1,083
Views
37
CrossRef citations to date
0
Altmetric
Extra View

Assessing the causes and consequences of co-polymerization in amyloid formation

, &
Pages 359-368 | Received 12 Aug 2013, Accepted 07 Sep 2013, Published online: 11 Sep 2013

Abstract

How, and why, different proteins form amyloid fibrils is most often studied in vitro using a single purified protein sequence. However, many amyloid diseases involve co-aggregation of different protein species, including proteins with/without post-translational modifications (e.g., different strains of PrP), proteins of different length (e.g., β2-microglobulin and ΔN6, Aβ40, and Aβ42), sequence variants (e.g., Aβ and AβARC), and proteins from different organisms (e.g., bovine PrP and human PrP). The consequences of co-aggregation of different proteins upon the structure, stability, species transmission and toxicity of the resulting amyloid aggregates is discussed here, including the role of co-aggregation in expanding the repertoire of oligomeric and fibrillar structures and how this can affect their biological and biophysical properties.

This article refers to:

Introduction

It has been well documented that amyloid fibrils can be highly polymorphic, even when formed from proteins with the same amino acid sequence under the same incubation conditions.Citation1 This polymorphism can be manifested through the formation of fibrils with different numbers and orientations of protofilaments, and also within the protofilament substructure, for example in the residues that contribute to the β-strand segments within the fibrils (reviewed in refs. Citation2Citation3). In a comprehensive study of an 11-residue transthyretin (TTR) peptide the structure and orientation of protofilaments from three different fibril polymorphs was described in atomic detail using cryo-electron microscopy (EM), scanning transmission EM and magic angle spinning (MAS) NMR, revealing different numbers of protofilaments within each polymorph.Citation4 The different packing arrangements available for an amyloid fibril core has also been explored using X-ray crystallography of 4–7-residue peptides from a variety of amyloidogenic proteins, the results revealing eight classes of polymorphs arranged as different steric zippers.Citation5 The structures of the steric zippers differ in whether the β-strands are parallel or anti-parallel, whether the β-sheets pack face-to-face or back-to-back, and whether the sheets orientate themselves up or down relative to each otherCitation6 (see ).

Figure 1. The potential repertoire of fibril polymorphs. (A) The eight classes of steric zipper available to a single polypeptide chain (re-drawn from Figure 4 of Sawaya et al.).Citation5 (B) A schematic of different possible fibril polymorphs that can arise from co-aggregation of more than one protein (depicted are ΔN6 [PDB code 2XKU] in pink and β2m [PDB code 2XKS] in purple).Citation20

Figure 1. The potential repertoire of fibril polymorphs. (A) The eight classes of steric zipper available to a single polypeptide chain (re-drawn from Figure 4 of Sawaya et al.).Citation5 (B) A schematic of different possible fibril polymorphs that can arise from co-aggregation of more than one protein (depicted are ΔN6 [PDB code 2XKU] in pink and β2m [PDB code 2XKS] in purple).Citation20

Oligomeric intermediates of amyloid formation can also show polymorphism. HypF-N is the 91-residue N-terminal domain of HypF, a carbamoyl transferase in Escherichia coli.Citation7 In a recent study, oligomers of HypF-N formed under different solution conditions were shown to be similar, morphologically and tinctorially, but differed in their ability to cause cell dysfunction.Citation8 Further investigation using site-specific labeling with pyrene maleimide and binding of the dye 8-anilinonaphthalene-1-sulfonic acid (ANS) demonstrated that the oligomers that possess less structural order and expose more hydrophobic surface area are more toxic than oligomers with a more ordered core. The increased toxicity of one oligomer morphology over the other is thought to arise because the structural flexibility and hydrophobic exposure of the toxic oligomers allows them to cross the hydrophobic bilayer of the cell membrane, whereas the non-toxic oligomers can bind to, but not cross the membrane. In other examples the more structured oligomers were shown to be the most toxic.Citation9 These examples, and many others,Citation10-Citation12 show the ability of a single polypeptide sequence to adopt an array of oligomeric structures with differing physical, chemical, biophysical and biological functions, even for species of similar molecular weight. This poses significant challenges for the identification and structural characterization of each individual species in the amyloid cascade and in amyloid disease.

The above studies demonstrate that there is a large range of possible polymorphic structures available to an amyloidogenic protein or peptide both as oligomeric intermediates in fibril formation and in the final fibrillar form. The potential assortment of polymorphs can be greatly expanded when mixtures of proteins can co-aggregate, either de novo or via cross-seeding. The resulting increased range of fibril structures expands the repertoire of aggregation propensities and amyloid stabilities. This can be useful, for example by providing enhanced opportunity for tailoring the physical properties of amyloid fibrils for use as nanomaterials. However, it may also be detrimental, for example, by increasing the possibility for toxicity and disease.

Here we discuss the repertoire of amyloid polymorphs formed from different protein and peptide sequences, (the interactions between amyloidogenic proteins and other important factors in amyloid fibril formation, such as chaperones, small molecules, metal ions, lipids, and glycolipids, are not included).Citation13 We identify three different scenarios that can affect the likelihood of two proteins cross-seeding and/or co-polymerizing with each other. First, we discuss the requirements of sequence in determining co-aggregation propensity and its outcomes. Second, we describe scenarios in which one protein increases the amyloidogenic propensity of another. Finally, we review the extent to which both proteins need to have the ability to adopt a similar fibril structure in order to co-assemble. A select number of studies are discussed in detail in each section, and a more comprehensive list of protein-protein interactions that expand the repertoire of fibril polymorphs is included in .

Table 1. Interactions between amyloidogenic proteins

Condition 1: Co-Aggregation of Proteins/Peptides with Similar Primary Sequence

The intermolecular interactions that stabilize amyloid fibrils involve the peptide backbone and thus the ability to form fibrils is considered to be a generic property of all polypeptide chains.Citation14 However side-chain interactions play a crucial role in the formation of amyloid fibrils and in determining the ability of one protein to seed polymerization of another. The importance of amino acid sequence similarity in cross-seeding (adding fibrillar seeds of one polypeptide sequence to monomers of a second polypeptide sequence) was shown in a study of hen lysozyme (Row 1 in ).Citation15 Fibrillar seeds created from sequences that are 99.2% (I55T mutational variant of hen lysozyme) or 95% identical (turkey lysozyme) to that of the hen lysozyme monomer produced identical seeding behavior to homologous seeding. A seed with a sequence that is 60% identical to the hen lysozyme monomer (human lysozyme) showed faster fibril elongation than the rate of fibril formation of the unseeded monomer, but with a lag phase that is increased compared with self-seeding. Sequences with 36% identity or no identity to the hen lysozyme monomer had no effect on fibril formation.Citation15 Thus the efficiency of seeding depends on the similarity between both the seed’s and monomer’s amino acid sequence.Citation15

Similar primary sequences between amyloidogenic proteins may promote the likelihood of a cross-seeding event occurring, but this does not always result in a reciprocal ability for both partners to cross-seed each other’s assembly. Islet amyloid polypeptide (IAPP), the amyloidogenic peptide involved in type II diabetes mellitus, is derived from the 89 amino acid precursor protein pro-islet amyloid polypeptide. Aβ, the neurotoxic agent in Alzheimer disease (AD), is derived from the proteolytic processing of amyloid precursor protein which, depending on the splicing isoform, is composed of 365 to 770 amino acids. These precursor proteins are unrelated in sequence and have no obvious functional relationship. Despite this, residues 15–37 of Aβ (residues involved in the amyloid core of Aβ fibrilsCitation16) share 39% sequence identity and 65% sequence similarity with residues 10–33 of IAPP. The interaction between IAPP and Aβ40 was studied to determine whether cross-seeding could occur.Citation17 These in vitro studies showed that Aβ40 fibrils, IAPP fibrils, and IAPP amorphous aggregates are all equally effective at seeding elongation with IAPP monomer. Conversely, IAPP aggregates will not cross-seed polymerization of Aβ40, and IAPP fibrils are much less efficient at seeding Aβ40 monomer compared with Aβ40 fibrils, exhibiting only 2% of the seeding efficiency of Aβ fibrils on a weight basis.Citation17 Thus, there is a surprising lack of equivalence in cross-seeding of Aβ40 and IAPP (Row 2 of ).

The importance of primary sequence in determining the ability of proteins to co-aggregate is more complex than both partners simply having amyloid-prone primary sequences. In one example, the ability of a variety of different amyloid fibrils to act as a seed for elongation with Aβ40 was compared with the efficiency of homologous seeding (Aβ40 seeds).Citation17 The majority of fibril types, including fibrils formed from the yeast protein Ure2p and the human protein β2-microglobulin (β2m), were as inefficient at cross-seeding Aβ40 as non-amyloid protein aggregates such as collagen or denatured ovalbumin (Row 3 of ). Thus simply being amyloidogenic does not necessarily result in an ability to seed elongation with other non-homologous amyloidogenic proteins; sequence similarity is required as well.

Sequence similarity is not the only important factor in determining the propensity for co-aggregation in amyloid formation as the precise positioning of compatible residues within the structure also plays a significant role. For example, co-incubation of murine and human Aβ40 and Aβ42 peptides revealed that interspecies fibrils will form for both Aβ40 and Aβ42 alloforms (Row 4 of )Citation18 despite the three residue difference in the mouse and human amino acid sequences. This is because the residue differences between human and murine Aβ (Arg5Gly, Tyr10Phe, and His13Arg) are located toward the N-terminal region of the peptides, a region not thought to take part in the β-structure of Aβ fibrils.Citation16 Interestingly the mixed fibrils containing murine and human Aβ42 were more stable in solubilizing buffers than the homopolymeric human Aβ42 fibrils.Citation18 Murine Aβ peptide contains a larger amount of β-turn structure and is more strongly stabilized by hydrogen bonds than human Aβ.Citation19 The increased number of hydrogen bonds may stabilize the mixed fibrils, resulting in a greater resistance to buffer solubilization compared with homopolymeric human fibrils. This study demonstrates how mixing proteins with different primary sequences can result in the emergence of new fibril polymorphs with different biophysical properties.

The ability of human Aβ and its murine homolog to co-polymerize is consistent with the other studies outlined above, as reflected by the percentage of overall sequence similarity between the two peptides (~93% identity for Aβ40 and Aβ42). Other studies show that the presence of a non-homologous amyloidogenic protein can inhibit fibril formation because both the number and position of differences in the primary sequences are incompatible with fibril elongation and/or nucleation, e.g., human and murine β2m (70% sequence similarity).Citation20 In an elegant study in which heterotetramers of transythyretin (TTR) were formed by mixing its human and murine subunits (Row 5 of ) Kelly and coworkers demonstrated that incorporation of murine TTR subunits protects the amyloidogenic human TTR from aggregating by stabilizing the native state.Citation21,Citation22 Similar results were observed for the mutant Thr119Met of human TTR which protects wild-type TTR tetramers from dissociation both in vitro and in vivo (Row 6 of ).Citation23 In both cases amyloid formation is inhibited via kinetic stabilization of the resulting heterotetramers relative to their homotetrameric human counterparts. Thus, using variants of amyloidogenic sequences allows amyloid formation to be controlled and/or inhibited by introducing a non-homologous primary sequence into a fibril formation reaction.

Condition 2: Co-Aggregation Occurs by One Partner Protein Affecting the Rate of Fibril Formation of Another

Amyloidoses can occur when susceptible proteins are exposed to conditions that promote global unfolding of the native stateCitation24 or enhance the population of amyloidogenic intermediates.Citation25 Statistically, mixed fibril formation is less likely than the generation of fibrils from a single sequence, as the production of heterofibrils requires a change in conformation of two proteins instead of one. However, this assumes that both partners must undergo a change to an amyloid conformer independently. By contrast with this observation, monomer-monomer interactions can promote protein unfolding, and/or formation of an amyloidogenic fold.Citation20,Citation26 Thus if one protein partner acts to enhance the amyloid potential of another, the probability of forming heteropolymeric assemblies is increased.

An excellent example of one protein enhancing the amyloid propensity of another has been shown for β2m and its N-terminally truncated variant ΔN6.Citation20 β2m forms amyloid fibrils in vivo resulting in the disease dialysis-related amyloidosis (DRA).Citation27 However, in the absence of other co-factors or co-solvents human β2m will not form fibrils in vitro at neutral pH within an experimentally tractable timescale,Citation28,Citation29 while ΔN6 is highly aggregation prone.Citation20,Citation30 Surprisingly, when monomeric human β2m is mixed with monomeric ΔN6 fibril formation from both proteins occurs, resulting in heteropolymeric fibrils, even at neutral pH (Row 7 of ).Citation20,Citation30 Interestingly ΔN6 was shown to stimulate fibril formation of human β2m at sub-stoichiometric levels, with fibril formation occurring at a 1:99 ΔN6:β2m molar ratio.Citation20 An atomistic description of how ΔN6 converts β2m into an amyloidogenic form has been provided using NMR.Citation20 These studies showed that collision with ΔN6 results in increased conformational dynamics of human β2m, allowing β2m to undergo further structural rearrangements, critically via cis-trans isomerization of Pro32, that subsequently allows access to the amyloid state.

By enhancing the amyloid potential of β2m, both β2m and ΔN6 are incorporated into heteropolymeric fibrils when incubated at a 1:1 ratio.Citation30 As discussed in the introduction, aggregation of a single polypeptide sequence can result in a range of fibril structures dependent on the experimental conditions employed, such as pH, temperature, ionic strength and agitation.Citation31 In the case of mixed fibrils, for example those formed from β2m and ΔN6 where both proteins are incorporated into the fibril structure, the possibilities for polymorphism are vast. Perhaps unsurprisingly then, the heteropolymeric fibrils formed from β2m and ΔN6 do not resemble those formed from β2m alone at acidic pH or from ∆N6 alone at neutral pH. Instead biophysical experiments have shown that the heteropolymeric fibrils of β2m/ΔN6 form a unique polymorph that is thermodynamically less stable than both β2m and ∆N6 homopolymeric fibrils.Citation30 These findings show that the heteropolymorphic fibrils do not possess “average” structural or thermodynamic characteristics resulting from mixing of the two precursors. Instead, through the enhancement of the amyloidogenic potential of one protein by another, a new area of the protein (mis)folding landscape is sampled, and a new polymorph is created.

The ability to endow an amyloid-prone form on a less amyloidogenic protein can also result in altered stabilities of on-pathway intermediates. The Arctic mutation (E22G) of Aβ40 results in a highly amyloidogenic protein (Aβ40ARC) that forms both protofibrils and fibrils under conditions in which Aβ40 remains monomeric.Citation32 Interestingly in the absence of Aβ40, Aβ40ARC protofibrils will convert to fibrils, but in the presence of Aβ40 at a 1:1 molar ratio, the two proteins form mixed protofibrils that are incapable of further assembly. The mixed protofibrils are more stable than the protofibrils formed by Aβ40ARC alone and hence these species accumulate (Row 8 in ). Lashuel et al. postulate that these mixed protofibrils may be responsible for the enhancement of neurotoxicity and accelerated disease progression that is seen in patients carrying this mutation.Citation32 Co-aggregation of two different peptides, therefore, can alter the kinetics of fibril formation, and thereby alter the morphology and stability of intermediates, resulting in a prolonged lifetime of toxic species.

A highly specific enhancement of one protein’s amyloidogenic potential by another is demonstrated by the co-incubation of tau with a familial mutant of α-synuclein (Ala53Thr). This amino acid substitution results in increased α-synuclein and tau inclusions, suggesting that it not only increases the propensity for α-synuclein to form fibrils, but also promotes the formation of tau inclusions.Citation33 Further in vitro studies have shown that co-incubation of tau and α-synuclein induces polymerization of both proteins and this effect was observed for all six tau isoforms (Row 9 in ).Citation33 Immunogold labeling revealed that the individual fibrils are mostly homopolymers, but some fibrils were labeled for both α-synuclein and tau but in spatially separate domains, suggesting that end-to-end annealing of the homopolymers occurs.Citation33 Furthermore, at low concentrations of α-synuclein, the presence of tau was required for α-synuclein fibrils to form. Similarly, tau requires α-synuclein pre-aggregation to induce the requisite conformational change to allow tau inclusion formation. Interestingly, this effect was specific to α-synuclein; other amyloidogenic proteins such as Aβ40 are unable to initiate tau polymerization.Citation33

Enhancing a protein’s amyloid potential not only results in increased formation of fibrils, but new species in the amyloidogenic pathway can also form by mixing protein monomers. Aβ has been shown to promote the aggregation and toxicity of α-synuclein in a dose- and time-dependent manner.Citation34 The two proteins may also interact in vivo as evidenced by an overlap in the pathology of AD and Parkinson Disease in Lewy Body Disease (LBD), where the initial signs are dementia followed by parkinsonism.Citation35 In samples of human brains from sufferers of LBD Aβ and α-synuclein were shown to co-immunoprecipitate, but no interaction between the two proteins was observed in non-demented controls. In vitro studies have shown that solubilized Aβ42 induces formation of α-synuclein tetramers and higher-order oligomers (Row 10 in ).Citation36 Electron microscopy images of the mixture after 6 h show well-defined hybrid ring oligomeric structures 9–15 nm in diameter, which may form functional cation nanopores.Citation36 Mixing two amyloidogenic proteins, therefore, can change the stability and characteristics of species that form during amyloid assembly, in this case resulting in the formation of a hybrid nanopore that could be involved in disease progression.

Condition 3: Structure and Post-Translational Modifications Affect the Efficiency of Cross-Seeding

One of the most striking examples of post-translational modification altering protein co-aggregation is in prion diseases. In these disorders the efficiency of cross-species infectivity is partly determined by primary sequence, but also relies on compatible structures and post-translational modifications (typically glycosylation).Citation37 Furthermore, to efficiently cross-seed, both the seed and the monomer must be able to adopt the same amyloidogenic conformation.

The importance of structure in prion propagation was demonstrated with the bovine spongiform encephalopathy (BSE) outbreak in the UK in the mid-1980s which resulted in the infection of more than 2 million UK cattle.Citation37 This was followed by an outbreak of a novel human prion disease, variant Creutzfeldt-Jakob disease (vCJD) in the mid-1990s. CJD is not a new disease: “classical” CJD prions, formed from spontaneous misfolding of the prion protein (PrP), occur at a rate of approximately one case per 1 million population per year.Citation37 However, despite sharing identical primary sequences, classical CJD prions and vCJD prions behave very differently in their ability to convert PrPC (the non-infectious form of PrP) to PrPSc (the prion form) in a new species. This was shown when human classical CJD prions were used to infect wild-type mice, wherein no transmission was observed. However, when human vCJD prions were used instead, transmission was much more successful and the resultant mouse PrPSc showed faithful strain replication of the vCJD strain.Citation38 Fascinatingly, this shows that two proteins with identical amino acid sequences can experience different species barriers, confirming that prion propagation is conformation dependent (Row 11 in ).

Although the vCJD and classical CJD strains of PrP have identical primary sequences, they differ in their glycosylation patterns. Classical CJD strains are modified predominantly with low molecular mass glycosylation, whereas the vCJD strain has high molecular mass glycosylation.Citation26 Furthermore, the ratio of the three PrP glycoforms (unglycosylated, monoglycosylated, and diglycosylated) is faithfully maintained on passage in transgenic mice expressing human PrP.Citation26 Similarly, transmission of human prions and bovine prions to wild-type mice results in murine PrPSc with glycoform ratios that correspond to the initial inocula.Citation26 Interestingly, PrPC glycosylation occurs before conversion to PrPSc, thus the different glycoforms may be determined by initial PrP conformation. Different cell types may also glycosylate proteins differently, thus particular PrPSc glycoform strains may replicate most favorably in cell types with a similar PrPC glycoform expressed on the cell surface.Citation37 Glycosylation similarities (or differences) between PrPC and the infectious prion may help to explain the different incubation periods seen in individuals with prion disease.Citation26 Thus post-translational modifications can be crucial in determining the propensity of a protein to cross-seed, as well as affecting the location and timescale of disease developing.

In a recent study the importance of strain for cross-species infectivity was demonstrated using two morphologies of hamster prion protein, R and S.Citation39 The two strains were formed using the same stock of recombinant PrP under identical solvent conditions, but with different agitation modes; the S-strain was produced under shaking, the R-strain was produced with rotation. The R- and S-fibrils display substantial differences in their cross-β cores, epitope exposure and morphology.Citation40 Despite sharing the same amino acid sequence, the difference in morphology between the R- and S-fibrils has a profound effect on their ability to reproduce their strain faithfully in a different species (mouse). The hamster R-fibrils successfully recruit mouse-PrP into fibrils and maintain their morphology, thus acting both as a so-called “catalyst” and a template for fibril formation of mouse PrP. By contrast, the hamster S-fibril form also acts as a catalyst for structural conversion of mouse-PrP, but does not act as a template, i.e., the strain morphology was not propagated (Row 12 in ).Citation39 In the latter case, fibril elongation occurred but involved mouse-PrP switching to a new conformational state that resembled the R-fibril form. This demonstrates that primary sequence (condition 1) is not wholly responsible for the ability to act as an enhancer of amyloidogenicity in another protein (condition 2). Instead structure also affects the ability to transmit disease across a species barrier.

Yeast prion proteins have also provided a powerful model for developing understanding of both sequence and strain effects in prion propagation. When the yeast protein Rnq1 is aggregated into its prion form, it can promote prion formation of two other proteins, Sup35 and Ure2p.Citation41 Interestingly the efficiency with which the prion form of Rnq1 can cross-seed the latter two proteins is dependent on its strain. One Rnq1 prion strain will seed Ure2p inefficiently, but will seed Sup35 efficiently. However an alternative Rnq1 prion strain behaves in the opposite manner, inefficiently seeding Sup35, but efficiently seeding Ure2p. Both Rnq1 prion strains can propagate their morphology by homologous seeding, yet their ability to cross-seed other proteins is entirely dependent on their structure, rather than their (identical) amino acid sequence (Row 13 in ).Citation41 Further revelations from yeast models show that not all prions can cross-seed, and some can even inhibit homologous seeding.Citation41 For example, the prion form of Sup35 will inhibit Ure2p prion seeds from seeding non-prion Ure2p, through “poisoning” of the Ure2p prion seeds (Row 14 in ).Citation41

The importance of structure in amyloid propagation is not limited to prion interactions. Despite Aβ40 and Aβ42 differing by only two residues, the two peptides have different effects on each other’s aggregation propensity. In co-polymerization experiments Aβ42 monomers stimulate Aβ40 monomers to aggregate, whereas Aβ40 monomers shows a concentration-dependent inhibitory effect on Aβ42 fibril formation (Row 15 in ).Citation42 However, the same effect is not seen in cross-seeding experiments (Row 16 in ),Citation42 where one partner is in fibrillar form and one partner is monomeric. In this case Aβ42 fibrils are less efficient than fibrillar Aβ40 at inducing fibril formation of monomeric Aβ40. Why does Aβ42 efficiently stimulate Aβ40 amyloidogenesis only when it is monomeric, and not when in the fibrillar form? This phenomenon is explained by the importance of structure, as well as primary sequence for promoting amyloid formation. Accordingly, it is Aβ42 oligomers, rather than fibrils, that are the optimal template for incorporating Aβ40 monomers, and these are mainly populated during incubation of soluble Aβ40 and Aβ42 (co-polymerization), rather than when Aβ42 is added as fibrils (cross-seeding). Thus, although primary sequence is important in allowing Aβ42 and Aβ40 co-aggregation to occur (condition 1), and Aβ42 is an enhancer of Aβ40 amyloidogenesis (condition 2), the inclusion of the structure of Aβ42 aggregates must also be considered in assessing the likelihood of co-aggregation occurring.

A final example of a structure-dependent interaction of two amyloidogenic proteins is between PrPc and Aβ. Interactions between Aβ42 oligomers and PrPC leads to the inhibition of long-term potentiation in hippocampal slices from mice.Citation43 Interestingly, experiments using NMR have shown that PrPc does not interact with disordered monomeric Aβ40, but Aβ40 must first change its conformation, possibly forming a misfolded monomer or dimer.Citation44 These studies not only highlight the importance of structure in determining co-aggregation, but also show that interactions are not limited to the monomer-monomer or monomer-fibril stages in fibril formation. Instead, recognition between two amyloidogenic proteins can be specific for a key intermediary structure in the amyloid formation pathway.

Conclusions

This review broadly categorizes interactions between two amyloidogenic proteins into “conditions” that prescribe the likelihood of a co-polymerization or cross-seeding event occurring. The first condition is that both partners must have a similar amino acid sequence. Although the importance of the protein backbone in fibril stability means most proteins may have the propensity to form fibrils under defined conditions,Citation14 side-chain packing also plays a vital role.Citation5 Interestingly the ability to cross-seed is not necessarily reciprocal, even between closely related amyloid proteins.Citation17 The second condition is that at least one of the partners must enhance the amyloidogenic potential of the other, such as the effect of ∆N6 on β2mCitation20,Citation30 or α-synuclein on tau.Citation33 Finally, the third condition is that structure and/or post-translational modifications can affect the efficiency of cross-seeding, for example in prion propagation across different species. The three conditions are not mutually exclusive: primary sequence can influence fibril/oligomer structure, and a similarity in sequences between amyloidogenic proteins with their less amyloidogenic partners is important in defining the ability of proteins to co-polymerize. The network of hydrogen bonds within a fibril is known to be vital for monomer addition, and this would be influenced by both side-chain and fold. The conditions outlined here are also not comprehensive: another property that can regulate the propensity for co-polymerization is similar aggregation kinetics of the co-incubated monomers (Row 18 in ).Citation45 Additionally the ratio of the two components may also be vital: under normal physiological conditions the Aβ42:Aβ40 ratio in the brain is ~1:9.Citation46 However, in familial AD patients a ratio of 3:7 Aβ42:Aβ40 is observed, suggesting that a change in the proportion of the two alloforms may be significant in the development of AD.Citation47

Examples of co-polymerization between unrelated sequences in vivo are relatively rare. The BSE crisis followed by the outbreak of vCJD has shown that cross-species prion transmission can occur.Citation37 An increased risk of AD has been shown for type II diabetes sufferers, suggesting a possible link between Aβ and IAPP aggregation.Citation48 In fact, Aβ seems to be promiscuous, with in vivo interactions reported between Aβ and cystatin c (Row 19 of )Citation49,Citation50, TTR (Row 20 of )Citation51, and neuroserpins (Rows 21 and 22 of ).Citation52 The lack of more examples in vivo suggests that perhaps all three conditions described here need to be met, in a timely and spatially defined manner, for co-polymerization and/or cross-seeding to occur. Additionally the rapid increase in examples reported in the literature of co-polymerization and cross-seeding between different amyloidogenic proteins in vitro confirms the possibility for co-polymerization, giving weight to a possible in vivo relevance to at least some of these interactions.

Interactions between unrelated, or sequence distinct amyloidogenic proteins are fascinating and important because of the possibilities for polymorphism that arise from co-aggregation. In the introduction of this review the eight different classes of steric zippers that comprise the amyloid core were discussedCitation5 and detailed studies of polymorphism in homopolymeric fibrils have been performed.Citation1,Citation4,Citation53 Additionally polymorphism can occur at the earliest stages of aggregation, with subsets of oligomers showing structural polymorphism that also affects the ability of the species to be toxic (reviewed in ref. Citation54). Polymorphism has been induced in proteins with the same sequence through alteration of solution conditions such as pH, agitation, temperature and/or ionic strength.Citation31,Citation55 Fibril length can also be controlled via fragmentation, allowing the physical dimensions and surface interactions of fibrils to be studied.Citation56 However, more parameters are also open to manipulation in systems that co-aggregate, such as the ratio of the protein species involved and the identity of the species that are mixed (monomers, oligomers or fibrils). By adjusting these different factors, in addition to altering experimental constraints, the potential repertoire of species that can form, both at the level of oligomers and fibrils is vast (see the schematic in ). Furthermore, as all proteins are potentially able to form amyloid,Citation14 and some amyloidogenic proteins can cross-seed unrelated polypeptide sequences, the risk of co-polymerization could potentially extend to the entire proteome. Unravelling the different structural possibilities for oligomer and fibril architectures and defining their physical and biological characteristics will be challenging. However these multi-component assemblies will need to be defined for developing our understanding of the causative agents of amyloid disease, as well as to harness the future potentials of amyloid fibrils as designer nanomaterials.

Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

Acknowledgments

This work was supported by Medical Research Council Grant G0900958. We thank the Radford group for helpful discussions and critical reading of the manuscript.

Note Added in Proof

Recent work has shown that a specific and transient protofibrillar species of Abeta42, with a novel triple helical structure, is the most potent aggregate involved in the interaction between PrP and Abeta.66 This works supports the importance of structure for influencing amyloid interactions between different proteins.

10.4161/pri.26415

References

  • Meinhardt J, Sachse C, Hortschansky P, Grigorieff N, Fändrich M. Abeta(1-40) fibril polymorphism implies diverse interaction patterns in amyloid fibrils. J Mol Biol 2009; 386:869 - 77; http://dx.doi.org/10.1016/j.jmb.2008.11.005; PMID: 19038266
  • Eichner T, Radford SE. A diversity of assembly mechanisms of a generic amyloid fold. Mol Cell 2011; 43:8 - 18; http://dx.doi.org/10.1016/j.molcel.2011.05.012; PMID: 21726806
  • Fändrich M, Meinhardt J, Grigorieff N. Structural polymorphism of Alzheimer Abeta and other amyloid fibrils. Prion 2009; 3:89 - 93; http://dx.doi.org/10.4161/pri.3.2.8859; PMID: 19597329
  • Fitzpatrick AW, Debelouchina GT, Bayro MJ, Clare DK, Caporini MA, Bajaj VS, Jaroniec CP, Wang L, Ladizhansky V, Müller SA, et al. Atomic structure and hierarchical assembly of a cross-β amyloid fibril. Proc Natl Acad Sci U S A 2013; 110:5468 - 73; http://dx.doi.org/10.1073/pnas.1219476110; PMID: 23513222
  • Sawaya MR, Sambashivan S, Nelson R, Ivanova MI, Sievers SA, Apostol MI, Thompson MJ, Balbirnie M, Wiltzius JJ, McFarlane HT, et al. Atomic structures of amyloid cross-beta spines reveal varied steric zippers. Nature 2007; 447:453 - 7; http://dx.doi.org/10.1038/nature05695; PMID: 17468747
  • Eisenberg D, Jucker M. The amyloid state of proteins in human diseases. Cell 2012; 148:1188 - 203; http://dx.doi.org/10.1016/j.cell.2012.02.022; PMID: 22424229
  • Paschos A, Bauer A, Zimmermann A, Zehelein E, Böck A. HypF, a carbamoyl phosphate-converting enzyme involved in [NiFe] hydrogenase maturation. J Biol Chem 2002; 277:49945 - 51; http://dx.doi.org/10.1074/jbc.M204601200; PMID: 12377778
  • Campioni S, Mannini B, Zampagni M, Pensalfini A, Parrini C, Evangelisti E, Relini A, Stefani M, Dobson CM, Cecchi C, et al. A causative link between the structure of aberrant protein oligomers and their toxicity. Nat Chem Biol 2010; 6:140 - 7; http://dx.doi.org/10.1038/nchembio.283; PMID: 20081829
  • Cremades N, Cohen SI, Deas E, Abramov AY, Chen AY, Orte A, Sandal M, Clarke RW, Dunne P, Aprile FA, et al. Direct observation of the interconversion of normal and toxic forms of α-synuclein. Cell 2012; 149:1048 - 59; http://dx.doi.org/10.1016/j.cell.2012.03.037; PMID: 22632969
  • Barghorn S, Nimmrich V, Striebinger A, Krantz C, Keller P, Janson B, Bahr M, Schmidt M, Bitner RS, Harlan J, et al. Globular amyloid beta-peptide oligomer - a homogenous and stable neuropathological protein in Alzheimer’s disease. J Neurochem 2005; 95:834 - 47; http://dx.doi.org/10.1111/j.1471-4159.2005.03407.x; PMID: 16135089
  • Bernstein SL, Dupuis NF, Lazo ND, Wyttenbach T, Condron MM, Bitan G, Teplow DB, Shea JE, Ruotolo BT, Robinson CV, et al. Amyloid-β protein oligomerization and the importance of tetramers and dodecamers in the aetiology of Alzheimer’s disease. Nat Chem 2009; 1:326 - 31; http://dx.doi.org/10.1038/nchem.247; PMID: 20703363
  • Ono K, Condron MM, Teplow DB. Structure-neurotoxicity relationships of amyloid beta-protein oligomers. Proc Natl Acad Sci U S A 2009; 106:14745 - 50; http://dx.doi.org/10.1073/pnas.0905127106; PMID: 19706468
  • McLaurin J, Yang DS, Yip CM, Fraser PE. Review: modulating factors in amyloid-beta fibril formation. J Struct Biol 2000; 130:259 - 70; http://dx.doi.org/10.1006/jsbi.2000.4289; PMID: 10940230
  • Dobson CM. Protein misfolding, evolution and disease. Trends Biochem Sci 1999; 24:329 - 32; http://dx.doi.org/10.1016/S0968-0004(99)01445-0; PMID: 10470028
  • Krebs MR, Morozova-Roche LA, Daniel K, Robinson CV, Dobson CM. Observation of sequence specificity in the seeding of protein amyloid fibrils. Protein Sci 2004; 13:1933 - 8; http://dx.doi.org/10.1110/ps.04707004; PMID: 15215533
  • Petkova AT, Ishii Y, Balbach JJ, Antzutkin ON, Leapman RD, Delaglio F, Tycko R. A structural model for Alzheimer’s beta -amyloid fibrils based on experimental constraints from solid state NMR. Proc Natl Acad Sci U S A 2002; 99:16742 - 7; http://dx.doi.org/10.1073/pnas.262663499; PMID: 12481027
  • O’Nuallain B, Williams AD, Westermark P, Wetzel R. Seeding specificity in amyloid growth induced by heterologous fibrils. J Biol Chem 2004; 279:17490 - 9; http://dx.doi.org/10.1074/jbc.M311300200; PMID: 14752113
  • Fung J, Frost D, Chakrabartty A, McLaurin J. Interaction of human and mouse Abeta peptides. J Neurochem 2004; 91:1398 - 403; http://dx.doi.org/10.1111/j.1471-4159.2004.02828.x; PMID: 15584916
  • Hilbich C, Kisters-Woike B, Reed J, Masters CL, Beyreuther K. Human and rodent sequence analogs of Alzheimer’s amyloid beta A4 share similar properties and can be solubilized in buffers of pH 7.4. Eur J Biochem 1991; 201:61 - 9; http://dx.doi.org/10.1111/j.1432-1033.1991.tb16256.x; PMID: 1915378
  • Eichner T, Kalverda AP, Thompson GS, Homans SW, Radford SE. Conformational conversion during amyloid formation at atomic resolution. Mol Cell 2011; 41:161 - 72; http://dx.doi.org/10.1016/j.molcel.2010.11.028; PMID: 21255727
  • Miroy GJ, Lai Z, Lashuel HA, Peterson SA, Strang C, Kelly JW. Inhibiting transthyretin amyloid fibril formation via protein stabilization. Proc Natl Acad Sci U S A 1996; 93:15051 - 6; http://dx.doi.org/10.1073/pnas.93.26.15051; PMID: 8986762
  • Reixach N, Foss TR, Santelli E, Pascual J, Kelly JW, Buxbaum JN. Human-murine transthyretin heterotetramers are kinetically stable and non-amyloidogenic. A lesson in the generation of transgenic models of diseases involving oligomeric proteins. J Biol Chem 2008; 283:2098 - 107; http://dx.doi.org/10.1074/jbc.M708028200; PMID: 18006495
  • Hammarström P, Schneider F, Kelly JW. Trans-suppression of misfolding in an amyloid disease. Science 2001; 293:2459 - 62; http://dx.doi.org/10.1126/science.1062245; PMID: 11577236
  • Johnson SM, Wiseman RL, Sekijima Y, Green NS, Adamski-Werner SL, Kelly JW. Native state kinetic stabilization as a strategy to ameliorate protein misfolding diseases: a focus on the transthyretin amyloidoses. Acc Chem Res 2005; 38:911 - 21; http://dx.doi.org/10.1021/ar020073i; PMID: 16359163
  • Jahn TR, Parker MJ, Homans SW, Radford SE. Amyloid formation under physiological conditions proceeds via a native-like folding intermediate. Nat Struct Mol Biol 2006; 13:195 - 201; http://dx.doi.org/10.1038/nsmb1058; PMID: 16491092
  • Collinge J, Sidle KC, Meads J, Ironside J, Hill AF. Molecular analysis of prion strain variation and the aetiology of ‘new variant’ CJD. Nature 1996; 383:685 - 90; http://dx.doi.org/10.1038/383685a0; PMID: 8878476
  • Gejyo F, Yamada T, Odani S, Nakagawa Y, Arakawa M, Kunitomo T, Kataoka H, Suzuki M, Hirasawa Y, Shirahama T, et al. A new form of amyloid protein associated with chronic hemodialysis was identified as beta 2-microglobulin. Biochem Biophys Res Commun 1985; 129:701 - 6; http://dx.doi.org/10.1016/0006-291X(85)91948-5; PMID: 3893430
  • Calabrese MF, Eakin CM, Wang JM, Miranker AD. A regulatable switch mediates self-association in an immunoglobulin fold. Nat Struct Mol Biol 2008; 15:965 - 71; http://dx.doi.org/10.1038/nsmb.1483; PMID: 19172750
  • Platt GW, Radford SE. Glimpses of the molecular mechanisms of beta2-microglobulin fibril formation in vitro: aggregation on a complex energy landscape. FEBS Lett 2009; 583:2623 - 9; http://dx.doi.org/10.1016/j.febslet.2009.05.005; PMID: 19433089
  • Sarell CJ, Woods LA, Su Y, Debelouchina GT, Ashcroft AE, Griffin RG, Stockley PG, Radford SE. Expanding the repertoire of amyloid polymorphs by co-polymerization of related protein precursors. J Biol Chem 2013; 288:7327 - 37; http://dx.doi.org/10.1074/jbc.M112.447524; PMID: 23329840
  • Petkova AT, Leapman RD, Guo Z, Yau WM, Mattson MP, Tycko R. Self-propagating, molecular-level polymorphism in Alzheimer’s beta-amyloid fibrils. Science 2005; 307:262 - 5; http://dx.doi.org/10.1126/science.1105850; PMID: 15653506
  • Lashuel HA, Hartley DM, Petre BM, Wall JS, Simon MN, Walz T, Lansbury PT Jr.. Mixtures of wild-type and a pathogenic (E22G) form of Abeta40 in vitro accumulate protofibrils, including amyloid pores. J Mol Biol 2003; 332:795 - 808; http://dx.doi.org/10.1016/S0022-2836(03)00927-6; PMID: 12972252
  • Giasson BI, Forman MS, Higuchi M, Golbe LI, Graves CL, Kotzbauer PT, Trojanowski JQ, Lee VMY. Initiation and synergistic fibrillization of tau and alpha-synuclein. Science 2003; 300:636 - 40; http://dx.doi.org/10.1126/science.1082324; PMID: 12714745
  • Masliah E, Rockenstein E, Veinbergs I, Sagara Y, Mallory M, Hashimoto M, Mucke L. beta-amyloid peptides enhance alpha-synuclein accumulation and neuronal deficits in a transgenic mouse model linking Alzheimer’s disease and Parkinson’s disease. Proc Natl Acad Sci U S A 2001; 98:12245 - 50; http://dx.doi.org/10.1073/pnas.211412398; PMID: 11572944
  • Galasko D, Hansen LA, Katzman R, Wiederholt W, Masliah E, Terry R, Hill LR, Lessin P, Thal LJ. Clinical-neuropathological correlations in Alzheimer’s disease and related dementias. Arch Neurol 1994; 51:888 - 95; http://dx.doi.org/10.1001/archneur.1994.00540210060013; PMID: 8080388
  • Tsigelny IF, Crews L, Desplats P, Shaked GM, Sharikov Y, Mizuno H, Spencer B, Rockenstein E, Trejo M, Platoshyn O, et al. Mechanisms of hybrid oligomer formation in the pathogenesis of combined Alzheimer’s and Parkinson’s diseases. PLoS One 2008; 3:e3135; http://dx.doi.org/10.1371/journal.pone.0003135; PMID: 18769546
  • Collinge J. Prion diseases of humans and animals: their causes and molecular basis. Annu Rev Neurosci 2001; 24:519 - 50; http://dx.doi.org/10.1146/annurev.neuro.24.1.519; PMID: 11283320
  • Hill AF, Desbruslais M, Joiner S, Sidle KC, Gowland I, Collinge J, Doey LJ, Lantos P. The same prion strain causes vCJD and BSE. Nature 1997; 389:448 - 50, 526; http://dx.doi.org/10.1038/38925; PMID: 9333232
  • Makarava N, Ostapchenko VG, Savtchenko R, Baskakov IV. Conformational switching within individual amyloid fibrils. J Biol Chem 2009; 284:14386 - 95; http://dx.doi.org/10.1074/jbc.M900533200; PMID: 19329794
  • Makarava N, Baskakov IV. The same primary structure of the prion protein yields two distinct self-propagating states. J Biol Chem 2008; 283:15988 - 96; http://dx.doi.org/10.1074/jbc.M800562200; PMID: 18400757
  • Bradley ME, Edskes HK, Hong JY, Wickner RB, Liebman SW. Interactions among prions and prion “strains” in yeast. Proc Natl Acad Sci U S A 2002; 99:Suppl 4 16392 - 9; http://dx.doi.org/10.1073/pnas.152330699; PMID: 12149514
  • Pauwels K, Williams TL, Morris KL, Jonckheere W, Vandersteen A, Kelly G, Schymkowitz J, Rousseau F, Pastore A, Serpell LC, et al. Structural basis for increased toxicity of pathological aβ42:aβ40 ratios in Alzheimer disease. J Biol Chem 2012; 287:5650 - 60; http://dx.doi.org/10.1074/jbc.M111.264473; PMID: 22157754
  • Laurén J, Gimbel DA, Nygaard HB, Gilbert JW, Strittmatter SM. Cellular prion protein mediates impairment of synaptic plasticity by amyloid-beta oligomers. Nature 2009; 457:1128 - 32; http://dx.doi.org/10.1038/nature07761; PMID: 19242475
  • Younan ND, Sarell CJ, Davies P, Brown DR, Viles JH. The cellular prion protein traps Alzheimer’s Aβ in an oligomeric form and disassembles amyloid fibers. FASEB J 2013; 27:1847 - 58; http://dx.doi.org/10.1096/fj.12-222588; PMID: 23335053
  • MacPhee CE, Dobson CM. Formation of mixed fibrils demonstrates the generic nature and potential utility of amyloid nanostructures. J Am Chem Soc 2000; 122:12707 - 13; http://dx.doi.org/10.1021/ja0029580
  • Suzuki N, Cheung TT, Cai XD, Odaka A, Otvos L Jr., Eckman C, Golde TE, Younkin SG. An increased percentage of long amyloid beta protein secreted by familial amyloid beta protein precursor (beta APP717) mutants. Science 1994; 264:1336 - 40; http://dx.doi.org/10.1126/science.8191290; PMID: 8191290
  • Scheuner D, Eckman C, Jensen M, Song X, Citron M, Suzuki N, Bird TD, Hardy J, Hutton M, Kukull W, et al. Secreted amyloid beta-protein similar to that in the senile plaques of Alzheimer’s disease is increased in vivo by the presenilin 1 and 2 and APP mutations linked to familial Alzheimer’s disease. Nat Med 1996; 2:864 - 70; http://dx.doi.org/10.1038/nm0896-864; PMID: 8705854
  • Janson J, Laedtke T, Parisi JE, O’Brien P, Petersen RC, Butler PC. Increased risk of type 2 diabetes in Alzheimer disease. Diabetes 2004; 53:474 - 81; http://dx.doi.org/10.2337/diabetes.53.2.474; PMID: 14747300
  • Kaeser SA, Herzig MC, Coomaraswamy J, Kilger E, Selenica ML, Winkler DT, Staufenbiel M, Levy E, Grubb A, Jucker M. Cystatin C modulates cerebral beta-amyloidosis. Nat Genet 2007; 39:1437 - 9; http://dx.doi.org/10.1038/ng.2007.23; PMID: 18026102
  • Mi W, Pawlik M, Sastre M, Jung SS, Radvinsky DS, Klein AM, Sommer J, Schmidt SD, Nixon RA, Mathews PM, et al. Cystatin C inhibits amyloid-beta deposition in Alzheimer’s disease mouse models. Nat Genet 2007; 39:1440 - 2; http://dx.doi.org/10.1038/ng.2007.29; PMID: 18026100
  • Li X, Masliah E, Reixach N, Buxbaum JN. Neuronal production of transthyretin in human and murine Alzheimer’s disease: is it protective?. J Neurosci 2011; 31:12483 - 90; http://dx.doi.org/10.1523/JNEUROSCI.2417-11.2011; PMID: 21880910
  • Fabbro S, Schaller K, Seeds NW. Amyloid-beta levels are significantly reduced and spatial memory defects are rescued in a novel neuroserpin-deficient Alzheimer’s disease transgenic mouse model. J Neurochem 2011; 118:928 - 38; http://dx.doi.org/10.1111/j.1471-4159.2011.07359.x; PMID: 21689108
  • Jiménez JL, Nettleton EJ, Bouchard M, Robinson CV, Dobson CM, Saibil HR. The protofilament structure of insulin amyloid fibrils. Proc Natl Acad Sci U S A 2002; 99:9196 - 201; http://dx.doi.org/10.1073/pnas.142459399; PMID: 12093917
  • Glabe CG. Structural classification of toxic amyloid oligomers. J Biol Chem 2008; 283:29639 - 43; http://dx.doi.org/10.1074/jbc.R800016200; PMID: 18723507
  • Gosal WS, Morten IJ, Hewitt EW, Smith DA, Thomson NH, Radford SE. Competing pathways determine fibril morphology in the self-assembly of beta2-microglobulin into amyloid. J Mol Biol 2005; 351:850 - 64; http://dx.doi.org/10.1016/j.jmb.2005.06.040; PMID: 16024039
  • Xue WF, Hellewell AL, Hewitt EW, Radford SE. Fibril fragmentation in amyloid assembly and cytotoxicity: when size matters. Prion 2010; 4:20 - 5; http://dx.doi.org/10.4161/pri.4.1.11378; PMID: 20305394
  • Guo JL, Covell DJ, Daniels JP, Iba M, Stieber A, Zhang B, Riddle DM, Kwong LK, Xu Y, Trojanowski JQ, et al. Distinct α-synuclein strains differentially promote tau inclusions in neurons. Cell 2013; 154:103 - 17; http://dx.doi.org/10.1016/j.cell.2013.05.057; PMID: 23827677
  • Chiou A, Hägglöf P, Orte A, Chen AY, Dunne PD, Belorgey D, Karlsson-Li S, Lomas DA, Klenerman D. Probing neuroserpin polymerization and interaction with amyloid-beta peptides using single molecule fluorescence. Biophys J 2009; 97:2306 - 15; http://dx.doi.org/10.1016/j.bpj.2009.07.057; PMID: 19843463
  • Kinghorn KJ, Crowther DC, Sharp LK, Nerelius C, Davis RL, Chang HT, Green C, Gubb DC, Johansson J, Lomas DA. Neuroserpin binds Abeta and is a neuroprotective component of amyloid plaques in Alzheimer disease. J Biol Chem 2006; 281:29268 - 77; http://dx.doi.org/10.1074/jbc.M600690200; PMID: 16849336
  • Ridgley DM, Ebanks KC, Barone JR. Peptide mixtures can self-assemble into large amyloid fibers of varying size and morphology. Biomacromolecules 2011; 12:3770 - 9; http://dx.doi.org/10.1021/bm201005k; PMID: 21879764
  • Jarrett JT, Berger EP, Lansbury PT Jr.. The carboxy terminus of the beta amyloid protein is critical for the seeding of amyloid formation: implications for the pathogenesis of Alzheimer’s disease. Biochemistry 1993; 32:4693 - 7; http://dx.doi.org/10.1021/bi00069a001; PMID: 8490014
  • Nussbaum JM, Schilling S, Cynis H, Silva A, Swanson E, Wangsanut T, Tayler K, Wiltgen B, Hatami A, Rönicke R, et al. Prion-like behaviour and tau-dependent cytotoxicity of pyroglutamylated amyloid-β. Nature 2012; 485:651 - 5; http://dx.doi.org/10.1038/nature11060; PMID: 22660329
  • Ono K, Takahashi R, Ikeda T, Yamada M. Cross-seeding effects of amyloid β-protein and α-synuclein. J Neurochem 2012; 122:883 - 90; http://dx.doi.org/10.1111/j.1471-4159.2012.07847.x; PMID: 22734715
  • Willander H, Presto J, Askarieh G, Biverstål H, Frohm B, Knight SD, Johansson J, Linse S. BRICHOS domains efficiently delay fibrillation of amyloid β-peptide. J Biol Chem 2012; 287:31608 - 17; http://dx.doi.org/10.1074/jbc.M112.393157; PMID: 22801430
  • Yan LM, Velkova A, Tatarek-Nossol M, Andreetto E, Kapurniotu A. IAPP mimic blocks Abeta cytotoxic self-assembly: cross-suppression of amyloid toxicity of Abeta and IAPP suggests a molecular link between Alzheimer’s disease and type II diabetes. Angew Chem Int Ed Engl 2007; 46:1246 - 52; http://dx.doi.org/10.1002/anie.200604056; PMID: 17203498
  • Nicoll AJ, Panico S, Freir DB, Wright D, Terry C, Risse E, Herron CE, O’Malley T, Wadsworth JD, Farrow MA, et al. Amyloid-β nanotubes are associated with prion protein-dependent synaptotoxicity. Nat Commun 2013; 4:2416; http://dx.doi.org/10.1038/ncomms3416; PMID: 24022506