1,530
Views
2
CrossRef citations to date
0
Altmetric
Review

Copper coordination modulates prion conversion and infectivity in mammalian prion proteins

ORCID Icon
Pages 1-6 | Received 07 Oct 2022, Accepted 22 Dec 2022, Published online: 03 Jan 2023

ABSTRACT

In mammals the cellular form of the prion protein (PrPC) is a ubiquitous protein involved in many relevant functions in the central nervous system. In addition to its physiological functions PrPC plays a central role in a group of invariably fatal neurodegenerative disorders collectively called prion diseases. In fact, the protein is a substrate in a process in which it converts into an infectious and pathological form denoted as prion. The protein has a unique primary structure where the unstructured N-terminal moiety possesses characteristic sequences wherein histidines are able to coordinate metal ions, in particular copper ions. These sequences are called octarepeats for their characteristic length. Moreover, a non-octarepeat fifth-copper binding site is present where copper coordination seems to control infectivity. In this review, I will argue that these sequences may play a significant role in modulating prion conversion and replication.

Introduction

The cellular form of the prion protein (PrPC) is a physiological and ubiquitous molecule present in all mammal tissues [Citation1]. It is particularly abundant in the central nervous system (CNS) where it is involved in the pathogenesis of various neurodegenerative diseases [Citation2], known as transmissible spongiform encephalopathies (TSEs). The protein is either the substrate for the conversion, replication and accumulation into a pathogenic isoform denominated prion or PrPSc [Citation3], or it may also serve as a promiscuous receptor for amyloidal forms of other proteins involved in neurodegenerative disorders [Citation4]. Indeed, PrPC has been found to interact and facilitate the uptake of various amyloid structures of proteins such as alpha-synuclein, tau and more recently the TAR DNA-binding protein 43 or TDP-43 [Citation5–7]. The cellular prion protein, PrPC, is involved in different functions within the CNS, which include, among others, NMDAR regulation and neuritogenesis [Citation8,Citation9]. The protein has also been found involved in regulating the direction of synaptic plasticity in the hippocampus through protein kinase A [Citation10]. For the many functions carried out by PrPC, the protein can be defined as a pleiotropic protein [Citation1].

Prion protein

The primary structure of PrPC can be divided into an N-terminal unstructured, intrinsically disordered protein and well-structured and defined C-terminal moiety. At the amino terminal end an ER-localization sequence is found at residues 1 to 22 and the first amino acid, a lysine, starts at position 23. The mature sequence terminates at position 231 with a serine. The latter is further modified by the insertion of a glycosylphophatidylinositol group (by means of another processed signal sequence in position 232–254), which in turn anchors the protein to the outer leaflet of the plasma membrane [Citation1].

In the mature form the protein is composed of either 208 or 209 amino acids (depending whether mouse or human sequences, respectively) with unique feature in its primary structure [Citation11].

As mentioned above the N-terminal moiety is unstructured and it is characterized by the presence of four unique octapeptide repeat sequences (PHGGGWGQ) where the highlighted (in bold) histidine (His) within the sequence carries out a major function of coordinating divalent anions such as copper and to a lesser extent zinc [Citation12,Citation13]. These octapeptides structures are followed by a hydrophobic domain, with two main features: a so-called non-octarepeat fifth-copper binding site (95-HNQWNKPSKPKTNLKH-110) and a characteristic palindromic sequence (111-AGAAAAGA-119), which seems important for conversion of PrPC to PrPSc [Citation14]. The non-octarepeat fifth-copper binding site contains two highlighted (in bold) hystidine (His) residues which both coordinate mostly copper ions [Citation15].

The structured part of PrPC comprises three alpha-helices and three short beta-sheets. The first two beta-sheets (beta0 and beta1) precede the first alpha-helix, then a second beta-sheet (beta2) anticipates the last two helices [Citation16]. The latter contain a single disulphide bridge and two sites for N-glycosylation [Citation17].

All these PrPC features are shown schematically in .

Figure 1. A cartoon representation of the primary structure of the mature mouse PrPC. The primary structure of the mature mouse PrPC is shown. The highlighted amino acid sequences are shown of the N-terminal part including the octarepeats and the non-octarepeat fifth-copper binding site. The histidine residues involved in copper coordination are shown in bold either in red or in black. The single disulphide link and the two glycosylation sites are shown at the C-terminal domain.

Figure 1. A cartoon representation of the primary structure of the mature mouse PrPC. The primary structure of the mature mouse PrPC is shown. The highlighted amino acid sequences are shown of the N-terminal part including the octarepeats and the non-octarepeat fifth-copper binding site. The histidine residues involved in copper coordination are shown in bold either in red or in black. The single disulphide link and the two glycosylation sites are shown at the C-terminal domain.

Copper coordination

Of the many unique characteristics of PrPC, certainly the distinctive capacity of coordinating copper ions has attracted much attention. Since the octapeptide repeat sequences and the non-octarepeat fifth-copper binding site are unique traits within the molecule consequently their role in both physiology and pathology has been explored in depth. Following the striking evidence that PrPC alters copper content in vivo [Citation18], the first report accounting for PrPC binding to copper(II) ions was published about 25 years ago [Citation19]. In this paper, using biophysical techniques such as near-UV circular dichroism (CD) and tryptophan (Trp) fluorescence spectroscopy, the authors offered evidence that copper(II) [or Cu(II)] induces conformational changes of PrP and indicate that additional equilibrium dialysis experiments are consistent with a stoichiometry of 2 ions per molecule. Only few years later by means of electron paramagnetic resonance (EPR) the ratio between copper(II) and octarepeat is found as 1:1 regardless the number of repeats present [Citation20]. The atomic details of the coordination of copper within a single octarepeat sequence has been resolved indicating the direct involvement of glycine residues (Gly), besides the histidine, which revealed equatorial coordination by the histidine imidazole, two deprotonated glycine amides, a glycine carbonyl, along with an axial water bridging the tryptophan (Trp) indole [Citation13]. Shortly after these seminal studies, another important step forward into the understanding of copper coordination in full-length PrPC indicated the relevance of the non-octarepeat fifth-copper binding site in addition to the octapeptide repeat sequences [Citation15].

The complete coordination structures of copper(II) binding to the various PrP sites have been reviewed recently [Citation21]. The three-dimensional structure of PrPC showing the coordination of copper ions at the unstructured N-terminal domain is shown in .

Figure 2. The three-dimensional (3D) structure of PrPC. The human 3D structure of PrPC is shown in which the unstructured N-terminal domain is presented as a model of copper coordination. The C-terminal domain is represented based on structures solved by NMR.

Figure 2. The three-dimensional (3D) structure of PrPC. The human 3D structure of PrPC is shown in which the unstructured N-terminal domain is presented as a model of copper coordination. The C-terminal domain is represented based on structures solved by NMR.

Physiology

Therefore, it would be difficult to consider PrPC simply an apoprotein, without the coordination of copper. In other words, it seems that PrPC requires the presence of copper in order to carry out its functions. The importance of copper coordination has been well-studied at least for its neuritogenesis function [Citation22]. In addition, copper coordination in the whole prion protein is key in the regulation of NMDR receptor where the Cu(II) and Cu(I) coordination play a major role [Citation8]. Instead, the role of copper in modulating prion conversion, replication and infectivity has been controversial for some time. Indeed, different reports indicate copper coordination to either promote or inhibit prion conversion [Citation23,Citation24]. In this review, I will try to present recent evidence that support the role of copper coordination in modulating prion conversion and infectivity.

Copper coordination and prion conversion and infectivity

In biological terms, prions are infectious proteins that lead to neurodegeneration in mammals. In laboratory animals, upon inoculation in the CNS or other routes, prions will lead to a given set of neuropathological and biochemical changes that will define the infectious proteins. The inoculation can occur in a variety of different tissues but then the pathological changes and incubation times may vary depending on the site of inoculation. Regardless the site of inoculation, the outcome will be neurodegeneration of the CNS.

Prion infectivity is defined by several criteria. First, any given prion specimen will lead to characteristic incubation times that depend on the prion itself (prion strain), the homology of PrPC between the prion inoculum and the animal used in the experiment, and to the genetic background of the recipient laboratory animal (usually an inbred or outbred wild-type mouse strain, or transgenic mouse strains). The combination of these parameters for any given prion strain will harbour invariably similar incubation times, neuropathological changes and biochemical features over time. In fact, models of prion disorders in laboratory animal are the most reliable experimental representation of the human disease counterparts.

So, what does it make an infectious prion? At the molecular level, the structure of the prion is essential for its infectivity. Indeed, in human prion diseases there is a variety of molecular evidence about the nature of prions. For long time the neuropathological occurrence of prion protein amyloid deposits in the CNS was believed to be the main characteristic of prions. Nevertheless, several human prion disorders present neurodegeneration without prion amyloid deposits, thus implying that prion amyloid formation is a non-obligatory feature of these disorders. This is the case for some forms of fatal familial insomnia [Citation25]. A prion can exist in an infectious form even without forming detectable amyloidal structure (see below for prion structures). Early studies on the required molecular features for infectious prion showed that the palindromic sequence (111-AGAAAAGA-119) within the N-terminal hydrophobic domain is essential for prion formation and conversion. Deletion mutants lacking the sequence AGAAAAGA did not convert and replicate using persistently scrapie-infected mouse neuroblastoma cells as a model system for prion replication [Citation26,Citation27].

Proteolytic processing of PrP has been known to provide further evidence of the relevance and importance of copper-PrP interactions because the resulting fragments could provide a high affinity binding site for Cu(II) [Citation28].

Supporting the role of the non-octarepeat fifth-copper binding site in prion conversion, transgenic mice were generated with an amino acid replacement at residue H95 with a glycine, which lead to shorter incubation times upon inoculation with infectious prions compared to wild-type animals [Citation29].

In addition, structural evidence has reinforced the crucial role of the hydrophobic domain in the formation of infectious prions [Citation30].

Studies with synthetic prions, which offer a relevant molecular model of native prions, served as an important indication about the role of copper coordination and prion conversion and infectivity in laboratory animals. Using a transgenic mouse model for prion replication denoted as Tg9949, it was possible to establish that the octapeptide repeat sequences were dispensable for prion formation, conversion, accumulation and infectivity [Citation31]. This model of transgenic mice expresses a truncated form of the prion protein as a transgene, where the sequence is between residues 89 to 230 (start at amino acid 89, which is a glycine). In this sequence, as previously shown by Flechsig and colleagues, the octapeptide repeat sequences are missing but the protein still does convert to an infectious prion [Citation31,Citation32]. After a series of studies that strengthen this evidence, both the non-octarepeat fifth-copper binding site and the palindromic sequence AGAAAAGA, indicate a clear role for the hydrophobic domain in the production of prions. In addition, complete deletion of both octapeptide repeat sequences and hydrophobic domain leads to a toxic molecule denoted as PrP121-230 [Citation33].

Therefore, it could be argued that the minimal structure for infectious prion should start at the hydrophobic domain and should cover the remaining of the molecule, at least up to amino acid 145 [Citation34].

Of note, there a series of human mutant prion proteins in which the octapeptide repeat sequences are expanded (up to nine) and the number of extra octapeptide repeat sequences inversely correlates with disease onset [Citation35].

One important evidence for the involvement of copper coordination in determining an infectious prion came from structural characterization of mutant pathogenic prion proteins.

Structural studies on human PrP with pathological Q212P, a mutation linked to Gerstmann-Sträussler-Scheinker (GSS) syndrome, is of relevance [Citation36,Citation37]. Using synchrotron radiation-based X-ray absorption fine structure technique, Cu(II) and Cu(I) coordination geometries in the mutant Q212P were compared to those of the wild-type protein. Analysing the extended X-ray absorption fine structure and the X-ray absorption near-edge structure, highlighted changes in copper coordination induced by the mutation Q212P in both oxidation states. While in the wild-type protein the copper-binding site has the same structure for both Cu(II) and Cu(I), in the mutant the coordination sites drastically change from the oxidized to the reduced form of the copper ion. Copper-binding sites in the mutant confirmed the loss of short- and long-range interactions [Citation37].

More recently, another important contribution came from findings that highly conserved histidines in the C-terminal domain are crucial to hold together both the N-terminal and C-terminal moieties of PrPC through Cu(II) coordination [Citation38].

These results clearly indicate that point mutations in the C-terminal portion of PrP could affect copper coordination at the N-terminus and in particular copper binding at the non-octarepeat fifth-copper binding site.

In addition, one should also consider that the N-terminal domain of PrPC is involved in the interaction with cellular partners in carrying out its function and thereferore copper coordination could compete with these protein binding sites and/or promote the formation of ternary complexes [Citation39].

In order to model these effects, mutant proteins bearing either H96 or H111 mutations (human numbering) should lead to PrP molecules structural distinct in the non-octarepeat fifth-copper binding site [Citation40]. In fact, the involvement of H96 and H111 in the non-octarepeat fifth-copper binding site shows that Cu(II) occupancy may play a crucial role in determining PrPC conformation. Indeed, altered Cu(II) coordination due to a mutation that abrogates copper coordination, a tyrosine in place of the histidine (H96Y), may lead to spontaneous PrPSc formation in cultured neuronal cells and accumulation in the acid compartments. My group has proposed a model whereby HuPrP coordinating copper with His96 and His111 in the non-octarepeat fifth-copper binding site is more resistant to prion conversion compared to a PrPC molecule coordinating Cu(II) with one histidine [Citation40,Citation41]. This work has been further supported by using a series of different PrP from several specie polymorphisms and mutants [Citation42].

Conclusions and further prospective

In this brief review, I highlighted the current literature that deals with copper coordination in the prion protein and the role of this metal in prion conversion and replication. It seems that copper may play a crucial role in determining whether a prion protein molecule becomes a prion.

Recently, a major step forward in the field of prion biology has been the resolution of the first structures of infectious prions [Citation43–49]. It would be important to unveil the role of copper in generating such diverse structures for different prion strains.

It would also be quite important to carry out further studies on different mutants in place of the histidines 96 and 111 (human numbering) or 95 and 110 (mouse numbering) in the non-octarepeat fifth-copper binding site, in order to further establish their role in rendering PrP an infectious prion. The role of the non-octarepeat fifth-copper binding site seems to be crucial in establishing the structural determinant that leads to the formation of prions.

Acknowledgments

The author wishes to thank Drs Gabriele Giachin and Fabio Moda for helpful discussions and suggestions. has been kindly provided by Ms Francesca Katharina Legname. has been kindly provided by Dr Gabriele Giachin.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

The author(s) reported there is no funding associated with the work featured in this article.

References

  • Legname G. Elucidating the function of the prion protein. PLoS Pathog. 2017 Aug;13(8):e1006458.
  • Prusiner SB. Biology and genetics of prions causing neurodegeneration. Annu Rev Genet. 2013;47:601–623.
  • Prusiner SB. Cell biology. A unifying role for prions in neurodegenerative diseases. Science. 2012 Jun 22;336(6088):1511–1513.
  • Scialo C, Legname G. The role of the cellular prion protein in the uptake and toxic signaling of pathological neurodegenerative aggregates. Prog Mol Biol Transl Sci. 2020;175:297–323.
  • Aulic S, Masperone L, Narkiewicz J, et al. alpha-synuclein amyloids hijack prion protein to gain cell entry, facilitate cell-to-cell spreading and block prion replication. Sci Rep. 2017 Aug 30;7(1):10050.
  • De Cecco E, Celauro L, Vanni S, et al. The uptake of tau amyloid fibrils is facilitated by the cellular prion protein and hampers prion propagation in cultured cells. J Neurochem. 2020 Dec;155(5):577–591.
  • Scialo C, Celauro L, Zattoni M, et al. The cellular prion protein increases the uptake and toxicity of TDP-43 fibrils. Viruses. 2021 Aug 17;13(8):1625.
  • Gasperini L, Meneghetti E, Pastore B, et al. Prion protein and copper cooperatively protect neurons by modulating NMDA receptor through S-nitrosylation. Antioxid Redox Signal. 2015 Mar 20;22(9):772–784.
  • Amin L, Nguyen XT, Rolle IG, et al. Characterization of prion protein function by focal neurite stimulation. J Cell Sci. 2016 Oct 15;129(20):3878–3891.
  • Caiati MD, Safiulina VF, Fattorini G, et al. PrPC controls via protein kinase A the direction of synaptic plasticity in the immature hippocampus. J Neurosci. 2013 Feb 13;33(7):2973–2983.
  • Biljan I, Ilc G, Giachin G, et al. NMR structural studies of human cellular prion proteins. Curr Top Med Chem. 2013;13(19):2407–2418.
  • Viles JH, Cohen FE, Prusiner SB, et al. Copper binding to the prion protein: structural implications of four identical cooperative binding sites. Proc Natl Acad Sci U S A. 1999 Mar 2;96(5):2042–2047.
  • Burns CS, Aronoff-Spencer E, Dunham CM, et al. Molecular features of the copper binding sites in the octarepeat domain of the prion protein. Biochemistry. 2002 Mar 26;41(12):3991–4001.
  • Millhauser GL. Copper binding in the prion protein. Acc Chem Res. 2004 Feb;37(2):79–85.
  • Burns CS, Aronoff-Spencer E, Legname G, et al. Copper coordination in the full-length, recombinant prion protein. Biochemistry. 2003 Jun 10;42(22):6794–6803.
  • Abskharon RN, Giachin G, Wohlkonig A, et al. Probing the N-terminal beta-sheet conversion in the crystal structure of the human prion protein bound to a nanobody. J Am Chem Soc. 2014 Jan 22;136(3):937–944.
  • Nakic N, Tran TH, Novokmet M, et al. Site-specific analysis of N-glycans from different sheep prion strains. PLoS Pathog. 2021 Feb;17(2):e1009232.
  • Brown DR, Qin K, Herms JW, et al. The cellular prion protein binds copper in vivo. Nature. 1997 Dec 18-25;390(6661):684–687.
  • Stockel J, Safar J, Wallace AC, et al. Prion protein selectively binds copper(II) ions. Biochemistry. 1998 May 19;37(20):7185–7193.
  • Aronoff-Spencer E, Burns CS, Avdievich NI, et al. Identification of the Cu2+ binding sites in the N-terminal domain of the prion protein by EPR and CD spectroscopy. Biochemistry. 2000 Nov 14;39(45):13760–13771.
  • Posadas Y, Lopez-Guerrero VE, Segovia J, et al. Dissecting the copper bioinorganic chemistry of the functional and pathological roles of the prion protein: relevance in Alzheimer’s disease and cancer. Curr Opin Chem Biol. 2022 Feb;66:102098.
  • Nguyen XTA, Tran TH, Cojoc D, et al. Copper binding regulates cellular prion protein function. Mol Neurobiol. 2019 Sep;56(9):6121–6133.
  • Sigurdsson EM, Brown DR, Alim MA, et al. Copper chelation delays the onset of prion disease. J Biol Chem. 2003 Nov 21;278(47):46199–46202.
  • Bocharova OV, Breydo L, Salnikov VV, et al. Copper(II) inhibits in vitro conversion of prion protein into amyloid fibrils. Biochemistry. 2005 May 10;44(18):6776–6787.
  • Delisle MB, Uro-Coste E, Gray F, et al. Fatal familial insomnia. Clin Exp Pathol. 1999;47(3–4):176–180.
  • Hokfelt T, Broberger C, Zhang X, et al. Neuropeptide Y: some viewpoints on a multifaceted peptide in the normal and diseased nervous system. Brain Res Brain Res Rev. 1998 May;26(2–3):154–166.
  • Norstrom EM, Mastrianni JA. The AGAAAAGA palindrome in PrP is required to generate a productive PrPSc-PrPC complex that leads to prion propagation. J Biol Chem. 2005 Jul 22;280(29):27236–27243.
  • Lau A, McDonald A, Daude N, et al. Octarepeat region flexibility impacts prion function, endoproteolysis and disease manifestation. EMBO Mol Med. 2015 Mar;7(3):339–356.
  • Eigenbrod S, Frick P, Bertsch U, et al. Substitutions of PrP N-terminal histidine residues modulate scrapie disease pathogenesis and incubation time in transgenic mice. PLoS One. 2017;12(12):e0188989.
  • Abskharon R, Wang F, Wohlkonig A, et al. Structural evidence for the critical role of the prion protein hydrophobic region in forming an infectious prion. PLoS Pathog. 2019 Dec;15(12):e1008139.
  • Legname G, Baskakov IV, Nguyen HO, et al. Synthetic mammalian prions. Science. 2004 Jul 30;305(5684):673–676.
  • Flechsig E, Shmerling D, Hegyi I, et al. Prion protein devoid of the octapeptide repeat region restores susceptibility to scrapie in PrP knockout mice. Neuron. 2000 Aug;27(2):399–408.
  • Shmerling D, Hegyi I, Fischer M, et al. Expression of amino-terminally truncated PrP in the mouse leading to ataxia and specific cerebellar lesions. Cell. 1998 Apr 17;93(2):203–214.
  • Ghetti B, Dlouhy SR, Giaccone G, et al. Gerstmann-Straussler-Scheinker disease and the Indiana kindred. Brain Pathol. 1995 Jan;5(1):61–75.
  • Stevens DJ, Walter ED, Rodriguez A, et al. Early onset prion disease from octarepeat expansion correlates with copper binding properties. PLoS Pathog. 2009 Apr;5(4):e1000390.
  • Ilc G, Giachin G, Jaremko M, et al. NMR structure of the human prion protein with the pathological Q212P mutation reveals unique structural features. PLoS One. 2010 Jul 22;5(7):e11715.
  • D’Angelo P, Della Longa S, Arcovito A, et al. Effects of the pathological Q212P mutation on human prion protein non-octarepeat copper-binding site. Biochemistry. 2012 Aug 7;51(31):6068–6079.
  • Schilling KM, Tao L, Wu B, et al. Both N-terminal and C-Terminal histidine residues of the prion protein are essential for copper coordination and neuroprotective self-regulation. J Mol Biol. 2020 Jul 24;432(16):4408–4425.
  • Slapsak U, Salzano G, Amin L, et al. The N terminus of the prion protein mediates functional interactions with the Neuronal Cell Adhesion Molecule (NCAM) fibronectin domain. J Biol Chem. 2016 Oct 14;291(42):21857–21868.
  • Giachin G, Mai PT, Tran TH, et al. The non-octarepeat copper binding site of the prion protein is a key regulator of prion conversion. Sci Rep. 2015 Oct 20;5:15253.
  • Salzano G, Giachin G, Legname G. Structural consequences of copper binding to the prion protein. Cells. 2019 Jul 25;8(8). DOI:10.3390/cells8080770.
  • Salzano G, Brennich M, Mancini G, et al. Deciphering copper coordination in the mammalian prion protein amyloidogenic domain. Biophys J. 2020 Feb 4;118(3):676–687.
  • Wang LQ, Zhao K, Yuan HY, et al. Genetic prion disease-related mutation E196K displays a novel amyloid fibril structure revealed by cryo-EM. Sci Adv. 2021 Sep 10;7(37):eabg9676.
  • Kraus A, Hoyt F, Schwartz CL, et al. High-resolution structure and strain comparison of infectious mammalian prions. Mol Cell. 2021 Nov 4;81(21):4540–4551 e6.
  • Hoyt F, Standke HG, Artikis E, et al. Cryo-EM structure of anchorless RML prion reveals variations in shared motifs between distinct strains. Nat Commun. 2022 Jul 13;13(1):4005.
  • Manka SW, Zhang W, Wenborn A, et al. 2.7 A cryo-EM structure of ex vivo RML prion fibrils. Nat Commun. 2022 Jul 13;13(1):4004.
  • Hallinan GI, Ozcan KA, Hoq MR, et al. Cryo-EM structures of prion protein filaments from Gerstmann-Straussler-scheinker disease. Acta Neuropathol. 2022 Sep;144(3):509–520.
  • Caughey B, Standke HG, Artikis E, et al. Pathogenic prion structures at high resolution. PLoS Pathog. 2022 Jun;18(6):e1010594.
  • Artikis E, Kraus A, Caughey B. Structural biology of ex vivo mammalian prions. J Biol Chem. 2022 Jun 23;298(8):102181.