Next Article in Journal
NOX1 Inhibition Attenuates Kidney Ischemia-Reperfusion Injury via Inhibition of ROS-Mediated ERK Signaling
Next Article in Special Issue
Molecular Mechanisms Underlying Muscle Wasting in Huntington’s Disease
Previous Article in Journal
Aberrant BMP2 Signaling in Patients Diagnosed with Osteoporosis
Previous Article in Special Issue
Mitochondrial Respiration Changes in R6/2 Huntington’s Disease Model Mice during Aging in a Brain Region Specific Manner
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Commentary

Implications of the Orb2 Amyloid Structure in Huntington’s Disease

1
Stowers Institute for Medical Research, Kansas City, MO 64110, USA
2
MRC Laboratory of Molecular Biology, Francis Crick Avenue, Cambridge CB2 0QH, UK
3
Department of Molecular and Integrative Physiology, University of Kansas Medical Center, Kansas City, KS 66160, USA
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(18), 6910; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21186910
Submission received: 24 August 2020 / Revised: 11 September 2020 / Accepted: 16 September 2020 / Published: 21 September 2020
(This article belongs to the Special Issue Molecular Basis and Molecular Targets in Huntington’s Disease)

Abstract

:
Huntington’s disease is a progressive, autosomal dominant, neurodegenerative disorder caused by an expanded CAG repeat in the huntingtin gene. As a result, the translated protein, huntingtin, contains an abnormally long polyglutamine stretch that makes it prone to misfold and aggregating. Aggregation of huntingtin is believed to be the cause of Huntington’s disease. However, understanding on how, and why, huntingtin aggregates are deleterious has been hampered by lack of enough relevant structural data. In this review, we discuss our recent findings on a glutamine-based functional amyloid isolated from Drosophila brain and how this information provides plausible structural insight on the structure of huntingtin deposits in the brain.

Polyglutamine (PolyQ)-related diseases are dominant, late-onset genetic disorders manifested by progressive neurodegeneration. A common feature of this group of diseases, which include Huntington’s disease (HD), dentatorubral-pallidoluysian atrophy, spinobulbar muscular atrophy, and six types of spinocerebellar ataxias, is the abnormal expansion of a CAG codon repeat, coding for a 10 to 35 long glutamine tract in the wild-type protein [1]. First documented by George Huntington in 1872, HD is one of the most common inherited neurodegenerative diseases, causing cognitive disruptions and chorea with no effective cure [2]. The connection between HD and the expansion of the glutamine tract in the huntingtin (HTT) gene, which codes for the multidomain and multifunctional HTT protein [3,4], was identified in the early 1990s [5]. In HD, intraneuronal deposits of HTT fragments that map onto the exon 1 (HTTex1) are found in cerebellum, striatum, and cortex [6]. This led to the use of HTTex1 to determine the consequences of HTT deposits in mice and neuronal cell lines [7,8,9,10]. HD toxicity is believed to stem from a gain-of-toxic-function of HTT aggregates [11], along with a loss of function through sequestration of HTT and other proteins into the aggregates [12,13]. However, there is an ongoing debate about the nature of the harmful proteinaceous species in the brain; prefibrillar oligomeric, generally α-helical, assemblies [14], or fibrillar amyloid assemblies [15].
In vitro, the polyQ tract encoded by HTTex1 drives the self-assembly to an amyloid fold [16,17]. The assembly kinetics, however, also depends on the polyQ-flanking regions [18,19,20,21,22,23]. The in vitro-assembled HTT amyloid is proposed to adopt an antiparallel β-sheet arrangement [24,25,26,27,28,29,30,31,32,33,34]. Yet, despite intense efforts, the atomic resolution 3D architecture of aggregated HTT, even assembled in the test tube, remains elusive. A recent study employed cryo-ET methods to analyze the architecture of HTTex1 amyloid-like filaments in the cellular context [35]. However, there is no atomic-level structural information of pathological polyQ aggregates from patients’ brain. Unexpectedly, the functional amyloid formed by Drosophila Orb2 protein, a member of the cytoplasmic polyadenylation element binding proteins, provides plausible structural insight on the endogenous polyQ-based amyloids. The aggregated state of Orb2 plays a causal role in memory stabilization [36,37,38,39,40,41]. Using cryo-EM, the structure of the biochemically active Orb2 aggregates extracted from adult Drosophila head have been recently solved [42]. The structure revealed that Orb2 aggregates are left-handed C3 helical amyloid filaments, defined by three molecules per layer that form, on average, 750 Å continuous in-register parallel β-sheets (Figure 1A). The filament structure is stabilized by a Q-based amyloid core (Figure 1B), while the rest of the protein, comprising the RNA-recognition motifs and protein interaction domain, extends from the Q-based amyloid core [42].
In spite of differences in β-sheets arrangement, β-parallel (in vivo-assembled Orb2) versus β-antiparallel (in vitro-assembled HTTex1), both proteins employ a similar arrangement of individual molecules. Of the 704 amino acids of Orb2, 31 amino acids (176-206) form an antiparallel hairpin-like structure with a hydrophilic core stabilized by 7 inter-digitated Q coming from opposing stands connected by a turn—four from one β1 strand (Q179, Q181, Q183, and Q185) and three from the opposing β2 strand (Q200, Q202, and Q204), separated by 14 residues (Q-x-Q-x-Q-x-Q-x14-Q-x-Q-x-Q motif). The interior of the Orb2 amyloid is formed by tightly packed and hydrogen-bonded Q sidechains, whereas its exterior residues contribute to the protofilament interfaces [42] (Figure 1B). Despite differences in Q-length between Orb2 and HTTex1, a recent study inferred a similar interdigitated antiparallel hairpin structure as the stereochemically favorable arrangement of in vitro-assembled HTTex1 [43]. The tight interdigitation of glutamine sidechains in the packing of antiparallel β-sheets in HTTQex1 requires these sidechains to adopt two different rotamers for different strands [43], indicating structural heterogeneity at single-residue level. Specifically, the Qs of the two β-strands in the HTTQex1 β-hairpin differ in their side chain dihedral angles; however, within each strand, all Q residues are the same rotamers with the same backbone and sidechain geometry [28,43]. In this antiparallel arrangement, inter-strand hydrogen bonds are only formed between β-strands with different sidechain rotamers, but not between β-strands with the same rotamer [28,43]. On the other hand, the packing of parallel β-sheets in the Orb2 core can be achieved with the same rotamer for all interdigitating glutamines [42]. In addition, the Orb2 parallel β-structure could be stabilized by specific features of its interdigitated cross-β packing: a slight tilt of the glutamine sidechains toward the N-termini of the β-strands and the positioning of strands in one β-sheet opposite the inter-strand spaces in the other β-sheet and vice versa. These arrangements allow the formation of additional, stabilizing hydrogen bonds between the glutamine sidechains groups in one β-sheet and the carbonyl oxygen atom of main chain peptides of the opposite β-sheet, with little or no effect on peptide group conformation, as well as a tighter packing of the interdigitated glutamine sidechains from both β-sheets [42].
The Orb2 and HTTex1 antiparallel hairpin arrangement differs in the overall amyloid architecture. The hairpin-like fold of the individual Orb2 chains is formed by the strands from two opposing parallel β-sheets [42] (Figure 2A). In contrast, the in vitro-assembled HTT model suggests the HTTQex1 β-hairpin is made of two hydrogen-bonded antiparallel strands in the same β-sheet [43] (Figure 2B). However, in vivo, the structure and/or arrangement of HTTex1 β-hairpin could be dictated by context-specific factors [30], which could lead to structurally different conformations to that assembled in vitro. Indeed, activity and structure of in vitro-assembled Orb2 amyloid is distinct from Orb2 amyloid isolated from adult brain [42,44].
Is it possible that the endogenous, aggregated HTT structure is distinct from what has been inferred from in vitro studies? If so, could it be similar to Orb2 structure? Possible Orb2 and HTT structural similarity in the native context, besides a similar monomeric fold (i.e., hairpin) in the amyloid state, is underscored by the observation that exogenously expressed Orb2 co-aggregates with HTTex1 in Drosophila motor neurons [36]. This observation could be explained by different scenarios: First, the co-localization could result from the incorporation of HTTex1 monomers into the Orb2 filament arrangement forming a heteroaggregate, or vice versa. Second, HTTex1 and Orb2 adopt a different structure, and the co-localization arises from a lateral surface association of HTTex1 and Orb2 filaments. Third, endogenous HTTex1 adopts a structure similar to Orb2. Here, the interdigitated cross-β structure observed in head-extracted Orb2 filaments could be readily extended on both sides of a parallel β-sheet made of only glutamine residues (Figure 2C). Such an arrangement would allow the formation of more stable, multilayered cross-β structures from sufficiently long polyQ sequences based on hairpins with similar β-strand lengths as minimal repeat units (Figure 2D). Indeed, contrary to earlier reports [24,45,46], ssNMR data consistently report a length-independent common structure of the polyQ amyloid [26,47]. This observation may reflect the existence of a unique polyQ structure in all polyQ diseases, where protein context (such as flanking regions), or cell-type-specific context (such as monomer availability), could determine the Qn threshold [48] and supramolecular filament polymorphism [47,49]. In the future, the elucidation of structure of HTT and other polyQ aggregates from diseased brains would either refute or support this thesis.

Author Contributions

All three authors contributed to the preparation of this review. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Stowers Institute for Medical Research (to Kausik Si).

Acknowledgments

We thank Douglas V. Laurents for comments and M. Miller for illustrations.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

PolyQPolyglutamine
HDHuntington’s Disease
HTTHuntingtin
HTTex1Huntingtin Exon 1
CPEBCytoplasmic Polyadenylation Element Binding Protein
Orb2The Drosophila protein encoded by the oo18 RNA-binding (orb) gene
Cryo-EMCryo-Electron Microscopy
Cryo-ETCryo-Electron Tomography
ssNMRSolid-state Nuclear Magnetic Resonance

References

  1. Orr, H.T.; Zoghbi, H.Y. Trinucleotide Repeat Disorders. Annu. Rev. Neurosci. 2007, 30, 575–621. [Google Scholar] [CrossRef] [PubMed]
  2. Smith, M.A.; Brandt, J.; Shadmehr, R. Motor disorder in Huntington’s disease begins as a dysfunction in error feedback control. Nature 2000, 403, 544–549. [Google Scholar] [CrossRef] [PubMed]
  3. Saudou, F.; Humbert, S. The Biology of Huntingtin. Neuron 2016, 89, 910–926. [Google Scholar] [CrossRef] [Green Version]
  4. Guo, Q.; Huang, B.; Cheng, J.; Seefelder, M.; Engler, T.; Pfeifer, G.; Oeckl, P.; Otto, M.; Moser, F.; Maurer, M.; et al. The cryo-electron microscopy structure of huntingtin. Nature 2018, 555, 117–120. [Google Scholar] [CrossRef]
  5. MacDonald, M.E.; Ambrose, C.M.; Duyao, M.P.; Myers, R.H.; Lin, C.; Srinidhi, L.; Barnes, G.; Taylor, S.A.; James, M.; Groot, N.; et al. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington’s disease chromosomes. Cell 1993, 72, 971–983. [Google Scholar] [CrossRef]
  6. Kim, Y.J.; Yi, Y.; Sapp, E.; Wang, Y.; Cuiffo, B.; Kegel, K.B.; Qin, Z.H.; Aronin, N.; DiFiglia, M. Caspase 3-cleaved N-terminal fragments of wild-type and mutant huntingtin are present in normal and Huntington’s disease brains, associate with membranes, and undergo calpaindependent proteolysis. Proc. Natl. Acad. Sci. USA 2001, 98, 12784–12789. [Google Scholar] [CrossRef] [Green Version]
  7. Scherzinger, E.; Lurz, R.; Turmaine, M.; Mangiarini, L.; Hollenbach, B.; Hasenbank, R.; Bates, G.P.; Davies, S.W.; Lehrach, H.; Wanker, E.E. Huntingtin-encoded polyglutamine expansions form amyloid-like protein aggregates in vitro and in vivo. Cell 1997, 90, 549–558. [Google Scholar] [CrossRef] [Green Version]
  8. Schilling, G.; Klevytska, A.; Tebbenkamp, A.T.N.; Juenemann, K.; Cooper, J.; Gonzales, V.; Slunt, H.; Poirer, M.; Ross, C.A.; Borchelt, D.R. Characterization of Huntingtin Pathologic Fragments in Human Huntington Disease, Transgenic Mice, and Cell Models. J. Neuropathol. Exp. Neurol. 2007, 66, 313–320. [Google Scholar] [CrossRef] [Green Version]
  9. Jennifer Morton, A.; Howland, D.S. Large genetic animal models of huntington’s disease. J. Huntingt. Dis. 2013, 2, 3–19. [Google Scholar] [CrossRef] [Green Version]
  10. Stricker-Shaver, J.; Novati, A.; Yu-Taeger, L.; Nguyen, H.P. Genetic rodent models of huntington disease. In Advances in Experimental Medicine and Biology; Springer LLC: New York, NY, USA, 2018; Volume 1049, pp. 29–57. [Google Scholar]
  11. Adegbuyiro, A.; Sedighi, F.; Pilkington, A.W.; Groover, S.; Legleiter, J. Proteins Containing Expanded Polyglutamine Tracts and Neurodegenerative Disease. Biochemistry 2017, 56, 1199–1217. [Google Scholar] [CrossRef] [Green Version]
  12. Nucifora, J.; Sasaki, M.; Peters, M.F.; Huang, H.; Cooper, J.K.; Yamada, M.; Takahashi, H.; Tsuji, S.; Troncoso, J.; Dawson, V.L.; et al. Interference by huntingtin and atrophin-1 with CBP-mediated transcription leading to cellular toxicity. Science 2001, 291, 2423–2428. [Google Scholar] [CrossRef] [Green Version]
  13. Bence, N.F.; Sampat, R.M.; Kopito, R.R. Impairment of the ubiquitin-proteasome system by protein aggregation. Science 2001, 292, 1552–1555. [Google Scholar] [CrossRef]
  14. Hoffner, G.; Djian, P. Polyglutamine Aggregation in Huntington Disease: Does Structure Determine Toxicity? Mol. Neurobiol. 2015, 52, 1297–1314. [Google Scholar] [CrossRef] [PubMed]
  15. Drombosky, K.W.; Rode, S.; Kodali, R.; Jacob, T.C.; Palladino, M.J.; Wetzel, R. Mutational analysis implicates the amyloid fibril as the toxic entity in Huntington’s disease. Neurobiol. Dis. 2018, 120, 126–138. [Google Scholar] [CrossRef] [PubMed]
  16. Nelson, R.; Sawaya, M.R.; Balbirnie, M.; Madsen, A.; Riekel, C.; Grothe, R.; Eisenberg, D. Structure of the cross-β spine of amyloid-like fibrils. Nature 2005, 435, 773–778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Geddes, A.J.; Parker, K.D.; Atkins, E.D.T.; Beighton, E. “Cross-β” conformation in proteins. J. Mol. Biol. 1968, 32, 343–358. [Google Scholar] [CrossRef]
  18. Thakur, A.K.; Jayaraman, M.; Mishra, R.; Thakur, M.; Chellgren, V.M.; Byeon, I.J.L.; Anjum, D.H.; Kodali, R.; Creamer, T.P.; Conway, J.F.; et al. Polyglutamine disruption of the huntingtin exon 1 N terminus triggers a complex aggregation mechanism. Nat. Struct. Mol. Biol. 2009, 16, 380–389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Williamson, T.E.; Vitalis, A.; Crick, S.L.; Pappu, R.V. Modulation of Polyglutamine Conformations and Dimer Formation by the N-Terminus of Huntingtin. J. Mol. Biol. 2010, 396, 1295–1309. [Google Scholar] [CrossRef] [Green Version]
  20. Rockabrand, E.; Slepko, N.; Pantalone, A.; Nukala, V.N.; Kazantsev, A.; Marsh, J.L.; Sullivan, P.G.; Steffan, J.S.; Sensi, S.L.; Thompson, L.M. The first 17 amino acids of Huntingtin modulate its sub-cellular localization, aggregation and effects on calcium homeostasis. Hum. Mol. Genet. 2007, 16, 61–77. [Google Scholar] [CrossRef]
  21. Atwal, R.S.; Xia, J.; Pinchev, D.; Taylor, J.; Epand, R.M.; Truant, R. Huntingtin has a membrane association signal that can modulate huntingtin aggregation, nuclear entry and toxicity. Hum. Mol. Genet. 2007, 16. [Google Scholar] [CrossRef] [Green Version]
  22. Bhattacharyya, A.; Thakur, A.K.; Chellgren, V.M.; Thiagarajan, G.; Williams, A.D.; Chellgren, B.W.; Creamer, T.P.; Wetzel, R. Oligoproline effects on polyglutamine conformation and aggregation. J. Mol. Biol. 2006, 355, 524–535. [Google Scholar] [CrossRef]
  23. Darnell, G.; Orgel, J.P.R.O.; Pahl, R.; Meredith, S.C. Flanking Polyproline Sequences Inhibit β-Sheet Structure in Polyglutamine Segments by Inducing PPII-like Helix Structure. J. Mol. Biol. 2007, 374, 688–704. [Google Scholar] [CrossRef] [PubMed]
  24. Sharma, D.; Shinchuk, L.M.; Inouye, H.; Wetzel, R.; Kirschner, D.A. Polyglutamine homopolymers having 8-45 residues form slablike β-crystallite assemblies. Proteins Struct. Funct. Genet. 2005, 61, 398–411. [Google Scholar] [CrossRef]
  25. Schneider, R.; Schumacher, M.C.; Mueller, H.; Nand, D.; Klaukien, V.; Heise, H.; Riedel, D.; Wolf, G.; Behrmann, E.; Raunser, S.; et al. Structural characterization of polyglutamine fibrils by solid-state NMR spectroscopy. J. Mol. Biol. 2011, 412, 121–136. [Google Scholar] [CrossRef] [Green Version]
  26. Sivanandam, V.N.; Jayaraman, M.; Hoop, C.L.; Kodali, R.; Wetzel, R.; Van Der Wel, P.C.A. The aggregation-enhancing huntingtin N-terminus is helical in amyloid fibrils. J. Am. Chem. Soc. 2011, 133, 4558–4566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Tanaka, M.; Morishima, I.; Akagi, T.; Hashikawa, T.; Nukina, N. Intra- and Intermolecular β-Pleated Sheet Formation in Glutamine-repeat Inserted Myoglobin as a Model for Polyglutamine Diseases. J. Biol. Chem. 2001, 276, 45470–45475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Smith, A.N.; Märker, K.; Piretra, T.; Boatz, J.C.; Matlahov, I.; Kodali, R.; Hediger, S.; Van Der Wel, P.C.A.; De Paëpe, G. Structural Fingerprinting of Protein Aggregates by Dynamic Nuclear Polarization-Enhanced Solid-State NMR at Natural Isotopic Abundance. J. Am. Chem. Soc. 2018, 140, 14576–14580. [Google Scholar] [CrossRef]
  29. Sharma, D.; Sharma, S.; Pasha, S.; Brahmachari, S.K. Peptide models for inherited neurodegenerative disorders: Conformation and aggregation properties of long polyglutamine peptides with and without interruptions. FEBS Lett. 1999, 456, 181–185. [Google Scholar] [CrossRef] [Green Version]
  30. Nekooki-Machida, Y.; Kurosawa, M.; Nukina, N.; Ito, K.; Oda, T.; Tanaka, M. Distinct conformations of in vitro and in vivo amyloids of huntingtin-exon1 show different cytotoxicity. Proc. Natl. Acad. Sci. USA 2009, 106, 9679–9684. [Google Scholar] [CrossRef] [Green Version]
  31. Buchanan, L.E.; Carr, J.K.; Fluitt, A.M.; Hoganson, A.J.; Moran, S.D.; De Pablo, J.J.; Skinner, J.L.; Zanni, M.T. Structural motif of polyglutamine amyloid fibrils discerned with mixed-isotope infrared spectroscopy. Proc. Natl. Acad. Sci. USA 2014, 111, 5796–5801. [Google Scholar] [CrossRef] [Green Version]
  32. Xiong, K.; Punihaole, D.; Asher, S.A. UV resonance raman spectroscopy monitors polyglutamine backbone and side chain hydrogen bonding and fibrillization. Biochemistry 2012, 51, 5822–5830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Monsellier, E.; Redeker, V.; Ruiz-Arlandis, G.; Bousset, L.; Melki, R. Molecular interaction between the chaperone Hsc70 and the N-terminal flank of huntingtin exon 1 modulates aggregation. J. Biol. Chem. 2015, 290, 2560–2576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Bugg, C.W.; Isas, J.M.; Fischer, T.; Patterson, P.H.; Langen, R. Structural features and domain organization of huntingtin fibrils. J. Biol. Chem. 2012, 287, 31739–31746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Bäuerlein, F.J.B.; Saha, I.; Mishra, A.; Kalemanov, M.; Martínez-Sánchez, A.; Klein, R.; Dudanova, I.; Hipp, M.S.; Hartl, F.U.; Baumeister, W.; et al. In Situ Architecture and Cellular Interactions of PolyQ Inclusions. Cell 2017, 171, 179–187. [Google Scholar] [CrossRef] [Green Version]
  36. Hervás, R.; Li, L.; Majumdar, A.; del Fernández-Ramírez, M.C.; Unruh, J.R.; Slaughter, B.D.; Galera-Prat, A.; Santana, E.; Suzuki, M.; Nagai, Y.; et al. Molecular Basis of Orb2 Amyloidogenesis and Blockade of Memory Consolidation. PLoS Biol. 2016, 14, e1002361. [Google Scholar] [CrossRef] [Green Version]
  37. Keleman, K.; Krüttner, S.; Alenius, M.; Dickson, B.J. Function of the Drosophila CPEB protein Orb2 in long-term courtship memory. Nat. Neurosci. 2007, 10, 1587–1593. [Google Scholar] [CrossRef]
  38. Krüttner, S.; Traunmüller, L.; Dag, U.; Jandrasits, K.; Stepien, B.; Iyer, N.; Fradkin, L.G.; Noordermeer, J.N.; Mensh, B.D.; Keleman, K. Synaptic Orb2A Bridges Memory Acquisition and Late Memory Consolidation in Drosophila. Cell Rep. 2015, 11, 1953–1965. [Google Scholar] [CrossRef] [Green Version]
  39. Krüttner, S.; Stepien, B.; Noordermeer, J.N.; Mommaas, M.A.; Mechtler, K.; Dickson, B.J.; Keleman, K. Drosophila CPEB Orb2A Mediates Memory Independent of Its RNA-Binding Domain. Neuron 2012, 76, 383–395. [Google Scholar] [CrossRef] [Green Version]
  40. Majumdar, A.; Cesario, W.C.; White-Grindley, E.; Jiang, H.; Ren, F.; Khan, M.R.; Li, L.; Choi, E.M.L.; Kannan, K.; Guo, F.; et al. Critical role of amyloid-like oligomers of Drosophila Orb2 in the persistence of memory. Cell 2012, 148, 515–529. [Google Scholar] [CrossRef] [Green Version]
  41. Li, L.; Sanchez, C.P.; Slaughter, B.D.; Zhao, Y.; Khan, M.R.; Unruh, J.R.; Rubinstein, B.; Si, K. A Putative Biochemical Engram of Long-Term Memory. Curr. Biol. 2016, 26, 3143–3156. [Google Scholar] [CrossRef] [Green Version]
  42. Hervas, R.; Rau, M.J.; Park, Y.; Zhang, W.; Murzin, A.G.; Fitzpatrick, J.A.J.; Scheres, S.H.W.; Si, K. Cryo-EM structure of a neuronal functional amyloid implicated in memory persistence in Drosophila. Science 2020, 367, 1230–1234. [Google Scholar] [CrossRef]
  43. Hoop, C.L.; Lin, H.K.; Kar, K.; Magyarfalvi, G.; Lamley, J.M.; Boatz, J.C.; Mandal, A.; Lewandowski, J.R.; Wetzel, R.; Van Der Wel, P.C.A. Huntingtin exon 1 fibrils feature an interdigitated β-hairpin-based polyglutamine core. Proc. Natl. Acad. Sci. USA 2016, 113, 1546–1551. [Google Scholar] [CrossRef] [Green Version]
  44. Cervantes, S.A.; Bajakian, T.H.; Soria, M.A.; Falk, A.S.; Service, R.J.; Langen, R.; Siemer, A.B. Identification and Structural Characterization of the N-terminal Amyloid Core of Orb2 isoform A. Sci. Rep. 2016, 6, 1–11. [Google Scholar] [CrossRef] [Green Version]
  45. Stanley, C.B.; Perevozchikova, T.; Berthelier, V. Structural formation of huntingtin exon 1 aggregates probed by small-angle neutron scattering. Biophys. J. 2011, 100, 2504–2512. [Google Scholar] [CrossRef] [Green Version]
  46. Perevozchikova, T.; Stanley, C.B.; McWilliams-Koeppen, H.P.; Rowe, E.L.; Berthelier, V. Investigating the structural impact of the glutamine repeat in huntingtin assembly. Biophys. J. 2014, 107, 411–421. [Google Scholar] [CrossRef] [Green Version]
  47. Lin, H.K.; Boatz, J.C.; Krabbendam, I.E.; Kodali, R.; Hou, Z.; Wetzel, R.; Dolga, A.M.; Poirier, M.A.; Van Der Wel, P.C.A. Fibril polymorphism affects immobilized non-amyloid flanking domains of huntingtin exon1 rather than its polyglutamine core. Nat. Commun. 2017, 8, 1–12. [Google Scholar] [CrossRef] [Green Version]
  48. Du, X.; Gomez, C.M. Spinocerebellum ataxia type 6: Molecular mechanisms and calcium channel genetics. In Advances in Experimental Medicine and Biology; Springer LLC: New York, NY, USA, 2018; Volume 1049, pp. 147–173. [Google Scholar]
  49. Boatz, J.C.; Piretra, T.; Lasorsa, A.; Matlahov, I.; Conway, J.F.; van der Wel, P.C.A. Protofilament structure and supramolecular polymorphism of aggregated mutant huntingtin exon 1. J. Mol. Biol. 2020. [Google Scholar] [CrossRef]
Figure 1. General view of the Orb2 amyloid structure. Side view of the reconstructed Orb2 amyloid core showing the ~4.75 Å separation between β-strands, typical of the amyloid fold (A), and cross-sectional view of one molecular layer of the calculated atomic model (B) [42]. Glutamines are colored in dark blue, while histidines are colored in light blue. In some histidine residues, a major and minor occupancy, alternative sidechain conformations are shown (arrow heads). Leucine and serine residues are represented in gray.
Figure 1. General view of the Orb2 amyloid structure. Side view of the reconstructed Orb2 amyloid core showing the ~4.75 Å separation between β-strands, typical of the amyloid fold (A), and cross-sectional view of one molecular layer of the calculated atomic model (B) [42]. Glutamines are colored in dark blue, while histidines are colored in light blue. In some histidine residues, a major and minor occupancy, alternative sidechain conformations are shown (arrow heads). Leucine and serine residues are represented in gray.
Ijms 21 06910 g001
Figure 2. Molecular architecture of Orb2 and Huntingtin amyloids. (A) Schematic of the antiparallel hairpin-like fold adopted by head-extracted Orb2 filaments, derived from the cryo-EM structure. Three Orb2 molecules per molecular layer form continuous in-register parallel β-sheets. Different tone of blue represents the different amino acid composition for each β-strand of the hairpin. Amyloid forming sequence is indicated in the top. (B) Schematic of the antiparallel β-hairpin adopted by in vitro-assembled HTTex1 amyloid, derived from ssNMR data. One single HTTex1 molecule contributes to two molecular layers to form antiparallel β-sheets. Different tone of green represents the two differently structured β-strand types of the β-hairpin. (C) Model of a multilayer packing of the parallel polyQ β-sheets, obtained by extending the Orb2 inter-digitated cross-β structure on both sides. Blue dashed line represents the hairpin turn. The extended glutamine side chains form a steric zipper interface to allow an ~8 Å distance between β-sheets. (D) Proposed in vivo HTTex1 filament model based on the multilayer polyQ packing showed in (C). Stacks of hairpins or meanders of similar β-strand lengths (highlighted in red), are viewed across the filament axis.
Figure 2. Molecular architecture of Orb2 and Huntingtin amyloids. (A) Schematic of the antiparallel hairpin-like fold adopted by head-extracted Orb2 filaments, derived from the cryo-EM structure. Three Orb2 molecules per molecular layer form continuous in-register parallel β-sheets. Different tone of blue represents the different amino acid composition for each β-strand of the hairpin. Amyloid forming sequence is indicated in the top. (B) Schematic of the antiparallel β-hairpin adopted by in vitro-assembled HTTex1 amyloid, derived from ssNMR data. One single HTTex1 molecule contributes to two molecular layers to form antiparallel β-sheets. Different tone of green represents the two differently structured β-strand types of the β-hairpin. (C) Model of a multilayer packing of the parallel polyQ β-sheets, obtained by extending the Orb2 inter-digitated cross-β structure on both sides. Blue dashed line represents the hairpin turn. The extended glutamine side chains form a steric zipper interface to allow an ~8 Å distance between β-sheets. (D) Proposed in vivo HTTex1 filament model based on the multilayer polyQ packing showed in (C). Stacks of hairpins or meanders of similar β-strand lengths (highlighted in red), are viewed across the filament axis.
Ijms 21 06910 g002

Share and Cite

MDPI and ACS Style

Hervás, R.; Murzin, A.G.; Si, K. Implications of the Orb2 Amyloid Structure in Huntington’s Disease. Int. J. Mol. Sci. 2020, 21, 6910. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21186910

AMA Style

Hervás R, Murzin AG, Si K. Implications of the Orb2 Amyloid Structure in Huntington’s Disease. International Journal of Molecular Sciences. 2020; 21(18):6910. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21186910

Chicago/Turabian Style

Hervás, Rubén, Alexey G. Murzin, and Kausik Si. 2020. "Implications of the Orb2 Amyloid Structure in Huntington’s Disease" International Journal of Molecular Sciences 21, no. 18: 6910. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21186910

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