of parallel, in-register beta-strands of beta2-microglobulin in amyloid fibrils revealed by site-directed spin labeling and chemical labeling. J Biol Chem. 2010;285:17137–17147. doi: 10.1074/jbc.M110.117234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 283.Zhang Y, et al. Pulsed hydrogen-deuterium exchange mass spectrometry probes conformational changes in amyloid beta (Abeta) peptide aggregation. Proc Natl Acad Sci USA. 2013;110:14604–14609. doi: 10.1073/pnas.1309175110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 284.Der-Sarkissian A, Jao CC, Chen J, Langen R. Structural organization of alpha-synuclein fibrils studied by site-directed spin labeling. J Biol Chem. 2003;278:37530–37535. doi: 10.1074/jbc.M305266200. [DOI] [PubMed] [Google Scholar]
- 285.Varkey J, Langen R. Membrane remodeling by amyloidogenic and non-amyloidogenic proteins studied by EPR. J Magn Reson. 2017;280:127–139. doi: 10.1016/j.jmr.2017.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286.Cohen AS, Calkins E. Electron microscopic observations on a fiberous component in amyloid of diverse origins. Nature. 1959;183:1202–1203. doi: 10.1038/1831202a0. [DOI] [PubMed] [Google Scholar]
- 287.Cohen AS, Shirahama T. High resolution electron microscopic analysis of the amyloid fibril. J Cell Bio. 1967;33:679. doi: 10.1083/jcb.33.3.679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288.Astbury WT. X-ray studies of protein structure. Cold Spring Harb Symp Quant Biol. 1934;2:15–27. [Google Scholar]
- 289.Jimenez JL, Tennent G, Pepys M, Saibil HR. Structural diversity of ex vivo amyloid fibrils studied by cryo-electron microscopy. J Mol Biol. 2001;311:241–247. doi: 10.1006/jmbi.2001.4863. [DOI] [PubMed] [Google Scholar]
- 290.Schaffer J, et al. Recombinant versus natural human 111In-beta2-microglobulin for scintigraphic detection of Abeta2m amyloid in dialysis patients. Kidney Int. 2000;58:873–880. doi: 10.1046/j.1523-1755.2000.00237.x. [DOI] [PubMed] [Google Scholar]
- 291.Pras M, Schubert M, Zucker-Franklin D, Rimon A, Franklin EC. The characterization of soluble amyloid prepared in water. J Clin Invest. 1968;47:924–933. doi: 10.1172/JCI105784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 292.Fändrich M, Dobson CM. The behaviour of polyamino acids reveals an inverse side chain effect in amyloid structure formation. EMBO J. 2002;21:5682–5690. doi: 10.1093/emboj/cdf573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293.Sikorski P, Atkins E. New model for crystalline polyglutamine assemblies and their connection with amyloid fibrils. Biomacromolecules. 2005;6:425–432. doi: 10.1021/bm0494388. [DOI] [PubMed] [Google Scholar]
- 294.Ranson N, Stromer T, Bousset L, Melki R, Serpell LC. Insights into the architecture of the Ure2p yeast protein assemblies from helical twisted fibrils. Protein Sci. 2006;15:2481–2487. doi: 10.1110/ps.062215206. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 295.Graeber MB, Kösel S, Grasbon-Frodl E, Möller HJ, Mehraein P. Histopathology and APOE genotype of the first Alzheimer disease patient, Auguste D. Neurogenetics. 1998;1:223–228. doi: 10.1007/s100480050033. [DOI] [PubMed] [Google Scholar]
- 296.Nicoll AJ, et al. Amyloid-beta nanotubes are associated with prion protein-dependent synaptotoxicity. Nat Commun. 2013;4:2416. doi: 10.1038/ncomms3416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297.Martins IC, et al. Lipids revert inert Aβ amyloid fibrils to neurotoxic protofibrils that affect learning in mice. EMBO J. 2008;27:224–233. doi: 10.1038/sj.emboj.7601953. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 298.Jakhria T, et al. beta2-microglobulin amyloid fibrils are nanoparticles that disrupt lysosomal membrane protein trafficking and inhibit protein degradation by lysosomes. J Biol Chem. 2014;289:35781–35794. doi: 10.1074/jbc.M114.586222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 299.Fujioka S, et al. Update on novel familial forms of Parkinson’s disease and multiple system atrophy. Parkinsonism Relat Disord. 2014;20:S29–S34. doi: 10.1016/S1353-8020(13)70010-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 300.Pagano G, Ferrara N, Brooks DJ, Pavese N. Age at onset and Parkinson disease phenotype. Neurology. 2016;86:1400–1407. doi: 10.1212/WNL.0000000000002461. [DOI] [PMC free article] [PubMed] [Google Scholar]