Next Article in Journal
Chiral Dualism as an Instrument of Hierarchical Structure Formation in Molecular Biology
Next Article in Special Issue
Spontaneous Chiral Symmetry Breaking and Entropy Production in a Closed System
Previous Article in Journal
Stability and Dynamics of Viscoelastic Moving Rayleigh Beams with an Asymmetrical Distribution of Material Parameters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chiral Interface of Amyloid Beta (Aβ): Relevance to Protein Aging, Aggregation and Neurodegeneration

1
Departmemts: Virtual Reality Perception Lab. (VV. Dyakin) and Center for Neurochemistry (A. Lajtha), The Nathan S. Kline Institute for Psychiatric Research (NKI), Orangeburg, NY 10962, USA
2
Departments of Neurology, Pathology and Psychiatry, Center for Cognitive Neurology, New York University School of Medicine, New York, NY 10016, USA
*
Author to whom correspondence should be addressed.
Symmetry 2020, 12(4), 585; https://doi.org/10.3390/sym12040585
Submission received: 17 March 2020 / Revised: 31 March 2020 / Accepted: 2 April 2020 / Published: 7 April 2020
(This article belongs to the Special Issue Asymmetry in Biological Homochirality)

Abstract

:
Biochirality is the subject of distinct branches of science, including biophysics, biochemistry, the stereochemistry of protein folding, neuroscience, brain functional laterality and bioinformatics. At the protein level, biochirality is closely associated with various post-translational modifications (PTMs) accompanied by the non-equilibrium phase transitions (PhTs NE). PTMs NE support the dynamic balance of the prevalent chirality of enzymes and their substrates. The stereoselective nature of most biochemical reactions is evident in the enzymatic (Enz) and spontaneous (Sp) PTMs (PTMs Enz and PTMs Sp) of proteins. Protein chirality, which embraces biophysics and biochemistry, is a subject of this review. In this broad field, we focus attention to the amyloid-beta (Aβ) peptide, known for its essential cellular functions and associations with neuropathology. The widely discussed amyloid cascade hypothesis (ACH) of Alzheimer’s disease (AD) states that disease pathogenesis is initiated by the oligomerization and subsequent aggregation of the Aβ peptide into plaques. The racemization-induced aggregation of protein and RNA have been extensively studied in the search for the contribution of spontaneous stochastic stereo-specific mechanisms that are common for both kinds of biomolecules. The failure of numerous Aβ drug-targeting therapies requires the reconsolidation of the ACH with the concept of PTMs Sp. The progress in methods of chiral discrimination can help overcome previous limitations in the understanding of AD pathogenesis. The primary target of attention becomes the network of stereospecific PTMs that affect the aggregation of many pathogenic agents, including Aβ. Extensive recent experimental results describe the truncated, isomerized and racemized forms of Aβ and the interplay between enzymatic and PTMs Sp. Currently, accumulated data suggest that non-enzymatic PTMs Sp occur in parallel to an existing metabolic network of enzymatic pathways, meaning that the presence and activity of enzymes does not prevent non-enzymatic reactions from occurring. PTMs Sp impact the functions of many proteins and peptides, including Aβ. This is in logical agreement with the silently accepted racemization hypothesis of protein aggregation (RHPA). Therefore, the ACH of AD should be complemented by the concept of PTMs Sp and RHPA.

1. Introduction

The amyloid cascade hypothesis (ACH) has played a crucial role in the understanding of the Alzheimer’s disease (AD) etiology and pathogenesis. The deposition of β-amyloid (Aβ) and neurofibrillary tangles (NFTs) traditionally served as the essential neuropathological features of AD. However, for many years, the attention to the stereochemistry of underlying spontaneous events was under-appreciated. Stereochemical errors in biomolecular structures, including proteins and peptides, have a dramatic impact on cell physiology [1]. The discovery of free D-aspartic acid (D-Asp) in rodents and man open a new window for understanding the mechanisms of protein synthesis and degradation [2]. Proteins, including glycoproteins, are the subjects of the reversible enzymatic (Enz) [3] and irreversible spontaneous (Sp) [4] post-translational modifications (PTM Enz and PTMs Sp). (see Figure 1). All physiological and pathological forms of proteins are the consequence of PTMs. We are focusing on the aberrant forms of PTM, such as racemization Sp and isomerization Sp. The relevance of the spontaneous modifications of amino acids (AAs) within peptides and long-lived proteins to protein aging, accumulation and pathologies is being increasingly recognized in the recent studies. Accordingly, the primarily biomarkers of aging and neurodegeneration are becoming the protein-cell-specific PTMs Sp of amino acids (AAs) [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50].

2. Racemization of the Aβ

With the recognition of the fact that many proteins (Aβ, TAU, prion protein Prion (PrP), Huntingtin and α-synuclein) are the substrate of the aggregation-prone PTMs Sp [51], we are focusing, primarily, on the racemization of the Aβ. The amyloid precursor protein (APP) is one of the most studied proteins concerning pathological misfolding. The products of APP processing by α-, β- and γ-secretases range from 16 to 49 AAs.
Most studied amyloid beta (Aβ) peptides includes Aβ (1–16), Aβ (1–19), Aβ (20–34), Aβ (20–33), Aβ (20–40), Aβ (23–34), Aβ (34–40), Aβ (35–40), Aβ (1–35), Aβ (1–40) and Aβ (1–42) [5,52] are characterized by differential chain of PTMs and susceptibility to PTMs Sp. PTMs of Aβ (1–42), the primary constituent of Aβ plaques in the AD brain, are extensively studied. Misfolding and aggregation of Aβ peptides is the convincing example of a link between the perturbations of the molecular chirality, deteriorated enzyme-substrate recognition, abnormal cell signaling (including neurotransmission) and cognitive dysfunction [3]. The spontaneous aggregation of Aβ peptides into amyloid plaques and in the walls of the cerebral vasculature is the unresolved issue of Alzheimer’s disease (AD)-amyloid conundrum [6]. It is a common assumption that PTMs Sp can significantly alter the structure of the original polypeptide chain. The AAs that most frequently undergo racemization Sp and isomerization Sp in human proteins are aspartate (Asp), asparagine (Asn), glutamate (Glu), glutamine (Glu), serine (Ser), alanine (Ala) and proline [7]. For Aβ peptides, racemization-prone are found two non-essential AAs: serine (Ser) and aspartate (Asp) (Aβ-42 contains two Ser and three Asp residues (see Table 1)). For Asp, the mechanism acceleration of racemization Sp (about 10 5) is associated with the specific succinimide intermediates [8,9,10]. Both D-Ser and D-Asp play a crucial role in N-methyl-d-aspartate (NMDA) receptor-mediated neurotransmission. D-Ser26-Aβ1–40 possesses a strong tendency to form fibrils [11]. AD patients have increased brain D-Ser levels [12]. This fact agrees with the activated spontaneous racemization (RsSp) of Ser residue in Aβ, with an elevated level of D-Ser in amyloid plaques, impairment of the NMDA neurotransmission, memory loss and cognitive dysfunction. Racemization and isomerization of Asp are the most common types of non-enzymatic covalent modification that leads to an accumulation of aging proteins in numerous human tissues [13]. Asp-1, Asp-7 and Asp-23 of Aβ are crucial in the control of Aβ aging and aggregation [5].
Residues Asp-1 and Asp-7 of Aβ in amyloid plaques are a mixture of L-, D-, L-iso- and D-iso-aspartate [14]. D-Asp-7 enhances the aggregation process by shifting the equilibrium of Aβ from the soluble to the insoluble form [15]. Therefore, the set of PTMs Sp, including racemization Sp and isomerization Sp, is an efficient modifier of Aβ metabolism. In 2011, Kumar promoted the hypothesis that enzymatic phosphorylation of Aβ triggers the formation of toxic aggregates [16], which has been confirmed by later studies [3]. In 1994, Szendrei discovered that spontaneous isomerization of Asp affects the conformations of synthetic peptides [17]. However, the role of the PTMs Sp in Aβ aggregation and neurotoxicity remains in the shadow. Consequently, many structural details of misfolded Aβ have remained elusive for a long time [7,18]. This short review provides a summary of information regarding events of PTMs Sp in Aβ. The heterogeneity of Aβ proteolytic forms in AD brain is represented by at least 26 unique peptides, characterized by various N- and C-terminal truncations. The N- and C-terminal truncated fragments (in contrast to canonical Aβ) are allowing to distinguish between the soluble and insoluble aggregates. The N-terminal truncations are predominating in the insoluble material and C- terminal truncations segregating in the soluble aggregates [19,20]. Aβ peptides exhibit a high sensitivity of the secondary structure and fibril morphologies to the chirality of ligands [21] and enzymes of PTMs. Only for a small part of Aβ isoforms exist information regarding the pathways of PTMs Sp.
Currently, available data for Aβ42 peptides are summarized in Table 1, Table 2 and Table 3. The data in Table 2 and Table 3 demonstrate two essential facts: first, the coincidence of phosphorylation and spontaneous racemization/isomerization (enzymatic phosphorylation can be accompanied by the enzyme-driven or spontaneous racemization) events at the Ser-8, Ser-26 (Ser-26 residue is located within the turn region of Aβ), and Asp23 residues, second, currently available information covering the PTMs Sp of Aβ is limited only to 4 from the 16 types of AAs, which means that much remains to be explored. Most recent attempts to overcome previous limitations of the ACH are concentrated on many additional essential aspects of metabolism, contributing to progress in understanding [22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42]. However, most them, do not pay enough attention to the stereochemistry of the PTM in general and the impact of spontaneous racemization (Sp) on Aβ assembly, aggregation and functions. At the same time, the progress in methods of chiral discrimination has produced new, stereochemistry-oriented, experimental results regarding aberrant PTMs Sp. Growing evidence suggests that proteins undergo several unusual, previously unknown PTMs associates with the interplay between physiological protein modification, spontaneous aging-associated molecular processes [7,10,11,13], stress conditions [23,24], accidental co-localization of the enzyme and substrate or PTMs Sp. For Aβ, Asp and Ser are known as the most racemization prone residues. For the illustration purpose, we provide several of many existing molecular pathways where the racemization Sp of Ser can be critical.
First, the pathological role of mitochondrial enzymes Ser proteases (SerPs) is attributed to neurodegenerative disorders such as AD and Parkinson’s and disease [25,26,27]. The HtrA (high-temperature requirement) family represents a class of oligomeric SerPs [25]. Its members are classified by presence (in its AAs sequence) a catalytic triad contains His, Asp and Ser residues known as racemization prone.
Second, chaperone signaling complexes in AD involve a wide range of heat shock proteins (Hsp), including Hsp27, that are ingaged in protection against Aβ aggregation and toxicity [28]. Human Hsp27 is phosphorylated at three Ser residues (Ser15, Ser78 and Ser82), were Ser-78 and Ser-82 are the major phosphorylation sites [29,30]. It is evident that due to the stereo-specificity activity of both protein types (SerPs and Hsps) racemization Sp of Ser residues in each of them will contribute to the aberrant processing of substrates, including Aβ, inducing the cascade of aggregation, accompanied by neurodegeneration. The aggregation of protein and peptide indicates the decrease in the turnover rate. Accordingly, the previously short half-life-proteins are changing in the direction toward the long-lived one. Lowering turnover rate (i.e., protein aging) makes proteins the subject of the time-dependent PTMs Sp, including oxidation, nitration, glycation, isomerization and racemization [7]. The set of PTMs Sp and its effect on protein polymerization both are substrate specific. Tyrosine (Tyr) nitration, for example, significantly decreased the aggregation of Aβ1–40. [43].

3. Conclusions

In the manuscript, we assess the previous and current experimental results acquired in the specific areas of chiral proteomics—Aβ folding—from a broad perspective. For this purpose, we addressed the basic, fundamental and widely recognized facts and theories underlying the stereochemistry of Aβ. Due to progress in multidisciplinary fields, the view of the origin of biologic non-equilibrium chirality evolves from the physico-chemical nature of enantioselective autocatalytic reaction networks to a process that play an essential role in the pathogenesis of AD [44]. The phenomena of biochirality embrace two undivided branches of science biophysics and biochemistry. In 1990th, the nature of living organisms was associated with the absolute homochirality [50]. With the discovery of D-AAs in living organisms and the process of enzymic racemization, the concept of homochirality was replaced by the notion of prevalent chirality.
In the language of entropy, the transfer of protein/solvent system from the state of low-entropy (racemic mixture) to the high-entropy state (homochirality) is the order-disorder type transitions.
In terms of thermodynamics, this is the transition from the non-equilibrium to the equilibrium state. Accordingly, the enzymic PTMs, from a biophysical perspective, is the set of physiological non-equilibrium phase transitions. Enzymic racemization is an essential and necessary source of D-AAs in organisms. In contrast, the spontaneous racemization, as an aberrant PTMs, is the window for the irreversible transfer from non-equilibrium to equilibrium conditions.
In the words of proteomics, irreversible racemization is the conformation of protein from functional (physiological) to the dis-functional (inert or toxic) state of protein solvent, aggregates and depositions. The universal significance of the symmetry constraints is evident from the viral to the human proteome. The biologic significance of racemization-induced protein aggregation for the neuropathogenesis of AD was experimentally demonstrated as early as 1994 [45]. Currently accumulated data about PTMs of many proteins and peptides, including Aβ are coherent with the silently accepted racemization hypothesis of protein aggregation (RHPA). Therefore, after “three decades of struggles, ACH [46] of the neurodegeneration should be complemented by the concept of PTMs Sp and non-equilibrium phase transitions (PhTs NE) [47,48,49,50].

Author Contributions

V.V.D. contribute to review of biophysical aspect of molecular chirality, PTMs, protein aggregation, and neurodegeneration. A.L. contribute to review of biochemical aspect of molecular chirality, PTMs, protein aggregation, and neurodegeneration. T.M.W. contributed to review of biochemical aspects of PTMs, protein aggregation and neurodegeneration. All authors have read and agreed to the published version of the manuscript.

Funding

This manuscript is supported by NIH grants AG060882, AF058267 and AG066512 to TMW.

Acknowledgments

We gratefully thanks: Justin Lucas for criticism and correction. Csaba Vadasz for criticism and correction. Alexander G. Dadali for discussion of the thermodynamics of protein conformation, and Frank Fagnano for contribution to clarity of the argumentation.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

ACHAmyloid cascade hypothesis
PTMspost-translational modifications
PTMs Enzenzymic PTMs
PTMs Spspontaneous PTMs
PhTs NEnon-equilibrium phase transitions
amyloid beta
RHPAracemization hypothesis of protein aggregation
AlaAlanine
AsnAsparagine
SerSerine
Aspaspartic acid
GluGlutamate
IleIsoleucine
TyrTyrosine
ProProline

References

  1. Schreiner, E.; Trabuco, L.G.; Freddolino, P.L.; Schulten, K. Stereochemical errors and their implications for molecular dynamics simulations. BMC Bioinform. 2011, 12, 190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Dunlop, D.S.; Neidle, A.; McHale, D.; Dunlop, D.M.; Lajtha, A. The presence of free D-aspartic acid in rodents and man. Biochem. Biophys. Res. Commun. 1986, 141, 27–32. [Google Scholar] [CrossRef]
  3. Jamasbi, E.; Separovic, F.; Hossain, M.A.; Ciccotosto, G.D. Phosphorylation of a full-length amyloid-β peptide modulates its amyloid aggregation, cell binding and neurotoxic properties. Mol. BioSyst. 2017, 13, 1545–1551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Yokoyama, H.; Mizutani, R.; Noguchi, S.; Hayashida, N. Structural and biochemical basis of the formation of isoaspartate in the complementarity-determining region of antibody 64M-5 Fab. Sci. Rep. 2019, 9, 18494. [Google Scholar] [CrossRef] [Green Version]
  5. Moro, M.L.; Collins, M.J.; Cappellini, E. Alzheimer’s disease and amyloid β-peptide deposition in the brain: A matter of ‘aging’? Biochem. Soc. Trans. 2010, 38, 539–544. [Google Scholar] [CrossRef] [Green Version]
  6. Roher, A.E.; Kokjohn, T.A.; Clarke, S.G.; Sierks, M.R.; Maarouf, C.L.; Serrano, G.E.; Sabbagh, M.S.; Beach, T.G. APP/Aβ structural diversity and Alzheimer’s disease pathogenesis. Neurochem. Int. 2018, 110, 1–13. [Google Scholar] [CrossRef]
  7. McCudden, C.R.; Kraus, V.B. Biochemistry of amino acid racemization and clinical application to musculoskeletal disease. Clin. Biochem. 2006, 39, 1112–1130. [Google Scholar] [CrossRef]
  8. Radkiewicz, J.L.; Zipse, H.; Clarke, S.; Houk, K.N. Accelerated racemization of aspartic acid and asparagine residues via succinimide intermediates:  An ab initio theoretical exploration of mechanism. J. Am. Chem. Soc. 1996, 118, 9148–9155. [Google Scholar] [CrossRef]
  9. Helfman, P.M.; Bada, J.L.; Shou, M.Y. Considerations on the role of aspartic acid racemization in the aging process. Gerontology 1977, 23, 419–425. [Google Scholar] [CrossRef]
  10. Takahashi, O.; Kirikoshi, R.; Manabe, N. Academic editor Mihai V. Putz. Racemization of the succinimide intermediate formed in proteins and peptides: A computational study of the mechanism catalyzed by dihydrogen phosphate ion. Int. J. Mol. Sci. 2016, 10, 1698. [Google Scholar] [CrossRef] [Green Version]
  11. Kubo, T.; Kumagae, Y.; Miller, C.A.; Kaneko, I. Beta-amyloid racemized at the Ser26 residue in the brains of patients with Alzheimer’s disease: Implications in the pathogenesis of Alzheimer’s disease. J. Neuropathol. Exp. Neurol. 2003, 62, 248–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Madeira, C.; Lourenco, M.V.; Vargas-Lopes, C.; Suemoto, C.K.; Brandão, C.O.; Reis, T.; Leite, R.E.P.; Laks, J.; Jacob-Filho, W.; Pasqualucci, C.A. D-serine levels in Alzheimer’s disease: Implications for novel biomarker development. Transl. Psychiatry 2015, 5, e561. [Google Scholar] [CrossRef] [PubMed]
  13. Ritz-Timme, S.; Collins, M.J. Racemization of aspartic acid in human proteins. Ageing Res. Rev. 2002, 1, 43–59. [Google Scholar] [CrossRef]
  14. Roher, A.E.; Yablenson, J.D.; Clarke, S.; Wolkow, C.; Wang, R.; Cotter, R.J.; Reardon, I.M.; Ziircher-Neely, H.A.; Heinrikson, R.L.; Ball, M.J.; et al. Structural alterations in the peptide backbone of β-amyloid core protein may account for its deposition and stability in Alzheimer’s disease. J. Biol. Chem. 1993, 268, 3072–3083. [Google Scholar]
  15. Kuo, Y.M.; Webster, S.; Emmerling, M.R.; de Lima, N.; Roher, A.E. Irreversible dimerization/tetramerization and post-translational modifications inhibit proteolytic degradation of Aβ peptides of Alzheimer’s disease. Biochim. Biophys. Acta. 1998, 1406, 291–298. [Google Scholar] [CrossRef] [Green Version]
  16. Kumar, S.; Walter, J. Phosphorylation of amyloid beta (Aβ) peptides—A trigger for formation of toxic aggregates in Alzheimer’s disease. Aging 2011, 3, 803–812. [Google Scholar] [CrossRef]
  17. Szendrei, G.I.; Fabian, H.; Mantsch, H.H.; Lovas, S.; Nyéki, O.; Schön, I.; Otvos, L. Aspartate-bond isomerization affects the major conformations of synthetic peptides. Eur. J. Biochem. FEBS 1994, 226, 917–924. [Google Scholar] [CrossRef] [Green Version]
  18. Xiao, Y.; Ma, B.; McElheny, D.; Parthasarathy, S.; Long, F.; Hoshi, M.; Nussinov, R.; Ishii, Y. Aβ (1–42) fibril structure illuminates self-recognition and replication of amyloid in Alzheimer’s disease. Nat. Struct. Mol. Biol. 2015, 22, 499–505. [Google Scholar] [CrossRef] [Green Version]
  19. Wildburger, N.C.; Esparza, T.J.; LeDuc, R.D.; Fellers, R.T.; Thomas, P.M.; Cairns, N.J.; Kelleher, N.L.; Bateman, R.J.; David, L.; Brody, D.L. Diversity of amyloid-beta proteoforms in the Alzheimer’s disease brain. Sci. Rep. 2017, 7, 9520. [Google Scholar] [CrossRef] [Green Version]
  20. Jiang, N.; Leithold, L.H.E.; Post, J.; Ziehm, T.; Mauler, J.; Gremer, L.; Cremer, M.; Schartmann, E.; Shah, N.J.; Kutzsche, J.; et al. Preclinical pharmacokinetic studies of the tritium labelled D-enantiomeric peptide D3 developed for the treatment of Alzheimer’s disease. PLoS ONE 2015, 10, e0128553. [Google Scholar]
  21. Malishev, R.; Arad, E.; Bhunia, S.K.; Shaham-Niv, S.; Kolusheva, S.; Gazit, E.; Jelinek, R. Chiral modulation of amyloid beta fibrillation and cytotoxicity by enantiomeric carbon dots. Chem. Commun. 2018, 54, 7762–7765. [Google Scholar] [CrossRef] [PubMed]
  22. Zhou, Y.; Sun, Y.; Ma, Q.H.; Liu, Y. Alzheimer’s disease: Amyloid-based pathogenesis and potential therapies. Cell Stress. 2018, 2, 150–161. [Google Scholar] [CrossRef] [PubMed]
  23. Ravikirana, B.; Mahalakshmi, R. Unusual post-translational protein modifications: The benefits of sophistication. RSC Adv. 2014, 4, 33958–33974. [Google Scholar] [CrossRef]
  24. Osna, N.A.; Carter, W.G.; Ganesan, M.; Kirpich, I.A.; McClain, C.J.; Petersen, D.R.; Shearn, C.T.; Tomasi, M.L.; Kharbanda, K.K. Aberrant post-translational protein modifications in the pathogenesis of alcohol-induced liver injury. World J. Gastroenterol. 2016, 22, 6192–6200. [Google Scholar] [CrossRef] [PubMed]
  25. Grau, S.; Baldi, A.; Bussani, R.; Tian, X.; Stefanescu, R.; Przybylski, M.; Richards, P.; Jones, S.A.; Shridhar, V.; Tim Clausen, T.; et al. Implications of the serine protease HtrA1 in amyloid precursor protein processing. Proc. Natl. Acad. Sci. USA 2005, 102, 6021–6026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Gieldon, A.; Zurawa-Janicka, D.; Jarzab, M.; Wenta, T.; Golik, P.; Dubin, G.; Lipinska, B.; Ciarkowski, J. Distinct 3D architecture and dynamics of the human Htra2(Omi) protease and its mutated variants. PLoS ONE 2016, 11, e0161526. [Google Scholar] [CrossRef] [PubMed]
  27. Goo, H.G.; Rhim, H.; Kang, S. Pathogenic role of serine protease HtrA2/Omi in neurodegenerative diseases. Curr. Protein Pept. Sci. 2017, 18, 746–757. [Google Scholar] [CrossRef]
  28. Singh, M.K.; Sharma, B.; Tiwari, P.C. The small heat shock protein Hsp27: Present understanding and future prospects. J. Therm. Biol. 2017, 69, 149–154. [Google Scholar] [CrossRef]
  29. Landry, J.; Lambert, H.; Zhou, M.; Lavoie, J.N.; Hickey, E.; Weber, L.A.; Anderson, C.W. Human HSP27 is phosphorylated at serines 78 and 82 by heat shock and mitogen-activated kinases that recognize the same amino acid motif as S6 kinase II. J. Biol. Chem. 1992, 267, 794–803. [Google Scholar]
  30. Katsogiannou, M.; Andrieu, C.; Rocchi, P. Heat shock protein 27 phosphorylation state is associated with cancer progression. Front. Genet. 2014. [Google Scholar] [CrossRef] [Green Version]
  31. Lowenson, J.D.; Clarke, S.; Roher, A.E. Chemical modifications of deposited amyloid-β peptides. Methods Enzym. 1999, 309, 89–105. [Google Scholar]
  32. Takahashi, O.; Kirikoshi, R.; Manabe, N. Racemization of serine residues catalyzed by dihydrogen phosphate Ion: A computational Study. Catalysts 2017, 7, 363. [Google Scholar] [CrossRef] [Green Version]
  33. Shapira, R.; Austin, G.E.; Mirra, S.S. Neuritic plaque amyloid in Alzheimer’s disease is highly racemized. J. Neurochem. 1988, 50, 69–74. [Google Scholar] [CrossRef] [PubMed]
  34. Kaneko, I.; Morimoto, K.; Kubo, T. Drastic neuronal loss in vivo by beta-amyloid racemized at Ser (26) residue: Conversion of non-toxic [D-Ser (26)] beta-amyloid 1–40 to toxic and proteinase-resistant fragments. Neuroscience 2001, 104, 1003–1011. [Google Scholar] [CrossRef]
  35. Tomiyama, T.; Asano, S.; Furiya, Y.; Shirasawa, T.; Endo, N.; Mori, H. Racemization of Asp23 residue affects the aggregation properties of Alzheimer amyloid beta protein analogues. J. Biol. Chem. 1994, 269, 10205–10208. [Google Scholar]
  36. Shimizu, T.; Fukuda, H.; Murayama, S.; Izumiyama, N.; Shirasawa, T. Iso-aspartate formation at position 23 of amyloid beta peptide enhanced fibril formation and deposited onto senile plaques and vascular amyloids in Alzheimer’s disease. J. Neurosci. Res. 2002, 70, 451–461. [Google Scholar] [CrossRef]
  37. Warmack, R.A.; Boyer, D.R.; Zee, C.T.; Richards, L.R.; Sawaya, M.R.; Cascio, D.; Gonen, T.; Eisenberg, D.S.; Clarke, S.G. Structure of A-β (20–34) with Alzheimer’s-associated isomerization at Asp23 reveals a distinct protofilament interface. Nat. Commun. 2019, 10, 3357. [Google Scholar] [CrossRef] [Green Version]
  38. Kumar, S.; Singh, S.; Hinze, D.; Josten, M.; Sahl, H.G.; Siepmann, M.; Walter, J. Phosphorylation of amyloid-β peptide at serine-8 attenuates its clearance via insulin-degrading and angiotensin-converting enzymes. J. Biol. Chem. 2012, 287, 8641–8651. [Google Scholar] [CrossRef] [Green Version]
  39. Barykin, E.P.; Petrushanko, I.Y.; Kozin, S.A.; Telegi, G.B.; Cherno, A.S.; Lopina, O.D.; Radko, S.P.; Mitkevich, V.A.; Makarov, A.M. Phosphorylation of the amyloid-beta peptide inhibits zinc-dependent aggregation, prevents Na,K-ATPase inhibition, and reduces cerebral plaque deposition. Front. Mol. Neurosci. 2018, 11, 302. [Google Scholar] [CrossRef]
  40. Rezaei-Ghaleh, N.; Amininasab, M.; Kumar, S.; Walter, J.; Zweckstetterc, M. Phosphorylation modifies the molecular stability of β-amyloid deposits. Nat. Commun. 2016, 7, 11359. [Google Scholar] [CrossRef] [Green Version]
  41. Milton, N.G. Phosphorylation of amyloid-beta at the serine 26 residue by human cdc2 kinase. Neuroreport 2001, 12, 3839–3844. [Google Scholar] [CrossRef] [PubMed]
  42. Kumar, S.; Wirths, O.; Stüber, K.; Wunderlich, P.; Koch, P.; Theil, S.; Rezaei-Ghaleh, N.; Zweckstetter, M.; Bayer, T.A.; Brüstle, O.; et al. Phosphorylation of the amyloid β-peptide at Ser26 stabilizes oligomeric assembly and increases neurotoxicity. Acta Neuropathol. 2016, 131, 525–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zhao, J.; Wang, P.; Li, H.; Gao, Z. Nitration of Y10 in Aβ1–40: Is it a compensatory reaction against oxidative/nitrative stress and Aβ aggregation? Chem. Res. Toxicol. 2015, 28, 401–407. [Google Scholar] [CrossRef] [PubMed]
  44. Ribó, J.M.; Hochberg, D. Concept Paper. Chemical basis of biological homochirality during the abiotic evolution stages on Earth. Symmetry 2019, 11, 814. [Google Scholar] [CrossRef] [Green Version]
  45. Mori, H.; Ishii, K.; Tomiyama, T.; Furiya, Y.; Sahara, N.; Asano, S.; Endo, N.; Shirasawa, T.; Takio, K. Racemization: Its biological significance on neuropathogenesis of Alzheimer’s disease. Tohoku J. Exp. Med. 1994, 174, 251–262. [Google Scholar] [CrossRef] [Green Version]
  46. Kasim, J.K.; Kavianinia, I.; Harris, P.W.R.; Brimble, M.A. Three decades of amyloid beta synthesis: Challenges and advances. Front. Chem. 2019, 7, 472. [Google Scholar] [CrossRef]
  47. Dyakin, V.V.; Lucas, J. Non-equilibrium phase transition in biochemical—Systems. Chain of chirality transfer as Determinant of Brain Functional Laterality. Relevance to Alzheimer disease and cognitive psychology. In Proceedings of the Alzheimer’s Association International Conference (AAIC-2017), London, UK, 16–20 July 2017. [Google Scholar]
  48. Ornes, S. Core concept: How nonequilibrium thermodynamics speaks to the mystery of life. Proc. Natl. Acad. Sci. USA 2017, 114, 423–424. [Google Scholar] [CrossRef] [Green Version]
  49. Schrödinger, E. What is Life? The Physical Aspect of the Living Cell; Cambridge University Press: Cambridge, UK, 1994. [Google Scholar]
  50. Nansheng, Z. The role of homochirality in evolution. In Advances in BioChirality; Zucchi, C., Caglioti, L., Palyi, G., Eds.; Elsevier Science: Amsterdam, The Netherlands, 1999. [Google Scholar]
  51. Hsu, Y.H.; Chen, Y.-W.; Wu, M.-H.; Tu, L.H. Protein glycation by glyoxal promotes amyloid formation by Islet Amyloid polypeptide. Biophys. J. 2019, 116, 2304–2313. [Google Scholar] [CrossRef]
  52. Zhang, A.; Qi, W.; Good, T.A.; Fernandez, E.J. Structural differences between Aβ (1–40) intermediate oligomers and fibrils elucidated by proteolytic fragmentation and hydrogen/deuterium exchange. Biophys. J. 2009, 96, 1091–1104. [Google Scholar]
Figure 1. Spontaneous deamidation and isomerization of asparagine (Asn). Side-chain bonds of asparagine and aspartate are drawn as bold lines. Adopted from [4].
Figure 1. Spontaneous deamidation and isomerization of asparagine (Asn). Side-chain bonds of asparagine and aspartate are drawn as bold lines. Adopted from [4].
Symmetry 12 00585 g001
Table 1. The frequency (f) of the AAs appearance in Aβ (1–42).
Table 1. The frequency (f) of the AAs appearance in Aβ (1–42).
The Frequency (f) of the AAs Appearance in A-beta (1–42)
fAmino Acids
6GlyVal
4Ala
3AspGluPheHis Ile
2LysLeuSer
1MetAsn ArgGlnTyr
Table 2. Coincidence of enzymatic and spontaneous PTMs at Ser and Asp residues of Aβ in the neurodegenerative amyloid aggregates.
Table 2. Coincidence of enzymatic and spontaneous PTMs at Ser and Asp residues of Aβ in the neurodegenerative amyloid aggregates.
PeptideDiseaseResiduePTMs
Rcm.Ism.Ph.
SpontaneousEnzymic
A-β (40–42)ADSer-8[24, 35] [3, 40, 41, 42]
Ser-26[13, 35, 36] [43. 44]
Asp-23[35, 37, 38] [40]
A-β (20–34)Asp-23 [39]
Post-translational modification (PTMs): Spontaneous (Sp) and Enzymic (Enz), Racemization (Rcm). Isomerization (Ism). Phosphorylation (Ph).
Table 3. Coincidence of phosphorylation and racemization at Ser and Asp residues of Aβ (1–42).
Table 3. Coincidence of phosphorylation and racemization at Ser and Asp residues of Aβ (1–42).
A-Beta (1–42)
N-Terminal123456789101112131415161718192021222324252627282930313233343536373839404142C-Terminal
AspAlaGluPheArgHisAspSerGlyTyrGluValHisHisGlnLysLeuValPhePheAlaGluAspValGlySerAsnLysGlyAlaIleIleGlyLeuMetValGlyGlyValValIleAla
DAEFRHDSGYEVHHQKLVFFAEDVGSNKGAIIGLMVGGVVIA
* *
**
*** *** ***
PTM: Enzymic (***) and Spontaneous (Racemization *, Izomerization **).

Share and Cite

MDPI and ACS Style

Dyakin, V.V.; Wisniewski, T.M.; Lajtha, A. Chiral Interface of Amyloid Beta (Aβ): Relevance to Protein Aging, Aggregation and Neurodegeneration. Symmetry 2020, 12, 585. https://doi.org/10.3390/sym12040585

AMA Style

Dyakin VV, Wisniewski TM, Lajtha A. Chiral Interface of Amyloid Beta (Aβ): Relevance to Protein Aging, Aggregation and Neurodegeneration. Symmetry. 2020; 12(4):585. https://doi.org/10.3390/sym12040585

Chicago/Turabian Style

Dyakin, Victor V., Thomas M. Wisniewski, and Abel Lajtha. 2020. "Chiral Interface of Amyloid Beta (Aβ): Relevance to Protein Aging, Aggregation and Neurodegeneration" Symmetry 12, no. 4: 585. https://doi.org/10.3390/sym12040585

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop