Open access peer-reviewed chapter

Neuroprotective Function of Non-Proteolytic Amyloid-β Chaperones in Alzheimer’s Disease

Written By

Bhargy Sharma and Konstantin Pervushin

Submitted: 14 December 2018 Reviewed: 08 January 2019 Published: 01 February 2019

DOI: 10.5772/intechopen.84238

From the Edited Volume

Amyloid Diseases

Edited by Dmitry Kurouski

Chapter metrics overview

1,070 Chapter Downloads

View Full Metrics

Abstract

This chapter attempts to explore protective role of chaperone proteins in the neurodegenerative diseases caused by amyloidosis. These chaperones prevent amyloid pathology either directly, through chemical interactions with amyloidogenic species to mediate their refolding, solubilization and degradation, or indirectly, by scavenging reactive oxygen species produced as by-products of amyloid aggregation. Here we focus on structural and morphological changes during aggregation of amyloids which have been identified using Nuclear magnetic resonance spectroscopy, X-ray crystallography, Electron microscopy, Atomic force microscopy and other biophysical techniques as well as interactions between chaperone proteins and amyloid moieties. Non-proteolytic chaperones mediate amyloid clearance and metabolism through conformational changes due to proximity binding. In this chapter, we delineate these interactions as well as the molecular mechanism of chaperones used to sequester ROS products of amyloidosis with focus on amyloid-β peptides associated with the Alzheimer’s disease.

Keywords

  • amyloid-β
  • structural biology
  • aggregation
  • chaperone
  • neuroprotection
  • Alzheimer’s disease

1. Introduction

Abnormal deposition of amyloids or “Amyloidosis” is hallmark of several chronic cerebrovascular diseases including neurodegeneration culminating into dementia. Efforts to develop targeted drugs against amyloids have been hindered since there is no universal mechanism that leads to protein misfolding or aggregation, and the aggregates usually do not correspond directly to clinical symptoms of the diseases. A clearer understanding of molecular interactions of amyloids can drive the ongoing therapeutic efforts to prevent aggregation of nascent amyloids into pathological species and to design timely interventions. In addition to aging, precursor mutations in genes and proteins, gene multiplication, expansion of amyloidogenic sequences, and xenobiotics such as air pollutants are risk factors usually associated with amyloidosis disorders [1]. There are different amyloidogenic species causing a variety of neuropathic diseases such as Alzheimer’s disease (AD), Parkinson’s disease (PD), Poly-glutamine disorders like Huntington disease (HD), Prion diseases including Creutzfeldt-Jakob disease, Lewy body disease, Amyotrophic Lateral Sclerosis (ALS) as well as metabolic diseases such as type II diabetes (T2D) and corneal dystrophy to name a few [2]. AD is a pandemic form of dementia caused due to improper aggregation of amyloidogenic proteins-amyloid-β (Aβ), which is a cleavage product of amyloid precursor protein (APP) and tau, stabilizes microtubules in neurons [3]. Prognosis is also closely associated with aggregation of α-Synuclein (αS) protein into Lewy bodies usually concentrated in presynaptic terminals [4]. Human islet amyloid polypeptide (IAPP), also known as amylin, is secreted along with insulin from pancreatic β-islet cells. IAPP rich amyloid plaques, facilitated by insulin resistance, are hallmarks of T2D [5]. β-sheeted infectious isoform of cellular prion protein (PrP) causes transmissible spongiform encephalopathy (TSE), broadly known as prion disease [2].

In this chapter, we review various structures and conformations attained by Aβ peptides during the process of aggregation. We begin with introducing different morphologies and conformations attained by amyloids during this process. We outline mechanisms of amyloid pathology, either directly mediated by aggregates or indirectly through generation of oxidative stress. We have briefly alluded to chaperone functions of Heat shock protein (Hsp) family and their interactions to different amyloid structures as well as intracellular protease mechanisms. It is important to note that such intracellular proteases are almost entirely inefficient in dealing with large insoluble plaques. Here we emphasize on additional endogenous proteins which show potential as chaperones, albeit through different mechanisms. Towards the end, we discuss the role of oxidative stress in accelerating AD pathology and the effect of Aβ interactions with metal species. Mechanisms to alleviate oxidative stress and their possible protective role in AD have been discussed.

Advertisement

2. Structural and biophysical basis of amyloid aggregation

Membrane glycoprotein APP is cleaved by β-secretase and γ-secretase enzymes generating 36–49 residue long peptides, among which amyloidogenic Aβ(1–40) and Aβ(1–42) are well-known culprits in AD [6]. Unstructured monomeric amyloids polymerize to form fibrillar structures with characteristic cross-β morphology formed by in-register β-sheets which align parallel to the fibrillar axis whereas perpendicularly extended side-chains pack closely to form tight steric zippers [2]. Nuclear magnetic resonance (NMR) spectroscopy studies on amyloid monomers or fibrillar structures to identify characteristic dynamic structural features. X-ray crystallographic studies have helped in identifying structural motifs through fiber diffraction and mass-per-length studies through EM provide information regarding specific symmetries identified in filaments. NMR models of Aβ monomeric peptides show predominantly α-helices with propensity to convert into β-sheets [6]. The α-helical conformation is also observed among on-pathway transient intermediates, possibly mediated by interactions of hydrophilic N-terminal residues [7]. These intermediates then give way to β-sheet conformation in higher oligomeric species which eventually transition into mature fibrils. In vitro aggregation of Aβ monomers into fibrils can be carefully curated to obtained different fibril morphology. In solid state nuclear magnetic resonance (ssNMR) spectra, cross-peaks originating from hydrophobic core and C-terminal regions within amyloid fibrils, prepared without seeding indicate presence of polymorphs. Formation of homogenous fibrils from Aβ(1–42) seeds precludes aggregation of Aβ(1–40) [8].

αS binds to neuronal membranes in a highly α-helical state as opposed to the unfolded monomeric conformation [7]. The antipathic N-terminal sequence has tendency to form α-helical conformation, central region from aa 61–95 is highly hydrophobic and amyloidogenic, and the C-terminal domain provides flexibility to the protein without attaining any specific structure [7]. Prefibrillar αS shows conformational plasticity and the soluble monomers combine to form unstable dimeric molecules which further aggregate into higher oligomers and fibrils. IAPP structure in detergent micelles shows a helical conformation kinked around residue His18 at neutral pH and an extended α-helix in acidic conditions; thus the ability to attain specific conformation also depends on its chemical environment [5].

2.1 Morphological differences in amyloid aggregates

Disordered or misfolded monomers can convert into a range of amyloid species, e.g., spherical oligomers, amorphous aggregates, annular oligomers, protofibrils, inclusion bodies or insoluble fibrils, based on the pathological pathways (Table 1). Dynamic intermediate oligomers form transient structures facilitating on-pathway conversion from misfolded monomers to fibrils. The spherical and annular aggregates get generated off-pathway and contribute highly to amyloid mediated neurotoxicity [2, 9]. The Lewy bodies formed by αS are example of intracellular amyloid inclusions [10]. Formation of a specific aggregated form is usually controlled by process of nucleation. In primary nucleation, seeds are formed by spontaneous aggregation of monomers followed by fibril elongation, whereas fragments from mature fibrils recruit monomers to facilitate the polymerization in secondary nucleation process. However, this process is specific to monomeric species; seeding with Aβ(1–42) fibrils does not contribute to aggregation of Aβ(1–40) [8]. Comparison between kinetic profiles of seeded and non-seeded Aβ(1–40) suggest that aggregation follows the seeded fibrils template [8]. Monomeric Aβ(1–40) samples form striated ribbons under constant agitation, and twisted fibrils under undisturbed conditions [11]. NMR can also be used to study kinetics of oligomer interactions by utilizing the constant exchange of detectable monomers with other invisible oligomeric species, through saturation transfer difference (STD) experiments [12]. Amyloid fibrils show protease resistance and are insoluble in detergents. However, their affinity to small lipophilic molecules is amenable for detection through biophysical assays using Congo red and thioflavin dyes [13]. These fluorescent dyes can be used to quantify β-sheet rich amyloid fibrils under laboratory conditions where fluorescent intensity linearly correlates with fibril formation [14]. Morphology of amyloid fibrils can be studied using transmission electron microscopy (TEM) by observing diffraction patterns. Whether fibrils form a ribbon-like or a striated pattern, can be calculated by mass-per-length constraints [15]. Different amyloid structures can be categorized based on symmetry and periodicity with these methods [16]. Aβ40 fibrils structure have either 2-fold (2A) or 3-fold (3Q) rotational symmetry with a twisted morphology with periodicity of 120 ± 20 nm, and roughly 8 nm width [2, 11]. IAPP fibrils also lead to formation of striated and twisted ribbons [11].

Amyloid structureStructure/symmetryToxicityRelevanceRefs.
MonomerDisordered, random coilNon-toxicOn pathway*[20]
Small oligomersRich in β-sheet (spherical, annular, ADDLs)ToxicOn pathway[21, 22]
Amyloid inclusionsAmorphous aggregatesNon-toxicOff pathway[2, 23]
ProtofilamentsHomogenous, up to 200 nm longToxicOn pathway[24]
FibrilsCross-βNon-toxicOn pathway[25]
PlaqueAmorphous meshworks, fibril bundles, amyloid starsNon-toxic[26]

Table 1.

Summary of amyloid structures observed in AD.

On-pathway structures are intermediates involved in aggregation of amyloids into fibrils.


2.2 Structural properties of amyloid aggregates

Basic structural architecture of amyloid fibrils consists of a characteristic 4.7 Å repeat through the cross-β structure [2]. Still, there exist some variations in morphology of different Aβ fibrils; the C-terminus in Aβ(1–40) is hidden within the core, while the corresponding residues are exposed on the surface of Aβ(1–42) [10]. Solid state NMR studies of amyloid fibril structures employ 2D dipolar-assisted rotation resonance (DARR) experiments and frequency selective rotational-echo double resonance (REDOR) experiments along with distance measurements between 13C-13C and 13C-15N nuclei [8]. Studies of Aβ(1–42) fibrils using ssNMR suggested unidirectional protofibril growth with two molecules coming together to form parallel intermolecular β-sheets. The ssNMR structure of homogeneous Aβ(1–42) fibrils (PDB ID: 2MXU) exhibits triple-β-strands encompassing residues 12–18, 24–33 and 36–40 respectively, connected by coil-and-turns at residues 19–23 and 34–42 [8]. The other structure of Aβ(1–42) fibrils with oxidized Met35 (PDB:2BEG), suggests structural inhomogeneity in N-terminal residues from 1 to 17 and presence of two β-strands in residues 18–26 and 31–42, respectively [17]. The AD related Aβ(1–42) polymorph shows double-horseshoe like cross-β structure where N terminus of monomeric Aβ(1–42) has L shape and the C terminus takes an S-shape. The structure comprises five in-register parallel strands with each fibril layer comprising of two molecules with hydrophobic side chains buried maximally [16]. In addition to hydrophobic core, parallel β-sheets also show polar zipper interactions through intermolecular hydrogen bonds. A recent 4.0 Å resolution ssNMR-cryoEM hybrid structure (PDB ID: 5OQV) of intertwined Aβ(1–42) protofilaments showed an approximate 21 screw helical symmetry with 4.67 Å rise [18]. This implicated step-wise shift between the subunits is similar to tau dimers. Three hydrophobic clusters involving Ala, Val, Phe, Leu, Ile, and Met residues expand along the fibril axis and contribute to overall stability of fibrils [18]. Tertiary interactions in Aβ(1–42) fibrils are different compared to Aβ(1–40) fibrils owing to differences in side-chain packaging in hydrophobic core of protofilaments [17]. The N-terminus of Aβ(1–40) peptides is disordered and highly prone to proteolysis whereas the ordered region from tyrosine residue onwards acquires a double-layered structure with a “β-arch” motif where two β-strands are separated by a short loop [19]. This motif is stabilized by formation of a salt bridge between Asp23 and Lys28 across the bend. Interactions between Lys28 and Ala42 observed in these fibrils differ to those found in Aβ(1–40) [8]. Similarly, salt bridges between Asp1 and Lys28, Asp7 and Arg5, Glu11 and His6 and His13, and Asn-Gln ladders further contribute to stability of the Aβ(1–42) fibril structure [16, 18]. Point mutations introduced in basic amino acid sequences lead to varied fibril architecture compared to the wild-type Aβ40 fibrils [2].

NMR structure of tau monomers shows detectable propensity to β-sheet, poly-proline helices and transient α-helical conformations. Aggregation of the R3 fragment consisting of 26 amino acid residues is strongly associated with formation of fibrils in the presence of polyanions as well as during self-assembly of pristine tau [7]. Phosphorylation of serine and threonine residues stabilize the α-helical conformation. αS fibrils acquire an overall assembly that mimics a Greek key as seen in atomic resolution structure of purified protein [7].

2.3 Role of amyloidosis in AD pathology

Amyloidosis in cerebrovascular system is mediated through aggregation of wild type proteins as the consequence of multivalent interactions in intrinsically disordered proteins or regions of proteins, mutations in amyloidogenic precursors, expansion of repeats in the amyloidogenic sequences, actions of proteases or chemical modifications on the precursor sequences, overexpression of the precursor, liquid–liquid phase separation, actions of small metabolites or age-related cell death [27, 28, 29, 30, 31]. The number of human proteins capable of causing amyloidosis has exceeded 50 and are associated with many neurological disorders based of variations in the disease precursors [2]. Different oligomeric and morphologically distinct species affect cells and tissues in different manner. Amyloid oligomers can be cytotoxic with the long-term potentiation disrupting membranes to cause ion permeability and homeostasis imbalance [2]. Fully matured fibrils contribute to the disease by disturbing lipid membranes and thus interfere with the general membrane bound cellular organelles, or by forming physical barriers which disrupt inter-cellular communications [2]. Structural inhomogeneity was observed in fibrils from the different clinical variants of amyloids in AD. Differences in morphology of Aβ aggregates can change the phenotype observed in different forms of AD. While a single Aβ structure is found in brain cortex in the cases of typical prolonged duration AD, different polymorphs are found in the cases of rapidly progressive form of AD [32]. There are six isoforms of tau protein in brain and CNS, longest one containing a stretch of 441 amino acids, with high tendency for phosphorylation and self-assembly. Tau filaments cause different tauopathies, including neurofibrillary tangles in AD [7]. Neurofibrillary tangles and neuropil threads are predominantly tau aggregates localized in neuronal cell bodies and processes, respectively, whereas the mature Aβ plaques are present extracellularly [33].

2.4 Toxicity of amyloid aggregates

Aβ(1–42) is more potent fibrillogenic Aβ variant in human brain, yet its concentration accounts for only about 5–10% of Aβ(1–40) concentration produced [33]. Accumulation of Aβ(1–42) in brain and cerebrovascular system due to its aggregation propensity, and by extension, the relative ratio of Aβ(1–42): Aβ(1–40) is a biomarker for AD [33]. Aβ(1–40) fibrils are less toxic compared to Aβ(1–42) fibrils which show differential toxicity and Aβ(1–43) fibrils are most cytotoxic [10]. Soluble Aβ in the form of low-n oligomers, Aβ-derived diffusible ligands and protofibrils are among major toxic contributors towards AD pathology [34]. Interactions of oligomeric species with other molecular complexes, metal ions and cellular membranes also impact the extent of their toxic effects in brain. N-terminus truncated Aβ peptides (Aβ(n–42), where n can vary from 2 to 11) are highly toxic and are usually present in Aβ deposits in AD brain [7]. Peptides cleaved at positions 3 or 11 are prone to pyroglutamylation and prominent components of Aβ deposits, likely owing to their resistance to proteolytic degradation [35]. αS is a major component of Lewy bodies and neurites causing neurotoxicity through dopaminergic mechanism and these aggregates are key identifiers for PD and Lewy body dementia [7]. Detection of αS aggregates early in peripheral nervous system can serve as early biomarkers before motor disabilities develop in PD patients [36]. Horizontal transfer of amyloid aggregates and oligomers between cells also contributes to propagation of disease pathology [37]. IAPP oligomers can impair insulin secretion in pancreatic cells and cause cell death and cellular uncoupling [7].

Advertisement

3. β-Amyloid chaperones

The main function of molecular chaperones is to facilitate their protein targets acquire proper functional conformation or fold into correct oligomeric assemblies [38]. Several proteins have been shown to bind misfolded amyloids and inhibit amyloid aggregation or promote refolding of polymorphs. They usually play other physiological roles, yet show tendency to prevent amyloidosis under specific stimuli. We will discuss some of these non-proteolytic chaperones in more detail below. Additionally, albumin binding to polymeric Aβ, α1-antitrypsin, immunoglobulins A and G are endogenous human cerebrospinal fluid (CSF) proteins with capability to control amyloid formation, although the inhibitory activity is lesser by two orders of magnitude compared to plasma [39].

3.1 Heat shock proteins

Heat shock proteins (Hsp) are the most commonly known chaperones in human body, divided into five classes Hsp70, Hsp90, Hsp60, Hsp40, and small Hsps, differing in protein size. Chaperones in Hsp60 and Hsp70 family mediate folding, Hsp33 family hold partially folded proteins till other chaperones or degradation system can act, and Hsp104 promotes solubilization of aggregated proteins [38]. Mutations in αA- or αB-crystallin and other small heat shock proteins have been linked to increased risk of certain amyloidosis conditions [40]. Hsp70 and Hsp90 are capable of utilizing ATP molecules to perform energy costly unfolding of stable misfolded aggregates and then convert them into properly folded conformations. Similarly other members like Hsp110 can act as disaggregases to forcibly solubilize preformed fibrillar aggregates [41]. Hsp70 recognizes KFERQ motif in proteins to mediate degradation by transporting them to lysosomal compartments inside the cell, a typical example of chaperone mediated autophagy [41]. Disaggregation machinery in humans mainly involves proteins from Hsp family namely, Hsp110, Hsp105, Hsp100, and Hsp70/40 co-chaperone cognates [41]. Proteins in Hsp40 co-chaperone family, mainly DnaJB6 and B8, can decrease the aggregation of polyglutamine peptides in HD in addition to modulating Hsp70 activity [41]. Binding of misfolded proteins and interaction with ATPase domain of Hsp70 are achieved via the highly conserved J domain. Anti-aggregation activity of Hsps, especially Hsp70, is at least partially dependent on presence of ATP [42]. ADP-bound forms of Hsp70 is generated as a result of the ATP hydrolysis by its co-chaperones. This form has high affinity for the hydrophobic residues in misfolded peptides and holds the unfolded ensemble till it spontaneously achieved proper conformational populations. Small Hsps (12–42 KDa) do not require ATP since they usually function as holdases and assist other proteins in Hsp chaperone complex [41]. Hsp104 is part of proteostasis network which regulates prion assembly in yeast, Hsp110 is its human counterpart which facilitates chaperone function of Hsp70 and Hsp40 [43]. Hsp90 and its co-chaperones interact with tau and a close control on their interaction can cause the oligomers to turn into benign species instead of toxic aggregates [44]. Free Hsp90 binds and releases substrates in its monomeric extended form, and attains a dimeric ATP-bound closed conformation mediate by co-chaperones p23/Sba1, Hsp70/90 organizing protein (HOP), Cdc37 and other proteins [41, 45]. Hsp60 chaperonins are mainly located in mitochondria and form heptameric, double ring complexes which provide isolated environment for protein folding with co-chaperonins Hsp10 forming the lid of this cavity. Unlike Hsp60, the cytosolic chaperonins in neurons such as TCP-1 Ring Complex (TRiC or CCT), do not require co-chaperones for their function and form a double ring complex with each ring consisting of eight subunits. Hsp60 can interact with mutant αS in PD brain [41]. B-chain of monomeric or dimeric insulin can bind to IAPP monomers, thus preventing their aggregation [7]. Presence of chaperone proteins such as Hsp27 (HSPB1), αB-crystallin (HSPB5) and Hsp70 (HSPA1A) can decrease cell toxicity, possibly by binding Aβ oligomers and converting them to larger less toxic aggregates, however it not clear whether they bring about any change in secondary or tertiary structure [44]. HSPA6 is induced in neurons post heat shock, unlike other members of this family HSPA1A and HSP8, thus has a unique probable role as a human neuronal chaperone [46]. Hsp70 and Hsp90 have differential roles in case of inflammation and macrophage recruitment. Hsp90 can also stabilize neurotoxic proteins and should be carefully controlled to achieve required therapeutic outcomes.

3.2 Intracellular degradation machinery

Molecular chaperones are capable of binding and folding intracellular soluble amyloid aggregates, however solubilization of amyloid plaques is out of scope for these proteins. While these chaperones may not be able to completely reverse neurodegenerative symptoms, they play crucial anti-apoptotic functions through protein folding and degradation of unfolded or misfolded proteins. There are fundamental mechanisms underlying a chaperone function, namely, unfolded protein response (UPR), protein compartmentalization, heat shock response, chaperone-mediated autophagy and lysosome system, ER associated degradation (ERAD) and ubiquitin-proteasome system (UPS) [46]. Extracellular chaperones mediate proper protein folding and refolding by providing isolated environment or through intermolecular interactions. If proper folding cannot be achieved, they may play a role in mediating intracellular proteasomal degradation or microglial digestion of unfolded protein fragments. Cytosolic chaperone systems like Hsp70 and Hsp40, comprising heat shock cognate 70 (Hsc70), and their interacting proteins like C-terminus of Hsc70-interacting protein (CHIP) can mediate targeting of misfolded proteins to proteasome machinery [38]. Intracellular pathways for degradation of misfolded proteins include the UPS, 26S proteasome and lysosome-mediated phagocytosis [9]. Hsp70 recruits misfolded proteins, especially αS, for degradation by directing them to proteasomes or autophagy-lysosomal pathway [44]. UPS comprises of a cascade of enzymes E1 (Ub activating enzyme), E2 (Ub conjugating enzyme) and E3 (Ub ligase) facilitating binding of ubiquitin (Ub) to target proteins. Ligase enzymes in this cascade can employ chaperone proteins such as Hsc70 for recognition of exposed hydrophobic regions misfolded peptides and together mediate proteolysis of misfolded intracellular proteins. E4 enzyme from UPS can recruit additional Ub molecules to the protein substrate which can be either directed for degradation or other protein interactions based on their topology [41]. 26S proteasome particles degrades polyUb-conjugated proteins with the help of its two 19S regulatory particles responsible for recognition and de-ubiquitination and a 20S core particle through which proteolysis of these substrates into short peptides is mediated [41]. ERAD withholds misfolded proteins from continuing with on-going cargo to golgi apparatus for further downstream processes. UPR is initiated by migration of transcription factors to the nucleus causing upregulation of ER chaperones-encoding gene expression in response to increasing in unfolded proteins, such as Aβ aggregates in neuronal ER, which require assistance of chaperones present in the ER [47, 48]. Reduction in UPS proteolytic activity is linked with disease pathology in AD, PD, ALS, HD as well as TSEs.

3.3 Non-proteolytic amyloid chaperones

Aging is a major risk factor for many neurodegenerative diseases. The UPS system is known to get deregulated with increase in neuronal age and therefore offers less resistance to pathological protein aggregates. Proteins such as Human serum albumin show competitive binding towards Aβ oligomers, with its different domains binding to many oligomeric molecules [12]. It binds almost 90% of plasma Aβ peptides potentially through dual binding mechanism involving Aβ(1-40) monomers and protofibrils likely utilizing residues involved in fibril formation.

Here we suggest non-proteolytic endogenous proteins which show potential as AD therapeutics directed against Aβ, namely, Lipocalin-type prostaglandin D synthase (L-PGDS, also known as β-trace), apolipoprotein E (ApoE), α2 macroglobulin (α2M), haptoglobin and clusterin as alternatives to intracellular Aβ degradation machinery. These proteins have been sporadically studied for their holdase or transporter activity, yet their Aβ chaperone function is still unexplored. L-PGDS is the second most abundant protein in human CSF after albumin [49]. It exhibits dual functions, as a lipophilic ligand transporter in cells and as isomerase in arachidonic acid pathway to convert prostaglandin H2 to prostaglandin D2 [50]. L-PGDS plays protective roles in different neurological diseases including genetic demyelinating disease, brain injuries and multiple sclerosis [51]. It plays protective role against cerebral ischemia as well [52]. It promotes recruitment of astrocytes and glial cells to the source of injury [53]. Early stress stimulus can upregulate L-PGDS suggesting its unknown, yet novel stress protection mechanism. L-PGDS deficient mice showed additional neuronal apoptosis strongly indicating its important protective function in neurons and surrounding oligodendrocytes [51]. L-PGDS binds to various Aβ peptides including fibrils and is colocalized in Aβ plaques [54]. Furthermore, Aβ is physiologically secreted to CSF under normal conditions where L-PGDS is abundant. L-PGDS ratio in CSF:serum is already identified as early biomarker for detection of potential damage to blood–brain barrier and quantifying its complex with transthyretin in CSF is also suggested to be a possible diagnostic marker [55]. L-PGDS colocalizes with amyloid plaques and mediates inhibition of aggregates through cysteine residue [54]. Based on previous studies and our own findings, we posit that secreted L-PGDS binds to monomeric and prefibrillar Aβ and inhibits amyloid aggregation in synergy with its ability to break down mature fibrils.

α2M is also an extracellular glycoprotein showing potential chaperone properties. It can prevent proteases from hydrolyzing proteins and mediates Aβ clearance through formation of α2M/protease complexes [56]. It protects cells from apoptosis through receptor interactions. α2M selectively binds proteins in non-native conformations and in process prevents their aggregation [56]. α2M decreases with age and its clearance mechanisms for Aβ get impaired [57]. A2M gene polymorphisms are associated with sporadic AD in some populations and α2M is also present in amyloid plaques [58]. ApoE isoforms E2, E3 and E4 can delay amyloid aggregation through differential activity as extracellular chaperones [59]. The suggested role of apoE4 as pathological chaperone is partly due to its role in delaying fibril formation from Aβ monomers leading to increase in toxic oligomeric species [60]. Isoform E3 and E2 show binding affinity to oligomeric Aβ as well, thus decreasing their neurotoxic effects [61]. Clusterin, or apolipoprotein J, is also extracellularly secreted and moonlights as a chaperone upon cellular stress stimulus [62]. Very similar to sHsps in its chaperone activity, clusterin preferentially interacts with off-pathway aggregates which are highly toxic and prone to precipitation [63]. Haptoglobin is very similar to α2M as an extracellular chaperone glycoprotein, though is activated in acidic environment and interacts with prefibrillar amyloids to prevent further aggregation [64]. These endogenous proteins and many others may play very important yet undiscovered role to maintain proteostasis in physiological environment.

Ig antibodies such as aducanumab, targeting specific oligomeric forms of Aβ have already reached clinical trials, showing promising therapeutic effects focusing on delaying cognitive decline. However recent withdrawals of some of these antibodies like bapineuzumab and solanezumab, have forced scientists to look for more robust options [65]. Enzymes such as insulin degrading enzymes, neprilysin, cathepsin B are capable of degrading amyloidogenic peptides. Enzymes belonging to peptidyl-prolyl cis/trans isomerase (PPIase) family, e.g., CypB, can convert proline containing peptides from cis to trans conformational isomers thereby facilitating their folding process [47]. Some inherently amyloidogenic proteins such as PrP, hTTR and BRICHOS-domain containing proteins can paradoxically also inhibit aggregation of other amyloids. BRICHOS-domain containing protein ITM2B (Bri2) is a CNS membrane protein which shows chaperone functions for amyloids by binding to tyrosine and other charged residues [43]. Tetrameric form of Transthyretin (TTR) can also prevent aggregation of amyloid oligomers [44]. In recent years, some small molecules have also been put forward as possible inhibitors of amyloid aggregation—such as polyphenols, e.g., epigallocatechin gallate (EGCG), curcumin, resveratrol, etc. [7, 12]. Understanding inhibition mechanisms of non-proteolytic endogenous chaperone proteins and other molecules is very crucial to develop therapies targeting amyloid aggregates in AD (Table 2).

ChaperoneFunctionAmyloid speciesChromosomal and cellular locationBrain expression levels (HPA) [66]Refs.
Hsp70 (DnaK)HoldaseNascent amyloids and oligomers6p21; intracellular66.3 TPM[42, 67, 68]
Hsp90HoldaseNascent amyloids and oligomers14q32; intracellular1021.5 TPM[69, 70]
Hsp40 (DnaJ)Holdase; disaggregase in ternary complexHigher order aggregates19p13; intracellular110.5 TPM[71]
αB crystallin (Hsp27)HoldaseAmyloid fibrils11q23; intracellular, membrane1888.7 TPM[72]
Hsp110Disaggregase; holdasePrefibrillar oligomers13q12; intracellular109.4 TPM[73]
Hsp28 (sHsp)HoldaseUnfolded or misfolded proteins7q11; intracellular98.7 TPM[74]
Hsc70 (HSPA8)Chaperone mediate autophagyNascent polypeptides11q24; intracellular1119.0 TPM[75]
Hsp60ChaperoninEarly oligomeric species2q33; intracellular141.6 TPM[76, 77]
STUB1 (CHIP)Ubiquitin-mediated protein degradationMisfolded protein aggregated16p13; intracellular137.7 TPM[78]
L-PGDS (β-trace)Aggregation inhibitorMonomers, fibrils9q34; secreted1224.7 TPM[54]
Clusterin (ApoJ)StabilizerSenile plaques/diffusible aggregates8p21; secreted10875.4 TPM[79]
α2-MacroglobulinInhibitorPrefibrillar species12p13; extracellular172.0 TPM[80]

Table 2.

Mode of function of endogenous amyloid chaperones.

Advertisement

4. Oxidative stress

Oxidative stress in human body is mainly mediated by reactive oxygen species (ROS) including hydrogen peroxide (H2O2), hydroxyl free radical (•OH), superoxide (O2•−), peroxyl (•OOR), hypochlorous acid (HOCl), nitric oxide (NO), peroxynitrite (ONOO) and other reactive nitrogen species. Increase in concentration of these reactive species can trigger downregulation of antioxidant enzymes such as superoxide dismutase (SOD), glutathione peroxidase, catalase, thioredoxin and small molecules such as melatonin and coumarin [81]. Endogenous sources of ROS include components in mitochondrial electron transport chain and NADPH oxidase, which can trigger pathological responses leading to lipid peroxidation, DNA damage and cell death mediated by toxic products. Amyloid plaques have been shown to contain multi-fold concentrations of metals like iron (Fe), copper (Cu) and zinc (Zn) [81]. Intracellular Aβ can also trigger production of highly toxic molecules such as 4-hydroxyl-2,3-nonenal (HNE) and malandialdehyde through interactions with Fe or Cu which can catalyze Fenton reactions to generate more OH ions. Hydroxyl radicals can also be generated from other brain ROS through Haber-Weiss reaction, particularly within mitochondria which are most prone to such oxidative damage [82]. Such ROS are generated as by-products of amyloid aggregation and are involved with metallobiology of dementia. High concentrations of Fe in brain have drastic effects in cases of dementia leading to cognitive decline through interactions with brain amyloids [83]. Metal ions are crucial for many brain functions and Fe and Cu even possess active sites for binding antioxidant SOD. However, increase in concentrations of some of these metals can cause high levels of toxicity. Twenty percent of cases in familial ALS is affected by mutation in a single enzyme SOD1 functionally affected by Cu/Zn [83]. Metal ions have propensity to bind high affinity, well-protected and redox-shielded binding sites of proteins, and in higher concentrations they can bind to other putative active sites in proteins involved in pathology of various degenerative diseases [84]. While zinc compounds have been tested for their probable therapeutic role in overcoming cognitive degeneration, Zn2+ ion has also been implicated as a contributor to formation of amyloid plaques which act as metal sinks [84]. Such mitochondrial damage and mutations have been associated with age related late-onset, non-autosomal dominant AD pathology [85].

4.1 Amyloid-β and oxidative stress

Aβ binds and reduces Fe3+ and Cu2+ in presence of endogenous reducing agents to generate H2O2 further producing other partially ROS [86, 87]. Studies have revealed the role of metal ions such as Fe, Cu and Zn in inducing Aβ aggregation and oligomerization [88, 89, 90]. Brain regions rich in Aβ(1–42) show increased oxidative stress, possibly mediated through redox interactions with the only methionine (Met35) present in peptide sequence [91]. Amyloids oligomers can also trigger ROS generation [2]. The three histidine residues-His6, His13 and His14, facilitate Aβ coordination with transition metal ions. These residues get protonated in acidic environment and may increasingly contribute to aggregation at low pH. Cu2+ interacts with Aβ and oxidizes sulfur of Met35 to activate formation of disulfide bonds leading to dimerization and other oligomeric formations. Soluble oxidized aggregates avoid clearance causing enrichment of brain regions with these agglomerates. Aβ peptides can successfully recruit ions of metals like Fe, Cu and Zn through sulfide group of Met35 and a chelating domain which involves Asp1 and all three histidines, in a bid to induce redox complexes capable of bringing about oxidative insults [7]. Dityrosine cross-linked Aβ dimers along with nitrotyrosine cross-linked proteins are also associated with oxidative stress [92]. Cell death from oxidative stress is a cumulative result of alteration in proteostasis, protein phosphorylation and glucose metabolism as downstream consequences of increase in Aβ42 concentration [91].

4.2 Pathophysiology of ROS

Excessive free radical production as a result of oxidative stress at cellular levels causes protein oxidation and lipid peroxidation [91]. Lipid peroxidation leads to break down of unsaturated fatty acids among other components of membrane phospholipids, leading to accumulation of isoprostanes, acrolein, thiobarbiturate-reactive substances, etc. [84, 92]. Glutamate receptors overstimulation can trigger downstream cell death cascades through increased calcium influx and generation of nitric oxide species [48]. 8-hydroxy-2′-deoxyguanosine (OHDG) is one of the oxidative markers for DNA, found in PD patients [36]. These products impair glucose transport and glutamate uptake, hence contributing to cell apoptosis. ROS cause imbalance in metal and ion homeostasis, for example Ca2+, which can trigger imbalance in downstream signaling cascades. Oxidative damage can lead to hydroxylation of nucleic acids and carbonylation of proteins. Carbonyls are markers of protein oxidation and have been found concentrated in frontal brain regions of AD patients [92]. Free radicals generated as result of amyloid oligomerization or aggregation can directly mediate mitochondrial damage which triggers neuronal death through downstream pathways, one of them being cytochrome C reduction [48]. OS in PD cases may be a result of deregulation of dopamine-iron redox pathway, since αS can alter expression of enzymes indirectly regulating dopamine synthesis [84]. αS is also known to directly interact with metal ions causing protein aggregation. ALS is mainly characterized by loss of motor neurons, which combined with SOD mutations diminishing its free radical scavenging abilities can exacerbate the impacts to oxidative injury [84]. Oxidative markers localized in plaques and NFTs are toxic products such as 3-nitrotyrosine, HNE, pyrraline and pentosidine, while metal enriched protein carbonyls including ferritin, catalyst, Cu/Zn-SOD and Mn-SOD [93].

4.3 Protective mechanisms from ROS

Chaperones can bind ROS generated as by-product of amyloids and thus prevent triggering breakdown of homeostasis. α2-Macroglobulin can directly bind Aβ and potentially act as a chaperone in addition to its zinc-binding capabilities which can help mitigate redox activity of Aβ [94]. Zn2+ is redox-inert and may be helpful in mitigating metal mediated Aβ redox activity. ApoE can mediate Aβ clearance as a chaperone depending on specific isoform interactions; ε4 may potentially increase Aβ pathology [95]. ApoE interaction with Aβ is modulated by metals [93]. There are two key requirements for metal interactions in dementia-either supplementation or chelation [83]. Metal chelators can help solubilization of Aβ plaques. Effects of chaperones on metal ions need to be carefully curated to maintain metal homeostasis in brain and other organs of nervous system. Several endogenous brain proteins show potential protective response against ROS in diseased brain, including β-trace protein, which is the second most abundant protein in CSF. Competitive binding of metal ions by chaperone proteins may be advantageous in decreasing generation of ROS by-products. Overexpression of metal-affinity proteins have been shown, as example of iron-binding ferritin. Antioxidant alpha-tocopherol has been reported to potentially slow AD progression in addition to action of metal chelators such as clioquinol and desferrioxamine [92]. L-PGDS scavenges ROS and in process protects against neuronal cell death with its ligand binding function intact [96]. It also reduces cytotoxicity mediated by oxidation of heme metabolites such as bilirubin [97]. We have found that L-PGDS can directly interact with Aβ-heme complex and lower its peroxidase activity (to be published). Other non-proteolytic chaperones such as α2M, clusterin and haptoglobin, also show neuroprotective potential in similar manner as L-PGDS and are good candidates for more comprehensive oxidative stress related studies.

Advertisement

5. Conclusions

Alzheimer’s disease is a debilitating neurodegenerative condition and is projected to be a major risk factor for global population by the year 2050 [98]. One of the key theories, known as “amyloid cascade hypothesis” postulates that oligomerization of amyloid-β (Aβ) in brain is the key pathological event in AD [99]. Although complete explanation of causation in AD is yet to be established, it has been universally accepted that amyloidosis (perhaps provoked by environmental factors) plays a crucial role in AD progression [99, 100, 101]. Conformational alterations in Aβ which lead to its conversion from soluble peptide to insoluble aggregates are considered as a key mechanism in pathogenesis of neurodegenerative diseases such as AD [102]. Early amyloid aggregates can act as biomarkers in most dementia related maladies and associate observed clinical symptoms to underlying pathophysiological mechanisms. Till date, we do not have any therapeutic solution for aggregation of amyloids. Current gold-standard biomarkers in neurodegeneration are neuroimaging systems of degeneration and detectable clinical symptoms represent pathological changes causing irreversible damage to nervous system [33, 36]. Success of theranostic efforts will rely on rational drug design based on a proper understanding of molecular structures and mechanisms involved in aggregation. Aβ mediates AD pathology through direct inhibition of neuronal interactions and signaling cascades, and triggers oxidative stress in the process. Here we have discussed mechanisms of action of amyloid chaperones with focus on Aβ chaperones to better understand their modus operandi. Apart from the protein quality control machinery and housekeeping complexes responsible for maintaining general proteostasis, we have discussed importance of other endogenous chaperones which step up in time of crisis, often as response to stress stimuli. Proteins such as L-PGDS, clusterin, α2M have are involved in many physiological processes, with elusive function as amyloid chaperones. These proteins can provide alternative mechanisms to control amyloid aggregates in events of failure of intracellular Hsp complexes and proteasome machinery. Additionally, their affinity to redox active components may provide additional protective mechanisms against Aβ mediated oxidative stress.

Advertisement

Conflict of interest

Authors declare no conflict of interest.

Advertisement

Abbreviations

Aaamino acid
ADAlzheimer’s disease
ALSamyotrophic lateral sclerosis
APPamyloid precursor protein
amyloid-β
CNScentral nervous system
CSFcerebrospinal fluid
DARRdipolar assisted rotation resonance
EMelectron microscopy
HDHuntington disease
HNEhydroxy-2,3-nonenal
Hscheat shock cognate
Hspheat shock protein
HSRheat shock response
IAPPislet amyloid polypeptide
NMDAN-methyl-D-aspartate
NMRnuclear magnetic resonance
OHDG8-hydroxy-2′-deoxyguanosine
PDParkinson’s disease
PDBprotein data bank
PrPprion protein
REDORrotational echo double resonance
ROSreactive oxygen species
SODsuperoxide dismutase
ssNMRsolid state nuclear magnetic resonance
T2Dtype II diabetes
TEMtransmission electron microscopy
TSEtransmissible spongiform encephalopathy
Ububiquitin
UPRunfolded protein response
UPSubiquitin proteasome system
αSα synuclein

References

  1. 1. Herguedas B, Garcia-Nafria J, Cais O, Fernandez-Leiro R, Krieger J, Ho H, et al. Structure and organization of heteromeric AMPA-type glutamate receptors. Science. 2016;352(6285):aad3873
  2. 2. Iadanza MG, Jackson MP, Hewitt EW, Ranson NA, Radford SE. A new era for understanding amyloid structures and disease. Nature reviews Molecular Cell Biology. 2018;19(12):755-773
  3. 3. Dubois B, Hampel H, Feldman HH, Scheltens P, Aisen P, Andrieu S, et al. Preclinical Alzheimer’s disease: Definition, natural history, and diagnostic criteria. Alzheimer’s & Dementia. 2016;12(3):292-323
  4. 4. Benjamin Dehay MB, Gorry P, Przedborski S, Vila M, Hunot S, Singleton A, et al. Targeting α-synuclein for treatment of Parkinson’s disease: Mechanistic and therapeutic considerations. Lancet Neurology. 2015;14:855-866
  5. 5. Wei L, Jiang P, Xu W, Li H, Zhang H, Yan L, et al. The molecular basis of distinct aggregation pathways of islet amyloid polypeptide. The Journal of Biological Chemistry. 2011;286(8):6291-6300
  6. 6. Chen GF, Xu TH, Yan Y, Zhou YR, Jiang Y, Melcher K, et al. Amyloid beta: Structure, biology and structure-based therapeutic development. 2017;38(9):1205-1235
  7. 7. Ke PC, Sani MA, Ding F, Kakinen A, Javed I, Separovic F, et al. Implications of peptide assemblies in amyloid diseases. Chemical Society Reviews. 2017;46(21):6492-6531
  8. 8. Xiao Y, Ma B, McElheny D, Parthasarathy S, Long F, Hoshi M, et al. Abeta(1-42) fibril structure illuminates self-recognition and replication of amyloid in Alzheimer’s disease. Nature Structural & Molecular Biology. 2015;22(6):499-505
  9. 9. Muchowski PJ, Wacker JL. Modulation of neurodegeneration by molecular chaperones. Nature Reviews. Neuroscience. 2005;6(1):11-22
  10. 10. Chuang E, Hori AM, Hesketh CD, Shorter J. Amyloid assembly and disassembly. Journal of Cell Science. 2018;131(8):jcs189928
  11. 11. Anant K, Paravastu RDL, Yau W-M, Tycko R. Molecular structural basis for polymorphism in Alzheimer’s β-amyloid fibrils. Proceedings of the National Academy of Sciences. 2008;105(47):18349-18354
  12. 12. Ahmed R, Melacini G. A solution NMR toolset to probe the molecular mechanisms of amyloid inhibitors. Chemical Communications (Camb). 2018;54(37):4644-4652
  13. 13. Rajamohamedsait HB, Sigurdsson EM. Histological staining of amyloid and pre-amyloid peptides and proteins in mouse tissue. Methods in Molecular Biology. 2012;849:411-424
  14. 14. Xue C, Lin TY, Chang D, Guo Z. Thioflavin T as an amyloid dye: Fibril quantification, optimal concentration and effect on aggregation. Royal Society Open Science. 2017;4(1):160696
  15. 15. Chen B, Thurber KR, Shewmaker F, Wickner RB, Tycko R. Measurement of amyloid fibril mass-per-length by tilted-beam transmission electron microscopy. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(34):14339-14344
  16. 16. Wälti MA, Ravotti F, Arai H, Glabe CG, Wall JS, Böckmann A, et al. Atomic-resolution structure of a disease-relevant Aβ(1-42) amyloid fibril. Proceedings of the National Academy of Sciences of the United States of America. 2016;113(34):E4976-E4984
  17. 17. Lührs T, Ritter C, Adrian M, Riek-Loher D, Bohrmann B, Döbeli H, et al. 3D structure of Alzheimer’s amyloid-β(1-42) fibrils. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(48):17342-17347
  18. 18. Gremer LSD, Schenk C, Reinartz E, Labahn J, Ravelli RBG, Tusche M, et al. Fibril structure of amyloid-β(1-42) by cryo-electron microscopy. Science. 2017;358(6359):116-119
  19. 19. Petkova AT, Ishii Y, Balbach JJ, Antzutkin ON, Leapman RD, Delaglio F, et al. A structural model for Alzheimer’s beta-amyloid fibrils based on experimental constraints from solid state NMR. Proceedings of the National Academy of Sciences of the United States of America. 2002;99(26):16742-16747
  20. 20. Roche J, Shen Y, Lee JH, Ying J, Bax A. Monomeric Abeta(1-40) and Abeta(1-42) peptides in solution adopt very similar Ramachandran map distributions that closely resemble random coil. Biochemistry. 2016;55(5):762-775
  21. 21. Laganowsky A, Liu C, Sawaya MR, Whitelegge JP, Zhao M, Pensalfini A, et al. Atomic view of a toxic amyloid small oligomer. Science. 2012;335:1228-1231
  22. 22. Kayed R, Pensalfini A, Margol L, Sokolov Y, Sarsoza F, Head E, et al. Annular protofibrils are a structurally and functionally distinct type of amyloid oligomer. The Journal of Biological Chemistry. 2009;284(7):4230-4237
  23. 23. Mainz A, Peschek J, Stavropoulou M, Back KC, Bardiaux B, Asami S, et al. The chaperone alphaB-crystallin uses different interfaces to capture an amorphous and an amyloid client. Nature Structural & Molecular Biology. 2015;22(11):898-905
  24. 24. Kotler SA, Ramamoorthy A. Preparation of stable amyloid-beta oligomers without perturbative methods. Methods in molecular biology. 1777;2018:331-338
  25. 25. Rambaran RN, Serpell LC. Amyloid fibrils: Abnormal protein assembly. Prion. 2008;2(3):112-117
  26. 26. Han S, Kollmer M, Markx D, Claus S, Walther P, Fandrich M. Amyloid plaque structure and cell surface interactions of beta-amyloid fibrils revealed by electron tomography. Scientific Reports. 2017;7:43577
  27. 27. Lakshmanan A, Cheong DW, Accardo A, Di Fabrizio E, Riekel C, Hauser CA. Aliphatic peptides show similar self-assembly to amyloid core sequences, challenging the importance of aromatic interactions in amyloidosis. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(2):519-524
  28. 28. Hardy J, Selkoe DJ. The amyloid hypothesis of Alzheimer’s disease: Progress and problems on the road to therapeutics. Science. 2002;297(5580):353-356
  29. 29. Cohen SIA, Linse S, Luheshi LM, Hellstrand E, White DA, Rajah L, et al. Proliferation of amyloid-β42 aggregates occurs through a secondary nucleation mechanism. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(24):9758-9763
  30. 30. Kumar EK, Haque N, Prabhu NP. Kinetics of protein fibril formation: Methods and mechanisms. International Journal of Biological Macromolecules. 2016;100:3-10
  31. 31. Wegmann S, Eftekharzadeh B, Tepper K, Zoltowska KM, Bennett RE, Dujardin S, et al. Tau protein liquid-liquid phase separation can initiate tau aggregation. EMBO Journal. 2018;37(7):e98049
  32. 32. 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(7636):217-221
  33. 33. Holtzman DM, Morris JC, Goate AM. Alzheimer’s disease: The challenge of the second century. Science Translational Medicine. 2011;3(77):77sr1
  34. 34. Walsh DM, Selkoe DJ. A beta oligomers—A decade of discovery. Journal of Neurochemistry. 2007;101(5):1172-1184
  35. 35. Perez-Garmendia R, Gevorkian G. Pyroglutamate-modified amyloid Beta peptides: Emerging targets for Alzheimer’s disease immunotherapy. Current Neuropharmacology. 2013;11:491-498
  36. 36. Sherer TB. Biomarkers for Parkinson’s disease. Science Translational Medicine. 2011;3(79):79ps14
  37. 37. Lazarev VF, Mikhaylova ER, Guzhova IV, Margulis BA. Possible function of molecular chaperones in diseases caused by propagating amyloid aggregates. Frontiers in Neuroscience. 2017;11:277
  38. 38. Uversky VN. Flexible nets of malleable guardians: Intrinsically disordered chaperones in neurodegenerative diseases. Chemical Reviews. 2011;111(2):1134-1166
  39. 39. Bohrmann B, Tjernberg L, Kuner P, Poli S, Levet-Trafit B, Näslund J, et al. Endogenous proteins controlling amyloid b-peptide polymerization. The Journal of Biological Chemistry. 1999;274(23):15990-15995
  40. 40. Hishiya A, Takayama S. Molecular chaperones as regulators of cell death. Oncogene. 2008;27(50):6489-6506
  41. 41. Ciechanover A, Kwon YT. Protein quality control by molecular chaperones in neurodegeneration. Frontiers in Neuroscience. 2017;11:185
  42. 42. Evans CG, Wisen S, Gestwicki JE. Heat shock proteins 70 and 90 inhibit early stages of amyloid beta-(1-42) aggregation in vitro. The Journal of Biological Chemistry. 2006;281(44):33182-33191
  43. 43. Landreh M, Rising A, Presto J, Jornvall H, Johansson J. Specific chaperones and regulatory domains in control of amyloid formation. The Journal of Biological Chemistry. 2015;290(44):26430-26436
  44. 44. Lindberg I, Shorter J, Wiseman RL, Chiti F, Dickey CA, McLean PJ. Chaperones in neurodegeneration. The Journal of Neuroscience. 2015;35(41):13853-13859
  45. 45. Hutt D, Balch WE. The proteome in balance. Science. 2010;329(5993):766-767
  46. 46. Roodveldt C, Outeiro TF, Braun JEA. Editorial: Molecular chaperones and neurodegeneration. Frontiers in Neuroscience. 2017;11:565
  47. 47. Dubnikov T, Cohen E. The emerging roles of early protein folding events in the secretory pathway in the development of neurodegenerative maladies. Frontiers in Neuroscience. 2017;11:48
  48. 48. Shen Y, He P, Zhong Z, McAllister C, Lindholm K. Distinct destructive signal pathways of neuronal death in Alzheimer’s disease. Trends in Molecular Medicine. 2006;12(12):574-579
  49. 49. Mase M, Yamada K, Shimazu N, Seiki K, Oda H, Nakau H, et al. Lipocalin-type prostaglandin D synthase (beta-trace) in cerebrospinal fluid: A useful marker for the diagnosis of normal pressure hydrocephalus. Neuroscience Research. 2003;47(4):455-459
  50. 50. Yoshihiro Urade OH. Biochemical, structural, genetic, physiological, and pathophysiological features of lipocalin-type prostaglandin D synthase. Biochimica et Biophysica Acta. 2000;1482(1-2):259-271
  51. 51. Taniguchi H, Mohri I, Okabe-Arahori H, Kanekiyo T, Kagitani-Shimono K, Wada K, et al. Early induction of neuronal lipocalin-type prostaglandin D synthase after hypoxic-ischemic injury in developing brains. Neuroscience Letters. 2007;420(1):39-44
  52. 52. Saleem S, Shah ZA, Urade Y, Doré S. Lipocalin-prostaglandin D synthase is a critical beneficial factor in transient and permanent focal cerebral ischemia. Neuroscience. 2009;160(1):248-254
  53. 53. Lee S, Jang E, Kim JH, Kim JH, Lee WH, Suk K. Lipocalin-type prostaglandin D2 synthase protein regulates glial cell migration and morphology through myristoylated alanine-rich C-kinase substrate: Prostaglandin D2-independent effects. The Journal of Biological Chemistry. 2012;287(12):9414-9428
  54. 54. Kanekiyo T, Ban T, Aritake K, Huang ZL, Qu WM, Okazaki I, et al. Lipocalin-type prostaglandin D synthase/beta-trace is a major amyloid beta-chaperone in human cerebrospinal fluid. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(15):6412-6417
  55. 55. Lovell MA, Lynn BC, Xiong S, Quinn JF, Kaye J, Markesbery WR. An aberrant protein complex in CSF as a biomarker of Alzheimer disease. Neurology 2008;70(23):2212-2218
  56. 56. French K, Yerbury JJ, Wilson MR. Protease activation of R2-macroglobulin modulates a chaperone-like action with broad specificity. Biochemistry. 2008;47:1176-1185
  57. 57. Birkenmeier G, Müller R, Huse K, Forberg J, Gläser C, Hedrich H, et al. Human alpha2-macroglobulin: Genotype-phenotype relation. Experimental Neurology. 2003;184:153-161
  58. 58. Michałowska-Wender G, Wawrzynek A, Rossa G, Kozubski W, Wender M. On the lack of a clear-cut association between alpha-2-macroglobulin deletion and the risk of Alzheimer disease in Poland. Folia Neuropathologica. 2014;52(4):417-420
  59. 59. Fagan AM, Watson M, Parsadanian M, Bales KR, Paul SM, Holtzman DM. Human and murine ApoE markedly alters A beta metabolism before and after plaque formation in a mouse model of Alzheimer’s disease. Neurobiology of Disease. 2002;9(3):305-318
  60. 60. Deroo S, Stengel F, Mohammadi A, Henry N, Hubin E, Krammer EM, et al. Chemical cross-linking/mass spectrometry maps the amyloid beta peptide binding region on both apolipoprotein E domains. ACS Chemical Biology. 2015;10(4):1010-1016
  61. 61. Ly S, Altman R, Petrlova J, Lin Y, Hilt S, Huser T, et al. Binding of apolipoprotein E inhibits the oligomer growth of amyloid-beta peptide in solution as determined by fluorescence cross-correlation spectroscopy. The Journal of Biological Chemistry. 2013;288(17):11628-11635
  62. 62. Humphreys DT, Carver JA, Easterbrook-Smith SB, Wilson MR. Clusterin has chaperone-like activity similar to that of small heat shock proteins. The Journal of Biological Chemistry. 1999;274(11):6875-6881
  63. 63. Poon S, Treweek TM, Wilson MR, Easterbrook-Smith SB, Carver JA. Clusterin is an extracellular chaperone that specifically interacts with slowly aggregating proteins on their off-folding pathway. FEBS Letters. 2002;513:259-266
  64. 64. Yerbury JJ, Kumita JR, Meehan S, Dobson CM, Wilson MR. Alpha2-macroglobulin and haptoglobin suppress amyloid formation by interacting with prefibrillar protein species. The Journal of Biological Chemistry. 2009;284(7):4246-4254
  65. 65. Behl C. Amyloid in Alzheimer’s disease: Guilty beyond reasonable doubt? Trends in Pharmacological Sciences. 2017;38(10):849-851
  66. 66. Uhlen M, Fagerberg L, Hallstrom BM, Lindskog C, Oksvold P, Mardinoglu A, et al. Proteomics. Tissue-based map of the human proteome. Science. 2015;347(6220):1260419
  67. 67. Fernandez-Funez P, Sanchez-Garcia J, de Mena L, Zhang Y, Levites Y, Khare S, et al. Holdase activity of secreted Hsp70 masks amyloid-β42 neurotoxicity in Drosophila. Proceedings of the National Academy of Sciences of the United States of America. 2016;113(35):E5212-E5221
  68. 68. Cummings CJ, Sun Y, Opal P, Antalffy N, Mestril R, Orr HT, et al. Over-expression of inducible HSP70 chaperone suppresses neuropathology and improves motor function in SCA1 mice. Human Molecular Genetics. 2001;10(14):1511-1518
  69. 69. Martinez-Ruiz A, Villanueva L, Gonzalez de Orduna C, Lopez-Ferrer D, Higueras MA, Tarin C, et al. S-nitrosylation of Hsp90 promotes the inhibition of its ATPase and endothelial nitric oxide synthase regulatory activities. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(24):8525-8530
  70. 70. Orenstein SJ, Cuervo AM. Chaperone-mediated autophagy: Molecular mechanisms and physiological relevance. Seminars in Cell & Developmental Biology. 2010;21(7):719-726
  71. 71. Chan HYE, Warrick JM, Gray-Board GL, Paulson HL, Bonini NM. Mechanisms of chaperone suppression of polyglutamine disease: Selectivity, synergy and modulation o protein solubility in Drosophila. Human Molecular Genetics. 2000;9(19):2811-2820
  72. 72. Raman B, Ban T, Sakai M, Pasta SY, Ramakrishna T, Naiki H, et al. AlphaB-crystallin, a small heat-shock protein, prevents the amyloid fibril growth of an amyloid beta-peptide and beta2-microglobulin. The Biochemical Journal. 2005;392(Pt 3):573-581
  73. 73. Nillegoda NB, Kirstein J, Szlachcic A, Berynskyy M, Stank A, Stengel F, et al. Crucial HSP70 co-chaperone complex unlocks metazoan protein disaggregation. Nature. 2015;524(7564):247-251
  74. 74. Haruo Shinohara YI, Goto S, Inagaki T, Kato K. Alpha B crystallin and HSP28 are enhanced in the cerebral cortex of patients with Alzheimer’s disease. Journal of the neurological sciences. 1993;119(04122):203-208
  75. 75. Gao X, Carroni M, Nussbaum-Krammer C, Mogk A, Nillegoda NB, Szlachcic A, et al. Human Hsp70 Disaggregase reverses Parkinson’s-linked alpha-synuclein amyloid fibrils. Molecular Cell. 2015;59(5):781-793
  76. 76. Mangione MR, Vilasi S, Marino C, Librizzi F, Canale C, Spigolon D, et al. Hsp60, amateur chaperone in amyloid-beta fibrillogenesis. Biochimica et Biophysica Acta. 2016;1860(11 Pt A):2474-2483
  77. 77. Walls KC, Coskun P, Gallegos-Perez JL, Zadourian N, Freude K, Rasool S, et al. Swedish Alzheimer mutation induces mitochondrial dysfunction mediated by HSP60 mislocalization of amyloid precursor protein (APP) and beta-amyloid. The Journal of Biological Chemistry. 2012;287(36):30317-30327
  78. 78. Urushitani M, Kurisu J, Tateno M, Hatakeyama S, Nakayama K, Kato S, et al. CHIP promotes proteasomal degradation of familial ALS-linked mutant SOD1 by ubiquitinating Hsp/Hsc70. Journal of Neurochemistry. 2004;90(1):231-244
  79. 79. Narayan P, Orte A, Clarke RW, Bolognesi B, Hook S, Ganzinger KA, et al. The extracellular chaperone clusterin sequesters oligomeric forms of the amyloid-beta(1-40) peptide. Nature Structural & Molecular Biology. 2011;19(1):79-83
  80. 80. Hughes SR, Khorkova O, Goyal S, Knaeblein J, Heroux J, Riedel NG, et al. a2-macroglobulin associates with b-amyloid peptide and prevents fibril formation. Proceedings of the National Academy of Sciences of the United States of America. 1998;95:3275-3280
  81. 81. Poprac P, Jomova K, Simunkova M, Kollar V, Rhodes CJ, Valko M. Targeting free radicals in oxidative stress-related human diseases. Trends in Pharmacological Sciences. 2017;38(7):592-607
  82. 82. Mecocci P, Boccardi V, Cecchetti R, Bastiani P, Scamosci M, Ruggiero C, et al. A long journey into aging, brain aging, and Alzheimer’s disease following the oxidative stress tracks. Journal of Alzheimer’s Disease. 2018;62(3):1319-1335
  83. 83. Adlard PA, Bush AI. Metals and Alzheimer’s disease: How far have we come in the clinic? Journal of Alzheimer's Disease. 2018;62(3):1369-1379
  84. 84. Barnham KJ, Masters CL, Bush AI. Neurodegenerative diseases and oxidative stress. Nature Reviews Drug Discovery. 2004;3(3):205-214
  85. 85. Carmichael G. Mitochondrial mutations may underlie late-onset AD. Lancet Neurology. 2004;3:513
  86. 86. Pramanik D, Ghosh C, Dey SG. Heme-Cu bound abeta peptides: Spectroscopic characterization, reactivity, and relevance to Alzheimer’s disease. Journal of the American Chemical Society. 2011;133(39):15545-15552
  87. 87. Pramanik D, Ghosh C, Mukherjee S, Dey SG. Interaction of amyloid β peptides with redox active heme cofactor: Relevance to Alzheimer’s disease. Coordination Chemistry Reviews. 2013;257(1):81-92
  88. 88. Atwood CS, Moir RD, Huang X, Scarpa RC, Bacarra NM, Romano DM, et al. Dramatic aggregation of Alzheimer Aβ by Cu(II) is induced by conditions representing physiological acidosis. Journal of Biological Chemistry. 1998;273(21):12817-12826
  89. 89. Xudong Huang CSA, Hartshorn MA, Multhaup G, Goldstein LE, Scarpa RC, Cuajungco MP, et al. The Aβ peptide of alzheimer’s disease directly produces hydrogen peroxide through metal ion reduction. Biochemistry. 1997;38(24):7609-7616
  90. 90. Curtain CC, Ali F, Volitakis I, Cherny RA, Norton RS, Beyreuther K, et al. Alzheimer's disease amyloid-beta binds copper and zinc to generate an allosterically ordered membrane-penetrating structure containing superoxide dismutase-like subunits. The Journal of Biological Chemistry. 2001;276(23):20466-20473
  91. 91. Butterfield DA, Boyd-Kimball D. Oxidative stress, amyloid-beta peptide, and altered key molecular pathways in the pathogenesis and progression of Alzheimer’s disease. Journal of Alzheimer's Disease. 2018;62(3):1345-1367
  92. 92. Smith DG, Cappai R, Barnham KJ. The redox chemistry of the Alzheimer’s disease amyloid beta peptide. Biochimica et Biophysica Acta. 2007;1768(8):1976-1990
  93. 93. Xudong Huang RDM, Tanzi RE, Bush AI, Rogers JT. Redox-active metals, oxidative stress, and Alzheimer’s disease pathology. Annals of the New York Academy of Sciences. 2004;1012:153-164
  94. 94. Wyatt AR, Constantinescu P, Ecroyd H, Dobson CM, Wilson MR, Kumita JR, et al. Protease-activated alpha-2-macroglobulin can inhibit amyloid formation via two distinct mechanisms. FEBS Letters. 2013;587(5):398-403
  95. 95. Cedazo-Minguez A. Apolipoprotein E and Alzheimer’s disease: Molecular mechanisms and therapeutic opportunities. Journal of Cellular and Molecular Medicine. 2007;11(6):1227-1238
  96. 96. Fukuhara A, Yamada M, Fujimori K, Miyamoto Y, Kusumoto T, Nakajima H, et al. Lipocalin-type prostaglandin D synthase protects against oxidative stress-induced neuronal cell death. The Biochemical Journal. 2012;443(1):75-84
  97. 97. Inui T, Mase M, Shirota R, Nagashima M, Okada T, Urade Y. Lipocalin-type prostaglandin D synthase scavenges biliverdin in the cerebrospinal fluid of patients with aneurysmal subarachnoid hemorrhage. Journal of Cerebral Blood Flow and Metabolism. 2014;34(9):1558-1567
  98. 98. Hebert LE, Weuve J, Scherr PA, Evans DA. Alzheimer disease in the United States (2010-2050) estimated using the 2010 census. Neurology. 2013;80(19):1778-1783
  99. 99. Hardy JA, Higgins GA. Alzheimer’s disease: The amyloid cascade hypothesis. Science. 1992;256(5054):184-185
  100. 100. Skaper SD. Alzheimer’s disease and amyloid: Culprit or coincidence? International Review of Neurobiology. 2012;102:277-316
  101. 101. Gong Y, Chang L, Viola KL, Lacor PN, Lambert MP, Finch CE, et al. Alzheimer’s disease-affected brain: Presence of oligomeric A beta ligands (ADDLs) suggests a molecular basis for reversible memory loss. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(18):10417-10422
  102. 102. Ru-Jing Ren EBD, Wang G, Seyfried NT, Levey AI. Proteomics of protein post-translational modifications implicated in neurodegeneration. Translational Neurodegeneration. 2014;3(1):23

Written By

Bhargy Sharma and Konstantin Pervushin

Submitted: 14 December 2018 Reviewed: 08 January 2019 Published: 01 February 2019