Next Article in Journal
An Insight into miR-1290: An Oncogenic miRNA with Diagnostic Potential
Previous Article in Journal
Increased Risky Choice and Reduced CHRNB2 Expression in Adult Male Rats Exposed to Nicotine Vapor
Previous Article in Special Issue
Cofilin Signaling in the CNS Physiology and Neurodegeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Prion Protein: The Molecule of Many Forms and Faces

by
Valerija Kovač
and
Vladka Čurin Šerbec
*
Centre for Immunology and Development, Blood Transfusion Centre of Slovenia, Šlajmerjeva 6, SI-1000 Ljubljana, Slovenia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(3), 1232; https://doi.org/10.3390/ijms23031232
Submission received: 2 December 2021 / Revised: 10 January 2022 / Accepted: 21 January 2022 / Published: 22 January 2022
(This article belongs to the Special Issue Axonopathy in Neurodegenerative Diseases)

Abstract

:
Cellular prion protein (PrPC) is a glycosylphosphatidylinositol (GPI)-anchored protein most abundantly found in the outer membrane of neurons. Due to structural characteristics (a flexible tail and structured core), PrPC interacts with a wide range of partners. Although PrPC has been proposed to be involved in many physiological functions, only peripheral nerve myelination homeostasis has been confirmed as a bona fide function thus far. PrPC misfolding causes prion diseases and PrPC has been shown to mediate β-rich oligomer-induced neurotoxicity in Alzheimer’s and Parkinson’s disease as well as neuroprotection in ischemia. Upon proteolytic cleavage, PrPC is transformed into released and attached forms of PrP that can, depending on the contained structural characteristics of PrPC, display protective or toxic properties. In this review, we will outline prion protein and prion protein fragment properties as well as overview their involvement with interacting partners and signal pathways in myelination, neuroprotection and neurodegenerative diseases.

1. Introduction

Prion protein (PrP) is a highly conserved ubiquitous glycoprotein. It exists in two forms; the normal or cellular isoform, PrPC, and the disease-associated infectious isoform or scrapie PrP, PrPSc. The pathological role of PrPSc has been extensively studied in prion disease and has been reviewed in several papers [1,2,3]. PrPC is expressed in a variety of different organs and tissues with high expression levels in the central and peripheral nervous systems. It is abundantly present on the cell surface of neurons [4,5,6] and has been shown to be involved in many physiological mechanisms. The function of the protein remains to be elucidated; nevertheless, intensive studies link PrPC to myelin homeostasis [7], neuroprotection [8,9], the circadian rhythm [10,11], metal ion homeostasis [12,13], mitochondrial homeostasis [14] and intercellular signaling [6,15,16]. In neurons, PrPC is present in the presynaptic and postsynaptic compartments of axon terminals where it is involved in anterograde and retrograde axonal transport [17,18,19,20]. PrPC is cleaved at the cell membrane by proteases, forming released and attached forms. In recent years, prion protein and prion protein released forms have received attention in correlation with neuroprotection in neurodegenerative diseases. In this review, we present prion protein and prion protein released forms, summarize their involvement in myelination, neuroprotection and neurodegenerative diseases and discuss the most recent discoveries in this field.

2. Prion Protein

Mature human PrPC is composed of a flexible unstructured N-terminal domain (amino acid residues 23–120) and a structured C-terminal domain (amino acid residues 121–231). It is anchored to the cell membrane with a glycosylphosphatidylinositol (GPI) anchor [21,22]. The flexible N-terminal domain contains an octarepeat region whereas the structured domain consists of three α-helices, two β-sheets, a disulfide bond connecting cysteines 179 and 214 and two N-glycans on amino acid residues 181 and 197 [23,24] (Figure 1).
PrPC can transform into a β-sheet-rich isoform PrPSc, which is prone to autocatalytic conversion and aggregation into insoluble aggregates [22,25,26]. An abnormal accumulation of the pathologic protein in the brain can cause the development of transmissible spongiform encephalopathies (TSEs), also known as prion diseases. Prion diseases include Creutzfeldt–Jakob disease (CJD), Gerstmann–Sträussler–Scheinker syndrome (GSS), fatal familial insomnia (FFI) and kuru in humans, bovine spongiform encephalopathy in cattle, scrapie in goats and sheep and chronic wasting disease in cervids. All prion diseases are rare fatal neurodegenerative disorders. The clinical and neuropathological features of prion diseases in humans are similar to those of Alzheimer’s disease (AD) such as rapid memory loss and loss of brain function as well as dementia, spongiform deformation of the brain, personality changes and difficulties with movement [15,27]. Although prion diseases occur due to the accumulation of toxic PrPSc aggregates in the brain, the mechanism that underlies the conversion of PrPC to PrPSc and the development of prion disease remains an unknown. Apart from being a substrate for the development of prion diseases, PrPC can serve as a receptor for cytotoxic amyloid-β (Aβ) oligomers [20,28] and toxic soluble aggregates of tau protein in AD and other tauopathies [29,30]. There are also opposing studies on PrPC binding of α-synuclein (α-syn) oligomers in Parkinson’s disease (PD) and other synucleinopathies, opening the debate on the role of PrPC in toxicity of α-synuclein [30,31,32,33].

3. Prion Protein Fragments

PrPC can undergo four posttranslational cleavages, forming PrP fragments (Figure 1). The α-cleavage and β-cleavage occur within the unstructured N-terminal domain whereas the γ-cleavage and PrP shedding occur within the structured C-terminal domain. Apart from the mentioned cleavages, PrPC has been cleaved under experimental conditions with phospholipase C, which cleaved PrPC within the GPI anchor [34,35]. The site of cleavage, length of fragment and membrane attachment allow fragments to take part in various mechanisms.

3.1. α-Cleavage

The α-cleavage is the most studied cleavage of PrPC. It occurs under physiological conditions in the central hydrophobic region of mature PrPC (amino acid residues 105–120 in human sequence 111/112) [36,37,38] (Figure 1). The cleavage releases an ~11 kDa fragment N1 whereas the ~18 kDa part C1 remains attached to the cell membrane by the GPI anchor [36,39]. For now, there is no unique enzyme responsible for the α-cleavage [24,40]. Although cleavage sites have been determined with respect to species, the α-cleavage is tolerant to sequence variation in this region as long as its hydrophobicity remains preserved [38]. Studies have shown that α-cleavage in the human brain, mouse models and neuronal cultures occurs in the presence of enzymes ADAM10 and ADAM17 [41,42,43]. ADAM10 contributes to a constitutive N1 production whereas ADAM17 mainly participates in N1 formation upon stimulation [44,45]. ADAM8 has also been shown to cleave PrPC to form N1 and C1 in muscles [46]. A role of ADAM8, ADAM10 and ADAM17 in the α-cleavage has also been supported in a biophysical study [47]. Fragment N1 has a relatively low stability; nevertheless, it was found to be present in body fluids, tissue homogenates or cell culture supernatants [39,48,49]. The cleavage was initially thought to take place in acidic endosomal compartments [50,51] but later studies demonstrated that the α-cleavage occurs during the vesicular trafficking of PrPC along the secretory pathway [52,53]. The α-cleavage uses PrPC as a substrate, leading to its reduction of the cell surface. As PrPC is also a substrate for prion replication and a key mediator of toxicity in prion diseases, AD and other neurodegenerative diseases, the cleavage has a positive biological effect. The flexible N-terminal part of PrPC is essential for the interaction of the protein with the binding partners that regulate PrPC uptake in trafficking [54,55]. Lacking N1, C1 forms complexes on the cell membrane [56] and is more stable and persistent at the cell surface than PrPC [50]. Fragment C1 can be cleaved at the cell surface and released into the extracellular space [57]. C1 was found to inhibit prion replication in mice [58,59] whereas fragment N1 is neuroprotective [60,61]; the absence of the α-cleavage is toxic for both cells and mice [47,62].

3.2. β-Cleavage

The β-cleavage takes place at the end of the octapeptide repeat region N-terminal of the α-cleavage site. The β-cleavage is mostly observed under pathological conditions and is similar to the α-cleavage. It seems to act protectively. It takes place around amino acid residue 90, forming fragment N2 (~9 kDa) and fragment C2 (~20 kDa) [36,37,48,63] (Figure 1). The β-cleavage of PrPC is mediated by reactive oxygen species (ROS) [37,63,64,65,66]. By removing ROS, the cleavage protects cells from oxidative stress [65]. Apart from ROS, the β-cleavage is induced by calpains [67], lysosomal proteases [68,69] or even ADAM8 [47]. Proteinase K cleaves the protease-resistant core of PrPSc (PrP27–30) near position 90, creating a fragment with a length similar to C2. Similar to fragment C1, fragment C2 can also be shed from the cell surface [70]. The formation of such a fragment indicates that proteases involved in the β-cleavage could also be involved in the cellular attempts to break down PrPSc [71,72].

3.3. γ-Cleavage

The most recently discovered protease cleavage of PrPC is the γ-cleavage. The cleavage site in PrPC remains to be determined but the sizes of the released fragment N3 (~20 kDa) and GPI-anchored fragment C3 (~5 kDa) suggest that protein cleavage occurs in the region between amino acid residues 170 and 200 [73,74] (Figure 1). Studies indicate that the γ-cleavage occurs late in the secretory pathway on an unglycosylated protein in the presence of members of the matrix metalloproteases (MMP) family [73]. The reason the γ-cleavage occurs only on unglycosylated PrPC is proposed to be due to the steric hindrance of proteases by glycans in the proximity of the proposed cleavage site [40,75]. The γ-cleavage has been found to exist in different species, tissues and cell culture models. The determination of its role requires further study although an indication of increased amounts of fragment C3 in a CJD brain may lead to a possible pathogenic significance [73].

3.4. Shedding of Prion Protein

There is also an important cleavage of PrP in proximity to the C-terminus. The cleavage sheds PrP into the extracellular space, leaving a small number of amino acid residues on the cell surface. The cleavage was described in early research [35,39,76,77] but has received more attention in recent years due to the involvement of shed PrP in diseases [40,63,78,79,80,81,82,83]. Similar to the α-cleavage, the shedding of PrP occurs in the presence of enzymes from the ADAM family. In vitro and in vivo experiments suggest that ADAM9 and ADAM10 are involved in the process of cleavage and the shedding of PrP [47,84,85,86] where ADAM10 is the primary sheddase for PrP and ADAM9 is the modulator of ADAM10 activity [24]. Shed PrP was first determined in hamsters. In the prion-infected brain of hamsters, shed PrP represented approximately 15% of the PrPSc molecules [76]. A further analysis showed that ADAM10 cleaved shed PrP between Gly228 and Arg229 and formed shed PrP that terminated at Gly228 [84]. An analysis exploring the cleavage site profile of ADAM10 revealed that cleavage is not induced by a unique sequence [87]. Consequently, the ADAM10 protease can produce variants of shed PrP depending on the protein sequence and conformation. Jansen and coworkers described the existence of unanchored PrP forms ending with Tyr225 and Tyr226 in patients with prion disease [88]. The authors characterized two patients with prion disease who carried stop mutations at positions Y226X and Q227X and expressed the respective forms. Using a monoclonal antibody V5B2 [89] that specifically binds to a fragment of PrP ending with Tyr226, we concurrently described the existence of a free form of PrP named PrP226* [90,91,92,93,94]. The distribution of PrP226* in the human brain has been associated with the distribution of PrPSc [90,94]. Due to the existence of more than one shed form, we hypothesized that the proteolytic site in the human sequence is not exclusively located between amino acid residues 228 and 229 but is located in the proximity of the C-terminus [95] (Figure 1). Recently, Linsenmeier et al. published a comprehensive study on the mechanism stimulating PrPC proteolytic shedding [81]. Using animal models and controls, they showed that PrP shedding negatively correlates with prion conversion and that shed PrP is abundantly present in amyloid plaques. They also studied the influence of the binding of PrP-directed antibodies to PrPC in relation to shedding propensity. The binding of whole anti-PrP antibodies to the C-terminal structured domain of PrPC or single-chain antibody derivatives, directed towards repetitive epitopes within the octarepeat region of the N-terminal domain stimulated shedding, when the binding of whole anti-PrP antibodies to the octarepeat region of the N-terminal domain locked the N-terminal domain structure and evoked PrPC surface clustering, endocytosis and degradation in lysosomes [81].

4. Prion Protein and Myelination

PrPC is abundantly expressed in the central nervous system (CNS) and in the peripheral nervous system [4,5]. Studies in primate brains, rodent brains and transgenic mice showed that it is enriched along axons and in presynaptic terminals where it is involved in anterograde and retrograde axonal transport [4,17,18,96,97,98]. Deletions in the PrPC α-cleavage region showed severe demyelination in both the spinal cord and cerebellar white matter in vivo [99,100] Later, it was confirmed that axonal PrPC and its α-cleavage are necessary for pro-myelination in the peripheral nervous system [101]. Using a co-isogenic PrP-knockout mice model, Kuffer et al. discovered that axonal PrPC promotes myelin maintenance in trans via binding to the adhesion G-protein-coupled receptor Adgrg6 on Schwann cells with an N-terminal flexible tail [7]. They also confirmed that mice lacking PrPC developed chronic demyelinating neuropathy, which suggests that myelination homeostasis in the peripheral nervous system is a bona fide physiological function of PrPC [7]. Myelin maintenance was found to be regulated through the binding of an N-terminal released fragment of PrPC (presumably N1 or shed PrP) to Adgrg6 on Schwann cells. The interaction activated Adgrg6, increased the cellular levels of cAMP and triggered a signaling cascade that promoted myelination [7]. The regulation of peripheral myelin maintenance by PrPC was confirmed in five different PrP-knockout mouse model strains that developed late-onset peripheral neuropathy [101,102,103]. Recently, there was an attempt to develop a treatment for peripheral demyelinating diseases based on binding between the N-terminal domain of PrPC and Adgrg6 [104]. In this study, they constructed an immunoadhesin molecule consisting of two flexible N-terminal domains of PrPC linked to a crystallizable fragment (Fc) of immunoglobulin G1 (FT2Fc) [104]. The molecule showed favorable pharmacokinetic properties and showed potential in vitro but failed to have a therapeutic effect on the early molecular signs of demyelination in PrP-knockout mice [104]. PrPC was also studied in connection to peripheral myelin development and regeneration after nerve injuries [105]. As PrP was found to be dispensable in this mechanism, it could be presumed that PrP has no major role in the peripheral nerve repair process or its absence might be compensated by other ligands [105].
Myelination and other physiological roles of PrPC have been intensively studied on animal models with a knocked-out or knocked-down PrP gene expression. Studies have shown limited negative effects in mice [102,106,107,108,109], cattle [110] and goats [68,111,112] whereas studies on PrP-knockout mice or goats showed defects in the nervous system and sensitivity to oxidative stress [6,101,111,113]. Several PrP-knockout mice models were generated with a mixed background [106,109,114,115,116]. As the studies are not reproducible among models, this might raise the question of whether any observed phenotypes were actually due to polymorphisms in genes flanking Prnp or the result of PrPC absence. To avoid this issue, it would be advisable to repeat key experiments using co-isogenic PrP-knockout mice.
Although the role of PrPC in the CNS needs to be elucidated, PrPC and PrPC released fragments are indispensable in peripheral nerve myelin homeostasis but they may be dispensable in nerve recovery.

5. Prion Protein and Ischemic Strokes

In the previous section, we observed that knockout animals are more vulnerable to oxidative stress. Studies support the idea that PrPC acts as an antioxidant by regulating glutathione reductase activity [117,118] and by regulating superoxide dismutase (SOD) through ion binding [119,120,121,122,123]. PrP-knockout mice showed a reduced protection against ROS whereas prion-infected mice showed increased levels of oxidative stress, most likely as a consequence of a PrPC loss of function [124,125,126]. Under oxidative stress conditions, PrP mRNA levels increase, which implies that oxidative stress upregulates PrPC expression [127]. Ischemic stroke is a condition where the loss of blood flow in a brain area causes hypoxic conditions and brain damage [128]. PrP-knockout animal models subject to ischemia showed intensive ischemic damage and a reduced chance of regeneration whereas the possibility of PrPC synthesis resulted in PrPC overexpression and decreased ischemic damage [127]. Studies on ischemic strokes have indicated that PrPC overexpression can reduce the lesion size compared with wild-type mice, ascribing PrPC a protective role in ischemia damage [129,130,131,132,133,134,135]. After an ischemic insult, PrPC is associated with neuroprotective and regenerative processes by interacting with various cytosolic and transmembrane signal proteins. Among others, PrPC has been associated with the upregulation of extracellular signal-regulated kinase (ERK1/2) [133,136,137], activation of the phosphatidylinositol 3-kinase/protein kinase B/Akt (PI3K/Akt) pathway [138,139,140,141,142], modulation of N-methyl-D-aspartate (NMDA) receptor-mediated toxicity [143], activation of the cAMP-dependent protein kinase A (PKA) pathway [144,145,146] and interaction with stress-inducible protein 1 (STI1) [146], all resulting in neuron survival, neurite outgrowth and neuroprotection.
PrPC is a receptor of Fyn kinase, a member of the Src family of tyrosine kinases (SFKs) [146]. Through Fyn kinase activation, PrPC mediates oligomer-induced toxicity in neurodegenerative diseases [147,148,149,150] and promotes neurite outgrowth by the phosphorylation of the GluN2A domain of the neuronal cell adhesion molecule (NCAM) [151]. Fyn kinase and other members of the SFK family are involved in ischemic damage [152,153,154,155]. The inhibition of SFKs in a global ischemia model and the inhibition of the Fyn-mediated phosphorylation of GluN2A in a model of neonatal HII resulted in an increased neuronal survival [156,157,158] whereas the overexpression of Fyn in the model of neonatal HII led to increased brain damage [159]. The inhibition of SFKs in a mouse model of an ischemia also resulted in a decreased ischemic volume and improved cerebral function after provocation [155]. As this effect was not seen in Fyn-knockout mice, we suspect that ligands other than Fyn kinase may also affect ischemia insult recovery [155].
PrPC fragments were also shown to be involved in ischemic stroke. Fragments N1 and N2 were shown to act protectively under cellular stress [160,161,162] and modulate the quiescence of neural stem cells in adult neurogenesis upon stroke [163] whereas PrPC fragments C1 and C2 were involved in regulating p53-dependent apoptosis and cell survival [164]. Fragment C1 was found to be enriched in small EVs (sEVs) where it acted similarly to viral surface proteins [165,166]. Due to this, it may affect the intercellular information exchange between sEVs and their target cells as well as contributing to their uptake [63]. Brenna et al. studied the similarities between the cellular uptake of brain-derived sEVs from PrP-knockout mice and wild-type mice after a stroke [128]. They showed that sEVs lacking PrP were taken up significantly faster with a greater efficiency and were more easily sorted into lysosomes than sEVs containing PrP and fragment C1 [128]. Fragment N1 was also found to be involved in regulating the interactions between microglia and other brain cells. A recent in vitro study on a mixed neuronal lineage and microglia co-culture system showed that fragment N1 stimulated a change in the cell morphology and metabolism and induced Cxcl10 secretion [167]. Furthermore, fragment N1 was shown to influence microglia to change the membrane composition to a higher GM1 content at the interaction sites with the surrounding cells in a co-culture yet only upon direct cell-to-cell contact [167]. Fragment N1 was also proposed to protect neurons against staurosporine-induced Caspase-3 activation in an ischemic model of the rat retina [60]. These results are supported by in vitro studies where the expression of PrPC was protective against staurosporine or anisomycin-induced apoptosis [144,146]. Fragment N1 is also related to neuroprotection in neurodegenerative diseases, which is discussed in more detail in the next section. In the presence of anchored PrPC, recombinant PrP (recPrP) can induce ERK1/2 and Akt signaling on mesenchymal stem cells that may support neuronal differentiation [168], promote neurite outgrowth and facilitate axonal growth cone guidance [169]. Recently, it was reported that recPrP promotes neurite outgrowth and Schwann cell migration through the ERK1/2 pathway [170]. The activation involved NMDA receptors, low density lipoprotein receptor-related protein-1 (LRP1), SFKs and Trk receptors; it seemed to take place independently of anchored PrPC [170]. In this mechanism, SFKs played a critical role in recPrP-initiated cell signaling by activating Trk receptors, which are upstream of ERK1/2 [170,171]. Although recPrP lacks glycosylation, it might be considered to be a suitable analog of shed PrP.
Prion protein and prion protein fragments are linked with intercellular communication and signaling, oxidative stress and neuroprotection and present an attractive target for the treatment and regulation of these mechanisms. Nevertheless, further studies should be conducted to confirm the effects of these molecules in the mentioned mechanisms.

6. Prion Protein and Neurodegeneration

Neurodegeneration is the progressive loss of the structure or function of neurons, which may ultimately involve cell death. On the molecular level, neurodegeneration is connected to accumulation of misfolded proteins. Accumulation of protein aggregates causes mitochondria dysfunction, induces oxidative stress and ultimately causes chronic inflammation. Neurodegeneration occurs in diseases such as prion disease, PD and AD due to the aggregation of PrPSc [26,172,173], α-syn [174,175,176,177] and Aβ isoforms [178,179] and tau protein [180,181,182,183], respectively. Prion protein or prion protein fragments have been found to interact with aggregating agents in different neurodegenerative diseases but their roles depend on the studied conditions [24,81,184,185].
It has been reported that PrPC binds a wide range of β-sheet-rich oligomers associated with neurodegenerative diseases [148,149,150]. PrPC engages metabotropic glutamate receptor 5 (mGluR5) and mediates oligomer-induced toxicity through Fyn kinase [175,186,187,188]. Activated Fyn kinase can phosphorylate the GluN2A and GluN2B subunits of NMDA receptors, which are then hyperactivated and cause calcium influx and cell death [20,189]. It has also been shown that PrPC can activate Fyn kinase-mediated Aβ oligomer toxicity by an interaction with LRP1 [190]. A recent study in this field suggested that, apart from LRP1, this process includes activated a2-macroglobulin and tissue-type plasminogen activator [191]. Studies have implied that binding between soluble protein aggregates and PrPC causes neurotoxicity and inhibits long-term potentiation (LTP) [30,192]. Opposing studies have also been published that report no significant effect of PrPC levels on Aβ-induced LTP in PrP-knockout mice [193], cell ablation or PrP overexpression [194]. The reasons for these discrepancies are unclear but they could be due to the use of different model systems and toxic or nontoxic species [195].
Aβ oligomers bind to PrPC at two binding sites within the flexible N-terminal part of PrPC, between amino acid residues 23–27 and 92–110 [192,195,196]. Apart from Aβ oligomers, PrPC has been reported to be a receptor for α-syn oligomers and tau aggregates. Similar to Aβ oligomers, anchored PrPC binds small soluble aggregates or shorter fibrils of α-syn oligomers or tau aggregates within the flexible N-terminal part [30,175,185,197,198,199]. PrPC has also been shown to uptake recombinant α-syn fibrils. A model system lacking PrPC showed a lower uptake of α-syn and α-syn fibrils in comparison with controls [177,185,197], resulting in less α-syn aggregation, astroglial activation and loss of dopaminergic neurons in the brains of PrP-knockout mice [185]. Furthermore, PrP-knockout mice did not exhibit α-syn-induced LTP impairment whereas treatment with an anti-PrP antibody prevented α-syn-induced LTP defects in a model of PD [175]. Although the mentioned studies support a PrPC and α-syn oligomer interplay, La Vitola et al. showed that PrPC was not mandatory for the mediation of α-syn oligomer detrimental effects in vitro or in vivo [33]. Although the discrepancy could not be explained in the study, it could also occur due to the use of a different protocol of soluble aggregate preparation or the use of different model systems. Anchored PrPC was also shown to bind tau aggregates and seemed to facilitate their uptake [30,198,200]. Absence of PrPC or pretreatment with anti-PrP blocking antibodies was shown to decrease the uptake of recombinant tau aggregates and abolish tau aggregate-induced toxicity [30,198,200].
Studies regarding recombinant PrP fragment N1 in neurodegenerative diseases have shown that these molecules can bind toxic Aβ oligomers at regions between amino acid residues 23–31 and 95–105. Fragment N1 neutralizes toxic Aβ oligomers by seizing them in the extracellular space and reduces oligomer-induced toxicity [61,195,201,202,203,204]. The protective effects of fragment N1 have also been observed in vivo in mice exposed to acute Aβ-induced toxicity [203]. Beland and coworkers observed increases in the α-cleavage of PrPC in the brains of AD patients [205]. As the N1 fragment abundantly binds Aβ oligomers, it may be indicated that the cleavage acts protectively in the development of diseases [205] whereas the inhibition of N1 production promotes AD progression [42].
PrP shedding reduces the level of cell-anchored PrPC [78]. This results in a decreased level of the substrate for prion replication and a decreased level of the receptor for toxic oligomers [85,206]. Similar to fragment N1, shed PrP is also believed to be protective in prion diseases and other neurodegenerative diseases [40,79,81]. As mentioned in the previous section, recPrP is similar to shed PrP. Although it lacks glycans, recPrP may be used as a model to predict the role of shed PrP in diseases. RecPrP was found to increase the development of synapses and neurite outgrowth in the presence of anchored PrPC [170,207]. Similar to fragment N1, recPrP also inhibited Aβ oligomer formation and neutralized Aβ oligomer toxicity in an AD model [203]. In vitro studies using recPrP and its derivatives showed that both the N-terminal and C-terminal domains of PrP are required for an efficient inhibition of Aβ fibril elongation [202,208] and support the protective role of shed PrP in the inhibition of Aβ fibril formation. RecPrP was also shown to bind tau aggregates and α-syn oligomers and may neutralize their toxicity [30]. Although PrPC shedding acts protectively, enhanced PrPC shedding could lead to negative biological activity such as inflammation in the CNS [83,209]. Jarosz-Griffiths et al. [82] recently reported on the protective role of PrP shedding. The authors reported that siRNA-mediated ADAM10 knockdown reduced PrPC shedding and increased Aβ oligomer binding whereas acitretin promoted PrPC shedding and decreased Aβ oligomer binding in the neuroblastoma cells and in human-induced pluripotent stem cells [82].
In a recent paper by Linsenmeier et al., researchers evaluated the role of shed PrP in different models [81]. Using a polyclonal antibody sPrPG228 that specifically recognized murine PrP ending with G228 [210] they showed that in prion-diseased mice, shed PrP colocalized with PrPSc in amyloid plaques. Similar to the model of prion disease, shed PrP was also distributed to Aβ deposits in the brains of 5xFAD mice where it was found bound to Aβ oligomers and seen in the center of many amyloid plaques. Due to the knowledge in this field thus far, the authors proposed that physiologically shed PrP may act protectively in prion diseases and AD by blocking toxic oligomers and/or by precipitating them into less toxic deposits [81,211].
RecPrP and N1 may also inhibit Aβ oligomerization, neutralize cytotoxicity of preexisting Aβ oligomers, prevent the binding of oligomers with cell surface PrPC and rescue the Aβ-induced impairment of LTP [212]. As recPrP and N1 both contain proposed binding sites of protein oligomers, both molecules were reported to also bind α-syn oligomers as well as mediate the co-clustering of α-syn oligomers and AD-associated amyloid-β oligomers [199].
PrPC is enriched in extracellular vesicles (EVs) [128,213,214]. Little is known regarding the physiological functions of PrPC in EVs. Several studies have suggested that PrPC in EVs protect cells against Aβ toxicity [214,215,216,217]. The mechanism behind the neutralization of toxic Aβ oligomers by EVs is not known; nevertheless, it is presumed that it is similar to the recPrP or N1-mediated process. It has been proposed that exosomal PrPC catches Aβ oligomers at the N-terminal PrP region (amino acid residues 23–31 and 95–105) [203], neutralizes the oligomers, promotes the formation of Aβ fibrils and upregulates internalization and degradation of the aggregates by microglia [214,215,216,217]. As recPrP and anchored PrPC have been shown to bind tau and α-syn oligomers [30], exosomal PrPs are expected to act in the same manner. By binding free toxic tau or α-syn oligomers in the extracellular space, exosomal PrPs prevent toxic oligomer binding to anchored PrPC and inhibit toxic signaling in the CNS of patients with diseases. Exosomes associated with PrPSc have been shown to be infectious and pose a danger of spreading prion disease [218,219,220,221,222]. Although there is no direct study yet, exosomal PrPC might also induce CNS inflammation. More work needs to be undertaken to examine other biological activities that exosomal PrPC may possess.
On the basis of the determined oligomer binding domains, researchers have designed potential treatment strategies for AD based on synthetic peptides [204,223] and functional Aβ oligomer-binding compounds [149]. The designed synthetic peptides have been shown to reduce the initial rate of Aβ fibrillization, inhibit the aggregation pathway of Aβ by reducing Aβ oligomer uptake and protect cultured hippocampal neurons from the oligomer-induced retraction of neurites and loss of cell membrane integrity [204] whereas D-peptide RD2D3 has been shown to be successful in interfering with the PrPC-Aβ oligomer assembly and has been proposed as a promising therapeutic agent in AD [223].

7. Conclusions

The reviewed studies support the fact that prion protein and/or prion protein fragments are involved in myelin homeostasis, ischemia and neurodegeneration where they may take on different roles (Figure 2). According to the current information, anchored PrP and/or released fragments (N1, shed PrP) interact with Adgrg6 to regulate peripheral nerve myelin homeostasis. Although there have been attempts to connect PrP to other Adgrg6-mediated processes, no direct involvement has been perceived. In strokes, the expression of PrP is upregulated. Anchored PrP takes part in mediating signaling pathways through transmembrane and cytosolic receptor proteins. Although further study is needed, the released forms may play decisive roles in neuroprotection and regeneration, including the regulation of interactions between microglia and brain cells and the promotion of neurogenesis. EVs and sEVs highly enriched in PrP fragments may be important delivery mechanisms in neuroprotection and neurodegeneration; further studies are needed to prove their roles. In neurodegenerative diseases, anchored PrP acts as a receptor for Aβ oligomers, α-syn oligomers and tau aggregates and may mediate oligomer-induced cytotoxicity. The point of interaction between the oligomer and PrP may be an attractive site for drug development but therapy may also include the regulation of other partners involved in this process. Arguing their protective role, released PrP fragments may bind toxic oligomers and enable their depletion. Supporting this role, shed PrP has been shown to bind PrPSc and Aβ oligomers in amyloid plaques, which may be less toxic than oligomers. To conclude, there are many indications suggesting that prion protein and prion protein fragments may have multiple (sometimes even intertwined) roles in strokes and neurodegeneration. To undoubtedly elucidate their role(s) in these processes, further studies are needed in these fields.

Author Contributions

V.K. conceptualized the manuscript scope and wrote the first draft; V.Č.Š. conceptualized the manuscript scope and critically reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The work was funded by Slovenian Research Agency (ARRS grant number P4-0176).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scheckel, C.; Aguzzi, A. Prions, prionoids and protein misfolding disorders. Nat. Rev. Genet. 2018, 19, 405–418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. O’Carroll, A.; Coyle, J.; Gambin, Y. Prions and Prion-like assemblies in neurodegeneration and immunity: The emergence of universal mechanisms across health and disease. Semin. Cell Dev. Biol. 2020, 99, 115–130. [Google Scholar] [CrossRef] [PubMed]
  3. Ritchie, D.L.; Barria, M.A. Prion Diseases: A Unique Transmissible Agent or a Model for Neurodegenerative Diseases? Biomolecules 2021, 11, 207. [Google Scholar] [CrossRef] [PubMed]
  4. Herms, J.; Tings, T.; Gall, S.; Madlung, A.; Giese, A.; Siebert, H.; Schurmann, P.; Windl, O.; Brose, N.; Kretzschmar, H. Evidence of presynaptic location and function of the prion protein. J. Neurosci. 1999, 19, 8866–8875. [Google Scholar] [CrossRef]
  5. Bendheim, P.E.; Brown, H.R.; Rudelli, R.D.; Scala, L.J.; Goller, N.L.; Wen, G.Y. Nearly ubiquitous tissue distribution of the scrapie agent precursor protein. Neurology 1992, 42, 149. [Google Scholar] [CrossRef]
  6. Wulf, M.-A.; Senatore, A.; Aguzzi, A. The biological function of the cellular prion protein: An update. BMC Biol. 2017, 15, 34. [Google Scholar] [CrossRef] [Green Version]
  7. Kuffer, A.; Lakkaraju, A.K.; Mogha, A.; Petersen, S.C.; Airich, K.; Doucerain, C.; Marpakwar, R.; Bakirci, P.; Senatore, A.; Monnard, A.; et al. The prion protein is an agonistic ligand of the G protein-coupled receptor Adgrg6. Nature 2016, 536, 464–468. [Google Scholar] [CrossRef]
  8. Carulla, P.; Bribián, A.; Rangel, A.; Gavín, R.; Ferrer, I.; Caelles, C. Neuroprotective role of PrPC against kainate-induced epileptic seizures and cell death depends on the modulation of JNK3 activation by GluR6/7–PSD-95 binding. Mol. Biol. Cell 2011, 22, 3041–3054. [Google Scholar] [CrossRef] [Green Version]
  9. Carulla, P.; Llorens, F.; Matamoros-Angles, A.; Aguilar-Calvo, P.; Espinosa, J.C.; Gavín, R. Involvement of PrPC in kainate-induced excitotoxicity in several mouse strains. Sci. Rep. 2015, 5, srep11971. [Google Scholar] [CrossRef] [Green Version]
  10. Collins, S.; McLean, C.A.; Masters, C.L. Gerstmann–Sträussler–Scheinker syndrome, fatal familial insomnia, and kuru: A review ofthese less common human transmissiblespongiform encephalopathies. J. Clin. Neurosci. 2001, 8, 387–397. [Google Scholar] [CrossRef]
  11. Dibner, C.; Schibler, U.; Albrecht, U. The Mammalian Circadian Timing System: Organization and Coordination of Central and Peripheral Clocks. Annu. Rev. Physiol. 2010, 72, 517–549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Cingaram, P.K.R.; Nyeste, A.; Dondapati, D.T.; Fodor, E.; Welker, E. Prion Protein Does Not Confer Resistance to Hippocampus-Derived Zpl Cells against the Toxic Effects of Cu2+, Mn2+, Zn2+ and Co2+ Not Supporting a General Protective Role for PrP in Transition Metal Induced Toxicity. PLoS ONE 2015, 10, e0139219. [Google Scholar] [CrossRef] [PubMed]
  13. Gasperini, L.; Meneghetti, E.; Pastore, B.; Benetti, F.; Legname, G. Prion protein and copper cooperatively protect neurons by modulating NMDA receptor through S-nitrosylation. Antioxid. Redox Signal. 2015, 22, 772–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Faris, R.; Moore, R.A.; Ward, A.; Race, B.; Dorward, D.W.; Hollister, J.R.; Fischer, E.R.; Priola, S.A. Cellular prion protein is present in mitochondria of healthy mice. Sci. Rep. 2017, 7, 41556. [Google Scholar] [CrossRef] [PubMed]
  15. Harris, D.A. Cellular biology of prion diseases. Clin. Microbiol. Rev. 1999, 12, 429–444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Slapšak, U.; Salzano, G.; Amin, L.; Abskharon, R.N.; Ilc, G.; Zupančič, B.; Biljan, I.; Plavec, J.; Giachin, G.; Legname, G. The N terminus of the prion protein mediates functional interactions with the neuronal cell adhesion molecule (NCAM) fibronectin domain. J. Biol. Chem. 2016, 291, 21857–21868. [Google Scholar] [CrossRef] [Green Version]
  17. Borchelt, D.R.; Koliatsos, V.E.; Guarnieri, M.; Pardo, C.A.; Sisodia, S.S.; Price, D.L. Rapid anterograde axonal transport of the cellular prion glycoprotein in the peripheral and central nervous systems. J. Biol. Chem. 1994, 269, 14711–14714. [Google Scholar] [CrossRef]
  18. Moya, K.L.; Hässig, R.; Créminon, C.; Laffont, I.; Giamberardino, L. Enhanced detection and retrograde axonal transport of PrPc in peripheral nerve: Cellular prion protein in peripheral nerve. J. Neurochem. 2003, 88, 155–160. [Google Scholar] [CrossRef]
  19. Haeberle, A.M.; Ribaut-Barassin, C.; Bombarde, G.; Mariani, J.; Hunsmann, G.; Grassi, J. Synaptic prion protein immuno-reactivity in the rodent cerebellum. Microsc. Res. Tech. 2000, 50, 66–75. [Google Scholar] [CrossRef]
  20. Um, J.W.; Nygaard, H.B.; Heiss, J.K.; Kostylev, M.A.; Stagi, M.; Vortmeyer, A. Alzheimer amyloid-β oligomer bound to postsynaptic prion protein activates Fyn to impair neurons. Nat. Neurosci. 2012, 15, 1227–1235. [Google Scholar] [CrossRef] [Green Version]
  21. Abskharon, R.; Wang, F.; Wohlkonig, A.; Ruan, J.; Soror, S.; Giachin, G.; Pardon, E.; Zou, W.; Legname, G.; Ma, J.; et al. Structural evidence for the critical role of the prion protein hydrophobic region in forming an infectious prion. PLoS Pathog. 2019, 15, e1008139. [Google Scholar] [CrossRef] [PubMed]
  22. Prusiner, S.B. Prions. Proc. Natl. Acad. Sci. USA 1998, 95, 13363–13383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zahn, R.; Liu, A.; Luhrs, T.; Riek, R.; von Schroetter, C.; Lopez Garcia, F.; Billeter, M.; Calzolai, L.; Wider, G.; Wuthrich, K. NMR solution structure of the human prion protein. Proc. Natl. Acad. Sci. USA 2000, 97, 145–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Dexter, E.; Kong, Q. Neuroprotective effect and potential of cellular prion protein and its cleavage products for treatment of neurodegenerative disorders part I. A literature review. Expert Rev. Neurother. 2021, 21, 969–982. [Google Scholar] [CrossRef]
  25. Castle, A.R.; Gill, A.C. Physiological Functions of the Cellular Prion Protein. Front. Mol. Biosci. 2017, 4, 19. [Google Scholar] [CrossRef] [Green Version]
  26. Pan, K.M.; Baldwin, M.; Nguyen, J.; Gasset, M.; Serban, A.; Groth, D.; Mehlhorn, I.; Huang, Z.W.; Fletterick, R.J.; Cohen, F.E.; et al. Conversion of alpha-helices into beta-sheets features in the formation of the scrapie prion proteins. Proc. Natl. Acad. Sci. USA 1993, 90, 10962–10966. [Google Scholar] [CrossRef] [Green Version]
  27. Walker, L.C.; Jucker, M. Neurodegenerative diseases: Expanding the prion concept. Annu. Rev. Neurosci. 2015, 38, 87–103. [Google Scholar] [CrossRef] [Green Version]
  28. Amin, L.; Harris, D.A. Aβ receptors specifically recognize molecular features displayed by fibril ends and neurotoxic oligomers. Nat. Commun. 2021, 12, 3451. [Google Scholar] [CrossRef]
  29. Castillo-Carranza, D.L.; Gerson, J.E.; Sengupta, U.; Guerrero-Muñoz, M.J.; Lasagna-Reeves, C.A.; Kayed, R. Specific targeting of tau oligomers in Htau mice prevents cognitive impairment and tau toxicity following injection with brain-derived tau oligomeric seeds. J. Alzheimer’s Dis. JAD 2014, 40 (Suppl. S1), S97–S111. [Google Scholar] [CrossRef] [Green Version]
  30. Corbett, G.T.; Wang, Z.; Hong, W.; Colom-Cadena, M.; Rose, J.; Liao, M.; Asfaw, A.; Hall, T.C.; Ding, L.; DeSousa, A.; et al. PrP is a central player in toxicity mediated by soluble aggregates of neurodegeneration-causing proteins. Acta Neuropathol. 2020, 139, 503–526. [Google Scholar] [CrossRef] [Green Version]
  31. Salazar, S.V.; Strittmatter, S.M. Cellular prion protein as a receptor for amyloid-beta oligomers in Alzheimer’s disease. Biochem. Biophys. Res. Commun. 2017, 483, 1143–1147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Vascellari, S.; Manzin, A. Parkinson’s Disease: A Prionopathy? Int. J. Mol. Sci. 2021, 22, 8022. [Google Scholar] [CrossRef] [PubMed]
  33. La Vitola, P.; Beeg, M.; Balducci, C.; Santamaria, G.; Restelli, E.; Colombo, L.; Caldinelli, L.; Pollegioni, L.; Gobbi, M.; Chiesa, R.; et al. Cellular prion protein neither binds to alpha-synuclein oligomers nor mediates their detrimental effects. Brain J. Neurol. 2019, 142, 249–254. [Google Scholar] [CrossRef] [PubMed]
  34. Stahl, N.; Borchelt, D.R.; Prusiner, S.B. Differential release of cellular and scrapie prion proteins from cellular membranes by phosphatidylinositol-specific phospholipase C. Biochemistry 1990, 29, 5405–5412. [Google Scholar] [CrossRef]
  35. Borchelt, D.R.; Rogers, M.; Stahl, N.; Telling, G.; Prusiner, S.B. Release of the cellular prion protein from cultured cells after loss of its glycoinositol phospholipid anchor. Glycobiology 1993, 3, 319–329. [Google Scholar] [CrossRef]
  36. Chen, S.G.; Teplow, D.B.; Parchi, P.; Teller, J.K.; Gambetti, P.; Autilio-Gambetti, L. Truncated forms of the human prion protein in normal brain and in prion diseases. J. Biol. Chem. 1995, 270, 19173–19180. [Google Scholar] [CrossRef] [Green Version]
  37. Mange, A.; Beranger, F.; Peoc’h, K.; Onodera, T.; Frobert, Y.; Lehmann, S. Alpha- and beta-cleavages of the amino-terminus of the cellular prion protein. Biol. Cell 2004, 96, 125–132. [Google Scholar] [CrossRef]
  38. Oliveira-Martins, J.B.; Yusa, S.-I.; Calella, A.M.; Bridel, C.; Baumann, F.; Dametto, P.; Aguzzi, A. Unexpected Tolerance of α-Cleavage of the Prion Protein to Sequence Variations. PLoS ONE 2010, 5, e9107. [Google Scholar] [CrossRef] [Green Version]
  39. Harris, D.A.; Huber, M.T.; Dijken, P.; Shyng, S.L.; Chait, B.T.; Wang, R. Processing of a cellular prion protein: Identification of N-and C-terminal cleavage sites. Biochemistry 1993, 32, 1009–1016. [Google Scholar] [CrossRef]
  40. Linsenmeier, L.; Altmeppen, H.C.; Wetzel, S.; Mohammadi, B.; Saftig, P.; Glatzel, M. Diverse functions of the prion protein—Does proteolytic processing hold the key? Biochim. Biophys. Acta 2017, 1864, 2128–2137. [Google Scholar] [CrossRef]
  41. Laffont-Proust, I.; Faucheux, B.A.; Hässig, R.; Sazdovitch, V.; Simon, S.; Grassi, J.; Hauw, J.-J.; Moya, K.L.; Haïk, S. The N-terminal cleavage of cellular prion protein in the human brain. FEBS Lett. 2005, 579, 6333–6337. [Google Scholar] [CrossRef] [Green Version]
  42. Pietri, M.; Dakowski, C.; Hannaoui, S.; Alleaume-Butaux, A.; Hernandez-Rapp, J.; Ragagnin, A.; Mouillet-Richard, S.; Haik, S.; Bailly, Y.; Peyrin, J.-M.; et al. PDK1 decreases TACE-mediated α-secretase activity and promotes disease progression in prion and Alzheimer’s diseases. Nat. Med. 2013, 19, 1124–1131. [Google Scholar] [CrossRef] [PubMed]
  43. Alleaume-Butaux, A.; Nicot, S.; Pietri, M.; Baudry, A.; Dakowski, C.; Tixador, P.; Ardila-Osorio, H.; Haeberlé, A.-M.; Bailly, Y.; Peyrin, J.-M.; et al. Double-Edge Sword of Sustained ROCK Activation in Prion Diseases through Neuritogenesis Defects and Prion Accumulation. PLoS Pathog. 2015, 11, e1005073. [Google Scholar] [CrossRef] [PubMed]
  44. Vincent, B.; Paitel, E.; Frobert, Y.; Lehmann, S.; Grassi, J.; Checler, F. Phorbol ester-regulated cleavage of normal prion protein in HEK293 human cells and murine neurons. J. Biol. Chem. 2000, 275, 35612–35616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Vincent, B.; Paitel, E.; Saftig, P.; Frobert, Y.; Hartmann, D.; De Strooper, B.; Grassi, J.; Lopez-Perez, E.; Checler, F. The disintegrins ADAM10 and TACE contribute to the constitutive and phorbol ester-regulated normal cleavage of the cellular prion protein. J. Biol. Chem. 2001, 276, 37743–37746. [Google Scholar] [CrossRef] [PubMed]
  46. Liang, J.; Wang, W.; Sorensen, D.; Medina, S.; Ilchenko, S.; Kiselar, J.; Surewicz, W.K.; Booth, S.A.; Kong, Q. Cellular Prion Protein Regulates Its Own α-Cleavage through ADAM8 in Skeletal Muscle*. J. Biol. Chem. 2012, 287, 16510–16520. [Google Scholar] [CrossRef] [Green Version]
  47. McDonald, A.J.; Dibble, J.P.; Evans, E.G.B.; Millhauser, G.L. A New Paradigm for Enzymatic Control of α-Cleavage and β-Cleavage of the Prion Protein. J. Biol. Chem. 2014, 289, 803–813. [Google Scholar] [CrossRef] [Green Version]
  48. Jimenez-Huete, A.; Lievens, P.M.; Vidal, R.; Piccardo, P.; Ghetti, B.; Tagliavini, F.; Frangione, B.; Prelli, F. Endogenous proteolytic cleavage of normal and disease-associated isoforms of the human prion protein in neural and non-neural tissues. Am. J. Pathol. 1998, 153, 1561–1572. [Google Scholar] [CrossRef] [Green Version]
  49. Kuczius, T.; Grassi, J.; Karch, H.; Groschup, M.H. Binding of N- and C-terminal anti-prion protein antibodies generates distinct phenotypes of cellular prion proteins (PrPC) obtained from human, sheep, cattle and mouse. FEBS J. 2007, 274, 1492–1502. [Google Scholar] [CrossRef]
  50. Shyng, S.L.; Huber, M.T.; Harris, D.A. A prion protein cycles between the cell surface and an endocytic compartment in cultured neuroblastoma cells. J. Biol. Chem. 1993, 268, 15922–15928. [Google Scholar] [CrossRef]
  51. Taraboulos, A.; Scott, M.; Semenov, A.; Avrahami, D.; Laszlo, L.; Prusiner, S.B. Cholesterol depletion and modification of COOH-terminal targeting sequence of the prion protein inhibit formation of the scrapie isoform. J. Cell Biol. 1995, 129, 121–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Zhao, H.; Klingeborn, M.; Simonsson, M.; Linné, T. Proteolytic cleavage and shedding of the bovine prion protein in two cell culture systems. Virus Res. 2006, 115, 43–55. [Google Scholar] [CrossRef] [PubMed]
  53. Walmsley, A.R.; Watt, N.T.; Taylor, D.R.; Perera, W.S.S.; Hooper, N.M. α-cleavage of the prion protein occurs in a late compartment of the secretory pathway and is independent of lipid rafts. Mol. Cell Neurosci. 2009, 40, 242–248. [Google Scholar] [CrossRef] [PubMed]
  54. Nunziante, M.; Gilch, S.; Schätzl, H.M. Essential Role of the Prion Protein N Terminus in Subcellular Trafficking and Half-life of Cellular Prion Protein. J. Biol. Chem. 2003, 278, 3726–3734. [Google Scholar] [CrossRef] [Green Version]
  55. Béland, M.; Roucou, X. The prion protein unstructured N-terminal region is a broad-spectrum molecular sensor with diverse and contrasting potential functions. J. Neurochem. 2012, 120, 853–868. [Google Scholar] [CrossRef]
  56. Nieznanski, K.; Rutkowski, M.; Dominik, M.; Stepkowski, D. Proteolytic processing and glycosylation influence formation of porcine prion protein complexes. Biochem. J. 2005, 387, 93–100. [Google Scholar] [CrossRef] [Green Version]
  57. Wik, L.; Klingeborn, M.; Willander, H.; Linné, T. Separate mechanisms act concurrently to shed and release the prion protein from the cell. Prion 2012, 6, 498–509. [Google Scholar] [CrossRef] [Green Version]
  58. Lewis, V.; Hill, A.F.; Haigh, C.L.; Klug, G.M.; Masters, C.L.; Lawson, V.A.; Collins, S.J. Increased proportions of C1 truncated prion protein protect against cellular M1000 prion infection. J. Neuropathol. Exp. Neurol. 2009, 68, 1125–1135. [Google Scholar] [CrossRef] [Green Version]
  59. Westergard, L.; Turnbaugh, J.A.; Harris, D.A. A naturally occurring C-terminal fragment of the prion protein (PrP) delays disease and acts as a dominant-negative inhibitor of PrPSc formation. J. Biol. Chem. 2011, 286, 44234–44242. [Google Scholar] [CrossRef] [Green Version]
  60. Guillot-Sestier, M.V.; Sunyach, C.; Druon, C.; Scarzello, S.; Checler, F. The alpha-secretase-derived N-terminal product of cellular prion, N1, displays neuroprotective function in vitro and in vivo. J. Biol. Chem. 2009, 284, 35973–35986. [Google Scholar] [CrossRef] [Green Version]
  61. Guillot-Sestier, M.V.; Checler, F. Cellular prion and its catabolites in the brain: Production and function. Curr. Mol. Med. 2012, 12, 304–315. [Google Scholar] [CrossRef] [PubMed]
  62. Yusa, S.; Oliveira-Martins, J.B.; Sugita-Konishi, Y.; Kikuchi, Y. Cellular prion protein: From physiology to pathology. Viruses 2012, 4, 3109–3131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Altmeppen, H.C.; Puig, B.; Dohler, F.; Thurm, D.K.; Falker, C.; Krasemann, S.; Glatzel, M. Proteolytic processing of the prion protein in health and disease. Am. J. Neurodegener. Dis. 2012, 1, 15–31. [Google Scholar] [PubMed]
  64. McMahon, H.E.M.; Mangé, A.; Nishida, N.; Créminon, C.; Casanova, D.; Lehmann, S. Cleavage of the Amino Terminus of the Prion Protein by Reactive Oxygen Species*. J. Biol. Chem. 2001, 276, 2286–2291. [Google Scholar] [CrossRef] [Green Version]
  65. Watt, N.T.; Taylor, D.R.; Gillott, A.; Thomas, D.A.; Perera, W.S.; Hooper, N.M. Reactive oxygen species-mediated beta-cleavage of the prion protein in the cellular response to oxidative stress. J. Biol. Chem. 2005, 280, 35914–35921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Pushie, M.J.; Vogel, H.J. Modeling by Assembly and Molecular Dynamics Simulations of the Low Cu2+ Occupancy Form of the Mammalian Prion Protein Octarepeat Region: Gaining Insight into Cu2+-Mediated β-Cleavage. Biophys. J. 2008, 95, 5084–5091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Engelke, A.D.; Gonsberg, A.; Thapa, S.; Jung, S.; Ulbrich, S.; Seidel, R.; Basu, S.; Multhaup, G.; Baier, M.; Engelhard, M.; et al. Dimerization of the cellular prion protein inhibits propagation of scrapie prions. J. Biol. Chem. 2018, 293, 8020–8031. [Google Scholar] [CrossRef] [Green Version]
  68. Benestad, S.L.; Austbø, L.; Tranulis, M.A.; Espenes, A.; Olsaker, I. Healthy goats naturally devoid of prion protein. Vet. Res. 2012, 43, 87. [Google Scholar] [CrossRef] [Green Version]
  69. Meier, P.; Genoud, N.; Prinz, M.; Maissen, M.; Rülicke, T.; Zurbriggen, A.; Raeber, A.J.; Aguzzi, A. Soluble dimeric prion protein binds PrP(Sc) in vivo and antagonizes prion disease. Cell 2003, 113, 49–60. [Google Scholar] [CrossRef] [Green Version]
  70. Parizek, P. Similar turnover and shedding of the cellular prion protein in primary lymphoid and neuronal cells. J. Biol. Chem. 2001, 276, 44627–44632. [Google Scholar] [CrossRef] [Green Version]
  71. Oesch, B.; Westaway, D.; Walchli, M.; McKinley, M.P.; Kent, S.B.; Aebersold, R.; Barry, R.A.; Tempst, P.; Teplow, D.B.; Hood, L.E.; et al. A cellular gene encodes scrapie PrP 27-30 protein. Cell 1985, 40, 735–746. [Google Scholar] [CrossRef]
  72. Rogers, M.; Yehiely, F.; Scott, M.; Prusiner, S.B. Conversion of truncated and elongated prion proteins into the scrapie isoform in cultured cells. Proc. Natl. Acad. Sci. USA 1993, 90, 3182–3186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Lewis, V.; Johanssen, V.A.; Crouch, P.J.; Klug, G.M.; Hooper, N.M.; Collins, S.J. Prion protein “gamma-cleavage”: Characterizing a novel endoproteolytic processing event. Cell Mol. Life Sci. 2016, 73, 667–683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Haigh, C.; Collins, S. Endoproteolytic cleavage as a molecular switch regulating and diversifying prion protein function. Neural Regen. Res. 2016, 11, 238–239. [Google Scholar] [CrossRef] [PubMed]
  75. Kojima, A.; Konishi, M.; Akizawa, T. Prion Fragment Peptides Are Digested with Membrane Type Matrix Metalloproteinases and Acquire Enzyme Resistance through Cu2+-Binding. Biomolecules 2014, 4, 510–526. [Google Scholar] [CrossRef] [Green Version]
  76. Stahl, N.; Baldwin, M.A.; Burlingame, A.L.; Prusiner, S.B. Identification of glycoinositol phospholipid linked and truncated forms of the scrapie prion protein. Biochemistry 1990, 29, 8879–8884. [Google Scholar] [CrossRef]
  77. Tagliavini, F.; Prelli, F.; Porro, M.; Salmona, M.; Bugiani, O.; Frangione, B. A soluble form of prion protein in human cerebrospinal fluid: Implications for prion-related encephalopathies. Biochem. Biophys. Res. Commun. 1992, 184, 1398–1404. [Google Scholar] [CrossRef]
  78. Altmeppen, H.C.; Prox, J.; Krasemann, S.; Puig, B.; Kruszewski, K.; Dohler, F.; Bernreuther, C.; Hoxha, A.; Linsenmeier, L.; Sikorska, B.; et al. The sheddase ADAM10 is a potent modulator of prion disease. Elife 2015, 4, e04260. [Google Scholar] [CrossRef]
  79. Glatzel, M.; Linsenmeier, L.; Dohler, F.; Krasemann, S.; Puig, B.; Altmeppen, H.C. Shedding light on prion disease. Prion 2015, 9, 244–256. [Google Scholar] [CrossRef] [Green Version]
  80. Altmeppen, H.C.; Prox, J.; Puig, B.; Dohler, F.; Falker, C.; Krasemann, S.; Glatzel, M. Roles of endoproteolytic α-cleavage and shedding of the prion protein in neurodegeneration. FEBS J. 2013, 280, 4338–4347. [Google Scholar] [CrossRef]
  81. Linsenmeier, L.; Mohammadi, B.; Shafiq, M.; Frontzek, K.; Bär, J.; Shrivastava, A.N.; Damme, M.; Song, F.; Schwarz, A.; Da Vela, S.; et al. Ligands binding to the prion protein induce its proteolytic release with therapeutic potential in neurodegenerative proteinopathies. Sci. Adv. 2021, 7, eabj1826. [Google Scholar] [CrossRef] [PubMed]
  82. Jarosz-Griffiths, H.H.; Corbett, N.J.; Rowland, H.A.; Fisher, K.; Jones, A.C.; Baron, J.; Howell, G.J.; Cowley, S.A.; Chintawar, S.; Cader, M.Z.; et al. Proteolytic shedding of the prion protein via activation of metallopeptidase ADAM10 reduces cellular binding and toxicity of amyloid-β; oligomers. J. Biol. Chem. 2019, 294, 7085–7097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Megra, B.W.; Eugenin, E.A.; Berman, J.W. The Role of Shed PrP(C) in the Neuropathogenesis of HIV Infection. J. Immunol. 2017, 199, 224–232. [Google Scholar] [CrossRef] [PubMed]
  84. Taylor, D.R.; Parkin, E.T.; Cocklin, S.L.; Ault, J.R.; Ashcroft, A.E.; Turner, A.J. Role of ADAMs in the ectodomain shedding and conformational conversion of the prion protein. J. Biol. Chem. 2009, 284, 22590–22600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Altmeppen, H.C.; Prox, J.; Puig, B.; Kluth, M.A.; Bernreuther, C.; Thurm, D.; Jorissen, E.; Petrowitz, B.; Bartsch, U.; De Strooper, B.; et al. Lack of a-disintegrin-and-metalloproteinase ADAM10 leads to intracellular accumulation and loss of shedding of the cellular prion protein in vivo. Mol. Neurodegener. 2011, 6, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. McDonald, A.J.; Millhauser, G.L. PrP overdrive: Does inhibition of α-cleavage contribute to PrP(C) toxicity and prion disease? Prion 2014, 8, 183–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Tucher, J.; Linke, D.; Koudelka, T.; Cassidy, L.; Tredup, C.; Wichert, R.; Pietrzik, C.; Becker-Pauly, C.; Tholey, A. LC-MS based cleavage site profiling of the proteases ADAM10 and ADAM17 using proteome-derived peptide libraries. J. Proteome Res. 2014, 13, 2205–2214. [Google Scholar] [CrossRef] [PubMed]
  88. Jansen, C.; Parchi, P.; Capellari, S.; Vermeij, A.J.; Corrado, P.; Baas, F.; Strammiello, R.; van Gool, W.A.; van Swieten, J.C.; Rozemuller, A.J. Prion protein amyloidosis with divergent phenotype associated with two novel nonsense mutations in PRNP. Acta Neuropathol. 2010, 119, 189–197. [Google Scholar] [CrossRef] [Green Version]
  89. Čurin Šerbec, V.; Bresjanac, M.; Popović, M.; Pretnar Hartman, K.; Galvani, V.; Rupreht, R.; Černilec, M.; Vranac, T.; Hafner, I.; Jerala, R. Monoclonal antibody against a peptide of human prion protein discriminates between Creutzfeldt-Jacob’s disease-affected and normal brain tissue. J. Biol. Chem. 2004, 279, 3694–3698. [Google Scholar] [CrossRef] [Green Version]
  90. Dvorakova, E.; Vranac, T.; Janouskova, O.; Černilec, M.; Koren, S.; Lukan, A.; Novakova, J.; Matej, R.; Holada, K.; Čurin Šerbec, V. Detection of the GPI-anchorless prion protein fragment PrP226* in human brain. BMC Neurol. 2013, 13, 126. [Google Scholar] [CrossRef] [Green Version]
  91. Koren, S.; Kosmač, M.; Colja Venturini, A.; Montanič, S.; Čurin Šerbec, V. Antibody variable-region sequencing as a method for hybridoma cell-line authentication. Appl. Microbiol. Biotechnol. 2008, 78, 1071–1078. [Google Scholar] [CrossRef] [PubMed]
  92. Kovač, V.; Hafner-Bratkovič, I.; Čurin Šerbec, V. Anchorless forms of prion protein—Impact of truncation on structure destabilization and prion protein conversion. Biochem. Biophys. Res. Commun. 2016, 481, 1–6. [Google Scholar] [CrossRef] [Green Version]
  93. Kovač, V.; Zupančič, B.; Ilc, G.; Plavec, J.; Čurin Šerbec, V. Truncated prion protein PrP226*—A structural view on its role in amyloid disease. Biochem. Biophys. Res. Commun. 2017, 484, 45–50. [Google Scholar] [CrossRef] [PubMed]
  94. Lukan, A.; Černilec, M.; Vranac, T.; Popović, M.; Čurin Šerbec, V. Regional distribution of anchorless prion protein, PrP226 *, in the human brain. Prion 2014, 8, 203–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Kovač, V.; Čurin Šerbec, V. Prion Proteins Without the Glycophosphatidylinositol Anchor: Potential Biomarkers in Neurodegenerative Diseases. Biomark. Insights 2018, 13, 1177271918756648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Salès, N.; Hässig, R.; Rodolfo, K.; Giamberardino, L.; Traiffort, E.; Ruat, M. Developmental expression of the cellular prion protein in elongating axons. Eur. J. Neurosci. 2002, 15, 1163–1177. [Google Scholar] [CrossRef]
  97. Salès, N.; Rodolfo, K.; Hässig, R.; Faucheux, B.; Giamberardino, L.; Moya, K.L. Cellular prion protein localization in rodent and primate brain. Eur. J. Neurosci. 1998, 10, 2464–2471. [Google Scholar] [CrossRef]
  98. Mironov, A.; Latawiec, D.; Wille, H.; Bouzamondo-Bernstein, E.; Legname, G.; Williamson, R.A. Cytosolic prion protein in neurons. J. Neurosci. 2003, 23, 7183–7193. [Google Scholar] [CrossRef]
  99. Baumann, F.; Tolnay, M.; Brabeck, C.; Pahnke, J.; Kloz, U.; Niemann, H.H. Lethal recessive myelin toxicity of prion protein lacking its central domain. EMBO J. 2007, 26, 538–547. [Google Scholar] [CrossRef]
  100. Radovanovic, I. Truncated prion protein and Doppel are myelinotoxic in the absence of oligodendrocytic PrPC. J. Neurosci. 2005, 25, 4879–4888. [Google Scholar] [CrossRef] [Green Version]
  101. Bremer, J.; Baumann, F.; Tiberi, C.; Wessig, C.; Fischer, H.; Schwarz, P. Axonal prion protein is required for peripheral myelin maintenance. Nat. Neurosci. 2010, 13, 310–318. [Google Scholar] [CrossRef] [PubMed]
  102. Nuvolone, M.; Hermann, M.; Sorce, S.; Russo, G.; Tiberi, C.; Schwarz, P.; Minikel, E.; Sanoudou, D.; Pelczar, P.; Aguzzi, A. Strictly co-isogenic C57BL/6J-Prnp−/−mice: A rigorous resource for prion science. J. Exp. Med. 2016, 213, 313–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Nishida, N.; Tremblay, P.; Sugimoto, T.; Shigematsu, K.; Shirabe, S.; Petromilli, C.; Erpel, S.P.; Nakaoke, R.; Atarashi, R.; Houtani, T.; et al. A mouse prion protein transgene rescues mice deficient for the prion protein gene from purkinje cell degeneration and demyelination. Lab. Investig. A J. Tech. Methods Pathol. 1999, 79, 689–697. [Google Scholar]
  104. Henzi, A.; Senatore, A.; Lakkaraju, A.K.K.; Scheckel, C.; Mühle, J.; Reimann, R.; Sorce, S.; Schertler, G.; Toyka, K.V.; Aguzzi, A. Soluble dimeric prion protein ligand activates Adgrg6 receptor but does not rescue early signs of demyelination in PrP-deficient mice. PLoS ONE 2020, 15, e0242137. [Google Scholar] [CrossRef]
  105. Henzi, A.; Aguzzi, A. The prion protein is not required for peripheral nerve de- and remyelination after crush injury. PLoS ONE 2021, 16, e0245944. [Google Scholar] [CrossRef]
  106. Büeler, H.; Fischer, M.; Lang, Y.; Bluethmann, H.; Lipp, H.P.; DeArmond, S.J.; Prusiner, S.B.; Aguet, M.; Weissmann, C. Normal development and behaviour of mice lacking the neuronal cell-surface PrP protein. Nature 1992, 356, 577–582. [Google Scholar] [CrossRef]
  107. Manson, J.C.; Clarke, A.R.; Hooper, M.L.; Aitchison, L.; McConnell, I.; Hope, J. 129/Ola mice carrying a null mutation in PrP that abolishes mRNA production are developmentally normal. Mol. Neurobiol. 1994, 8, 121–127. [Google Scholar] [CrossRef]
  108. Steele, A.D.; Lindquist, S.; Aguzzi, A. The prion protein knockout mouse: A phenotype under challenge. Prion 2007, 1, 83–93. [Google Scholar] [CrossRef] [Green Version]
  109. Rossi, D.; Cozzio, A.; Flechsig, E.; Klein, M.A.; Rülicke, T.; Aguzzi, A.; Weissmann, C. Onset of ataxia and Purkinje cell loss in PrP null mice inversely correlated with Dpl level in brain. EMBO J. 2001, 20, 694–702. [Google Scholar] [CrossRef] [Green Version]
  110. Richt, J.A.; Kasinathan, P.; Hamir, A.N.; Castilla, J.; Sathiyaseelan, T.; Vargas, F. Production of cattle lacking prion protein. Nat. Biotechnol. 2007, 25, 132–138. [Google Scholar] [CrossRef]
  111. Skedsmo, F.S.; Malachin, G.; Våge, D.I.; Hammervold, M.M.; Salvesen, Ø.; Ersdal, C.; Ranheim, B.; Stafsnes, M.H.; Bartosova, Z.; Bruheim, P.; et al. Demyelinating polyneuropathy in goats lacking prion protein. FASEB J. 2020, 34, 2359–2375. [Google Scholar] [CrossRef] [PubMed]
  112. Salvesen, Ø.; Tatzelt, J.; Tranulis, M.A. The prion protein in neuroimmune crosstalk. Neurochem. Int. 2019, 130, 104335. [Google Scholar] [CrossRef] [PubMed]
  113. Criado, J.R.; Sánchez-Alavez, M.; Conti, B.; Giacchino, J.L.; Wills, D.N.; Henriksen, S.J. Mice devoid of prion protein have cognitive deficits that are rescued by reconstitution of PrP in neurons. Neurobiol. Dis. 2005, 19, 255–265. [Google Scholar] [CrossRef] [PubMed]
  114. Sakaguchi, S.; Katamine, S.; Nishida, N.; Moriuchi, R.; Shigematsu, K.; Sugimoto, T.; Nakatani, A.; Kataoka, Y.; Houtani, T.; Shirabe, S.; et al. Loss of cerebellar Purkinje cells in aged mice homozygous for a disrupted PrP gene. Nature 1996, 380, 528–531. [Google Scholar] [CrossRef] [PubMed]
  115. Moore, R.C.; Lee, I.Y.; Silverman, G.L.; Harrison, P.M.; Strome, R.; Heinrich, C.; Karunaratne, A.; Pasternak, S.H.; Chishti, M.A.; Liang, Y.; et al. Ataxia in prion protein (PrP)-deficient mice is associated with upregulation of the novel PrP-like protein doppel. J. Mol. Biol. 1999, 292, 797–817. [Google Scholar] [CrossRef] [PubMed]
  116. Mallucci, G.R.; Ratté, S.; Asante, E.A.; Linehan, J.; Gowland, I.; Jefferys, J.G.; Collinge, J. Post-natal knockout of prion protein alters hippocampal CA1 properties, but does not result in neurodegeneration. Embo J. 2002, 21, 202–210. [Google Scholar] [CrossRef]
  117. White, A.R.; Collins, S.J.; Maher, F.; Jobling, M.F.; Stewart, L.R.; Thyer, J.M. Prion protein-deficient neurons reveal lower glutathione reductase activity and increased susceptibility to hydrogen peroxide toxicity. Am. J. Pathol. 1999, 155, 1723–1730. [Google Scholar] [CrossRef] [Green Version]
  118. Hutter, G.; Heppner, F.L.; Aguzzi, A. No Superoxide Dismutase Activity of Cellular Prion Protein in vivo. Biol. Chem. 2003, 384, 1279–1285. [Google Scholar] [CrossRef] [Green Version]
  119. Davies, P.; Brown, D.R. The chemistry of copper binding to PrP: Is there sufficient evidence to elucidate a role for copper in protein function? Biochem. J. 2008, 410, 237–244. [Google Scholar] [CrossRef] [Green Version]
  120. Brown, D.R.; Nicholas, R.S.; Canevari, L. Lack of prion protein expression results in a neuronal phenotype sensitive to stress. J. Neurosci. Res. 2002, 67, 211–224. [Google Scholar] [CrossRef]
  121. Brown, D.R.; Boon-Seng, W.; Hafiz, F.; Clive, C.; Haswell, S.J.; Jones, I.M. Normal prion protein has an activity like that of superoxide dismutase. Biochem. J. 1999, 344, 1–5. [Google Scholar] [CrossRef] [PubMed]
  122. Brown, D.R.; Schulz-Schaeffer, W.J.; Schmidt, B.; Kretzschmar, H.A. Prion protein-deficient cells show altered response to oxidative stress due to decreased SOD-1 activity. Exp. Neurol. 1997, 146, 104–112. [Google Scholar] [CrossRef] [PubMed]
  123. Sakudo, A.; Lee, D.-c.; Saeki, K.; Nakamura, Y.; Inoue, K.; Matsumoto, Y.; Itohara, S.; Onodera, T. Impairment of superoxide dismutase activation by N-terminally truncated prion protein (PrP) in PrP-deficient neuronal cell line. Biochem. Biophys. Res. Commun. 2003, 308, 660–667. [Google Scholar] [CrossRef]
  124. Guentchev, M.; Voigtländer, T.; Haberler, C.; Groschup, M.H.; Budka, H. Evidence for oxidative stress in experimental prion disease. Neurobiol. Dis. 2000, 7, 270–273. [Google Scholar] [CrossRef] [Green Version]
  125. Klamt, F.; Dal-Pizzol, F.; Conte da Frota, M.L.; Walz, R.; Andrades, M.E.; Silva, E.G. Imbalance of antioxidant defense in mice lacking cellular prion protein. Free Radic. Biol. Med. 2001, 30, 1137–1144. [Google Scholar] [CrossRef]
  126. Wong, B.S.; Liu, T.; Li, R.; Pan, T.; Petersen, R.B.; Smith, M.A.; Gambetti, P.; Perry, G.; Manson, J.C.; Brown, D.R.; et al. Increased levels of oxidative stress markers detected in the brains of mice devoid of prion protein. J. Neurochem. 2001, 76, 565–572. [Google Scholar] [CrossRef] [Green Version]
  127. McLennan, N.F.; Brennan, P.M.; McNeill, A.; Davies, I.; Fotheringham, A.; Rennison, K.A.; Ritchie, D.; Brannan, F.; Head, M.W.; Ironside, J.W. Prion protein accumulation and neuroprotection in hypoxic brain damage. Am. J. Pathol. 2004, 165, 227–235. [Google Scholar] [CrossRef] [Green Version]
  128. Brenna, S.; Altmeppen, H.C.; Mohammadi, B.; Rissiek, B.; Schlink, F.; Ludewig, P.; Krisp, C.; Schlüter, H.; Failla, A.V.; Schneider, C.; et al. Characterization of brain-derived extracellular vesicles reveals changes in cellular origin after stroke and enrichment of the prion protein with a potential role in cellular uptake. J. Extracell. Vesicles 2020, 9, 1809065. [Google Scholar] [CrossRef]
  129. Puig, B.; Yang, D.; Brenna, S.; Altmeppen, H.C.; Magnus, T. Show Me Your Friends and I Tell You Who You Are: The Many Facets of Prion Protein in Stroke. Cells 2020, 9, 1609. [Google Scholar] [CrossRef]
  130. Zeng, L.; Zou, W.; Wang, G. Cellular prion protein (PrP(C)) and its role in stress responses. Int. J. Clin. Exp. Med. 2015, 8, 8042–8050. [Google Scholar]
  131. Mitsios, N.; Saka, M.; Krupinski, J.; Pennucci, R.; Sanfeliu, C.; Miguel Turu, M.; Gaffney, J.; Kumar, P.; Kumar, S.; Sullivan, M.; et al. Cellular prion protein is increased in the plasma and peri-infarcted brain tissue after acute stroke. J. Neurosci. Res. 2007, 85, 602–611. [Google Scholar] [CrossRef] [PubMed]
  132. Pham, N.; Dhar, A.; Khalaj, S.; Desai, K.; Taghibiglou, C. Down regulation of brain cellular prion protein in an animal model of insulin resistance: Possible implication in increased prevalence of stroke in pre-diabetics/diabetics. Biochem. Biophys. Res. Commun. 2014, 448, 151–156. [Google Scholar] [CrossRef] [PubMed]
  133. Shyu, W.-C. Overexpression of PrPC by adenovirus-mediated gene targeting reduces ischemic injury in a stroke rat model. J. Neurosci. 2005, 25, 8967–8977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Spudich, A.; Frigg, R.; Kilic, E.; Kilic, Ü.; Oesch, B.; Raeber, A.; Bassetti, C.L.; Hermann, D.M. Aggravation of ischemic brain injury by prion protein deficiency: Role of ERK-1/-2 and STAT-1. Neurobiol. Dis. 2005, 20, 442–449. [Google Scholar] [CrossRef]
  135. Weise, J.; Crome, O.; Sandau, R.; Schulz-Schaeffer, W.; Bähr, M.; Zerr, I. Upregulation of cellular prion protein (PrPc) after focal cerebral ischemia and influence of lesion severity. Neurosci. Lett. 2004, 372, 146–150. [Google Scholar] [CrossRef]
  136. Weise, J.; Sandau, R.; Schwarting, S.; Crome, O.; Wrede, A.; Schulz-Schaeffer, W. Deletion of cellular prion protein results in reduced Akt activation, enhanced postischemic caspase-3 activation, and exacerbation of ischemic brain injury. Stroke 2006, 37, 1296–1300. [Google Scholar] [CrossRef] [Green Version]
  137. Wang, V.; Chuang, T.-C.; Hsu, Y.-D.; Chou, W.-Y.; Kao, M.-C. Nitric oxide induces prion protein via MEK and p38 MAPK signaling. Biochem. Biophys. Res. Commun. 2005, 333, 95–100. [Google Scholar] [CrossRef]
  138. Hemmings, B.A.; Restuccia, D.F. Pi3k-pkb/akt pathway. Cold Spring Harb. Perspect. Biol. 2012, 4, a011189. [Google Scholar] [CrossRef] [Green Version]
  139. Brazil, D.P.; Hemmings, B.A. Ten years of protein kinase B signalling: A hard Akt to follow. Trends Biochem. Sci. 2001, 26, 657–664. [Google Scholar] [CrossRef]
  140. Manning, B.D.; Cantley, L.C. AKT/PKB signaling: Navigating downstream. Cell 2007, 129, 1261–1274. [Google Scholar] [CrossRef] [Green Version]
  141. Mitteregger, G.; Vosko, M.; Krebs, B.; Xiang, W.; Kohlmannsperger, V.; Nölting, S. The role of the octarepeat region in neuroprotective function of the cellular prion protein. Brain Pathol. 2007, 17, 174–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Vassallo, N.; Herms, J.; Behrens, C.; Krebs, B.; Saeki, K.; Onodera, T.; Windl, O.; Kretzschmar, H.A. Activation of phosphatidylinositol 3-kinase by cellular prion protein and its role in cell survival. Biochem. Biophys. Res. Commun. 2005, 332, 75–82. [Google Scholar] [CrossRef] [PubMed]
  143. Black, S.A.G.; Stys, P.K.; Zamponi, G.W.; Tsutsui, S. Cellular prion protein and NMDA receptor modulation: Protecting against excitotoxicity. Front. Cell Dev. Biol. 2014, 2, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Chiarini, L.B.; Freitas, A.R.O.; Zanata, S.M.; Brentani, R.R.; Martins, V.R.; Linden, R. Cellular prion protein transduces neuroprotective signals. EMBO J. 2002, 21, 3317–3326. [Google Scholar] [CrossRef] [PubMed]
  145. Chen, S.; Mangé, A.; Dong, L.; Lehmann, S.; Schachner, M. Prion protein as trans-interacting partner for neurons is involved in neurite outgrowth and neuronal survival. Mol. Cell. Neurosci. 2003, 22, 227–233. [Google Scholar] [CrossRef]
  146. Lopes, M.H.; Hajj, G.N.; Muras, A.G.; Mancini, G.L.; Castro, R.M.; Ribeiro, K.C.; Brentani, R.R.; Linden, R.; Martins, V.R. Interaction of cellular prion and stress-inducible protein 1 promotes neuritogenesis and neuroprotection by distinct signaling pathways. J. Neurosci. 2005, 25, 11330–11339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Crestini, A.; Santilli, F.; Martellucci, S.; Carbone, E.; Sorice, M.; Piscopo, P.; Mattei, V. Prions and Neurodegenerative Diseases: A Focus on Alzheimer’s Disease. J. Alzheimer’s Dis. 2022, 85, 503–518. [Google Scholar] [CrossRef]
  148. Angelopoulou, E.; Paudel, Y.N.; Julian, T.; Shaikh, M.F.; Piperi, C. Pivotal Role of Fyn Kinase in Parkinson’s Disease and Levodopa-Induced Dyskinesia: A Novel Therapeutic Target? Mol. Neurobiol. 2021, 58, 1372–1391. [Google Scholar] [CrossRef]
  149. Grayson, J.D.; Baumgartner, M.P.; Santos Souza, C.D.; Dawes, S.J.; El Idrissi, I.G.; Louth, J.C.; Stimpson, S.; Mead, E.; Dunbar, C.; Wolak, J.; et al. Amyloid binding and beyond: A new approach for Alzheimer’s disease drug discovery targeting Aβo–PrPC binding and downstream pathways. Chem. Sci. 2021, 12, 3768–3785. [Google Scholar] [CrossRef]
  150. Briner, A.; Götz, J.; Polanco, J.C. Fyn Kinase Controls Tau Aggregation In Vivo. Cell Rep. 2020, 32, 108045. [Google Scholar] [CrossRef]
  151. Santuccione, A.; Sytnyk, V.; Leshchyns’ka, I.; Schachner, M. Prion protein recruits its neuronal receptor NCAM to lipid rafts to activate p59 fyn and to enhance neurite outgrowth. J. Cell Biol. 2005, 169, 341–354. [Google Scholar] [CrossRef] [Green Version]
  152. Cheung, H.H.; Takagi, N.; Teves, L.; Logan, R.; Wallace, M.C.; Gurd, J.W. Altered association of protein tyrosine kinases with postsynaptic densities after transient cerebral ischemia in the rat brain. J. Cereb. Blood Flow Metab. 2000, 20, 505–512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Takagi, N.; Cheung, H.H.; Bissoon, N.; Teves, L.; Wallace, M.C.; Gurd, J.W. The effect of transient global ischemia on the interaction of Src and Fyn with the N-methyl-D-aspartate receptor and postsynaptic densities: Possible involvement of Src homology 2 domains. J. Cereb. Blood Flow Metab. 1999, 19, 880–888. [Google Scholar] [CrossRef] [Green Version]
  154. Knox, R.; Jiang, X. Fyn in Neurodevelopment and Ischemic Brain Injury. Dev. Neurosci. 2015, 37, 311–320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Paul, R.; Zhang, Z.G.; Eliceiri, B.P.; Jiang, Q.; Boccia, A.D.; Zhang, R.L.; Chopp, M.; Cheresh, D.A. Src deficiency or blockade of Src activity in mice provides cerebral protection following stroke. Nat. Med. 2001, 7, 222–227. [Google Scholar] [CrossRef] [PubMed]
  156. Hou, X.-Y.; Liu, Y.; Zhang, G.-Y. PP2, a potent inhibitor of Src family kinases, protects against hippocampal CA1 pyramidal cell death after transient global brain ischemia. Neurosci. Lett. 2007, 420, 235–239. [Google Scholar] [CrossRef] [PubMed]
  157. Knox, R.; Brennan-Minnella, A.M.; Lu, F.; Yang, D.; Nakazawa, T.; Yamamoto, T.; Swanson, R.A.; Ferriero, D.M.; Jiang, X. NR2B Phosphorylation at Tyrosine 1472 Contributes to Brain Injury in a Rodent Model of Neonatal Hypoxia-Ischemia. Stroke 2014, 45, 3040–3047. [Google Scholar] [CrossRef]
  158. Du, C.-P.; Gao, J.; Tai, J.-M.; Liu, Y.; Qi, J.; Wang, W.; Hou, X.-Y. Increased tyrosine phosphorylation of PSD-95 by Src family kinases after brain ischaemia. Biochem. J. 2008, 417, 277–285. [Google Scholar] [CrossRef] [Green Version]
  159. Knox, R.; Zhao, C.; Miguel-Perez, D.; Wang, S.; Yuan, J.; Ferriero, D.; Jiang, X. Enhanced NMDA receptor tyrosine phosphorylation and increased brain injury following neonatal hypoxia–ischemia in mice with neuronal Fyn overexpression. Neurobiol. Dis. 2013, 51, 113–119. [Google Scholar] [CrossRef] [Green Version]
  160. Haigh, C.L.; Drew, S.C.; Boland, M.P.; Masters, C.L.; Barnham, K.J.; Lawson, V.A.; Collins, S.J. Dominant roles of the polybasic proline motif and copper in the PrP23-89-mediated stress protection response. J. Cell Sci. 2009, 122, 1518–1528. [Google Scholar] [CrossRef] [Green Version]
  161. Haigh, C.L.; McGlade, A.R.; Collins, S.J. MEK1 transduces the prion protein N2 fragment antioxidant effects. Cell. Mol. Life Sci. 2015, 72, 1613–1629. [Google Scholar] [CrossRef] [PubMed]
  162. Haigh, C.L.; Tumpach, C.; Drew, S.C.; Collins, S.J. The Prion Protein N1 and N2 Cleavage Fragments Bind to Phosphatidylserine and Phosphatidic Acid; Relevance to Stress-Protection Responses. PLoS ONE 2015, 10, e0134680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Collins, S.J.; Tumpach, C.; Groveman, B.R.; Drew, S.C.; Haigh, C.L. Prion protein cleavage fragments regulate adult neural stem cell quiescence through redox modulation of mitochondrial fission and SOD2 expression. Cell. Mol. Life Sci. 2018, 75, 3231–3249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Sunyach, C.; Cisse, M.A.; da Costa, C.A.; Vincent, B.; Checler, F. The C-terminal Products of Cellular Prion Protein Processing, C1 and C2, Exert Distinct Influence on p53-dependent Staurosporine-induced Caspase-3 Activation *. J. Biol. Chem. 2007, 282, 1956–1963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Pager, C.T.; Craft, W.W.; Patch, J.; Dutch, R.E. A mature and fusogenic form of the Nipah virus fusion protein requires proteolytic processing by cathepsin L. Virology 2006, 346, 251–257. [Google Scholar] [CrossRef] [PubMed]
  166. Diederich, S.; Sauerhering, L.; Weis, M.; Altmeppen, H.; Schaschke, N.; Reinheckel, T.; Erbar, S.; Maisner, A. Activation of the Nipah Virus Fusion Protein in MDCK Cells Is Mediated by Cathepsin B within the Endosome-Recycling Compartment. J. Virol. 2012, 86, 3736–3745. [Google Scholar] [CrossRef] [Green Version]
  167. Carroll, J.A.; Groveman, B.R.; Williams, K.; Moore, R.; Race, B.; Haigh, C.L. Prion protein N1 cleavage peptides stimulate microglial interaction with surrounding cells. Sci. Rep. 2020, 10, 6654. [Google Scholar] [CrossRef] [Green Version]
  168. Martellucci, S.; Santacroce, C.; Santilli, F.; Piccoli, L.; Delle Monache, S.; Angelucci, A.; Misasi, R.; Sorice, M.; Mattei, V. Cellular and Molecular Mechanisms Mediated by recPrPC Involved in the Neuronal Differentiation Process of Mesenchymal Stem Cells. Int. J. Mol. Sci. 2019, 20, 345. [Google Scholar] [CrossRef] [Green Version]
  169. Amin, L.; Nguyen, X.T.; Rolle, I.G.; D’Este, E.; Giachin, G.; Tran, T.H.; Serbec, V.C.; Cojoc, D.; Legname, G. Characterization of prion protein function by focal neurite stimulation. J. Cell Sci. 2016, 129, 3878–3891. [Google Scholar] [CrossRef] [Green Version]
  170. Mantuano, E.; Azmoon, P.; Banki, M.A.; Lam, M.S.; Sigurdson, C.J.; Gonias, S.L. A soluble derivative of PrP(C) activates cell-signaling and regulates cell physiology through LRP1 and the NMDA receptor. J. Biol. Chem. 2020, 295, 14178–14188. [Google Scholar] [CrossRef]
  171. Shi, Y.; Mantuano, E.; Inoue, G.; Campana, W.M.; Gonias, S.L. Ligand binding to LRP1 transactivates Trk receptors by a Src family kinase-dependent pathway. Sci. Signal. 2009, 2, ra18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Prusiner, S.B. Novel proteinaceous infectious particles cause scrapie. Science 1982, 216, 136–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Stahl, N.; Borchelt, D.R.; Hsiao, K.; Prusiner, S.B. Scrapie prion protein contains a phosphatidylinositol glycolipid. Cell 1987, 51, 229–240. [Google Scholar] [CrossRef]
  174. Desplats, P.; Lee, H.-J.; Bae, E.-J.; Patrick, C.; Rockenstein, E.; Crews, L.; Spencer, B.; Masliah, E.; Lee, S.-J. Inclusion formation and neuronal cell death through neuron-to-neuron transmission of α-synuclein. Proc. Natl. Acad. Sci. USA 2009, 106, 13010–13015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Ferreira, D.G.; Temido-Ferreira, M.; Vicente Miranda, H.; Batalha, V.L.; Coelho, J.E.; Szegö, É.M.; Marques-Morgado, I.; Vaz, S.H.; Rhee, J.S.; Schmitz, M.; et al. α-synuclein interacts with PrPC to induce cognitive impairment through mGluR5 and NMDAR2B. Nat. Neurosci. 2017, 20, 1569–1579. [Google Scholar] [CrossRef] [PubMed]
  176. del Río, J.A.; Ferrer, I.; Gavín, R. Role of cellular prion protein in interneuronal amyloid transmission. Prog. Neurobiol. 2018, 165–167, 87–102. [Google Scholar] [CrossRef] [Green Version]
  177. Urrea, L.; Segura-Feliu, M.; Masuda-Suzukake, M.; Hervera, A.; Pedraz, L.; García Aznar, J.M.; Vila, M.; Samitier, J.; Torrents, E.; Ferrer, I.; et al. Involvement of Cellular Prion Protein in α-Synuclein Transport in Neurons. Mol. Neurobiol. 2018, 55, 1847–1860. [Google Scholar] [CrossRef] [Green Version]
  178. Domert, J.; Rao, S.B.; Agholme, L.; Brorsson, A.-C.; Marcusson, J.; Hallbeck, M.; Nath, S. Spreading of amyloid-β peptides via neuritic cell-to-cell transfer is dependent on insufficient cellular clearance. Neurobiol. Dis. 2014, 65, 82–92. [Google Scholar] [CrossRef] [Green Version]
  179. Eisele, Y.S.; Obermüller, U.; Heilbronner, G.; Baumann, F.; Kaeser, S.A.; Wolburg, H.; Walker, L.C.; Staufenbiel, M.; Heikenwalder, M.; Jucker, M. Peripherally applied Aβ-containing inoculates induce cerebral β-amyloidosis. Science 2010, 330, 980–982. [Google Scholar] [CrossRef] [Green Version]
  180. Clavaguera, F.; Bolmont, T.; Crowther, R.A.; Abramowski, D.; Frank, S.; Probst, A.; Fraser, G.; Stalder, A.K.; Beibel, M.; Staufenbiel, M. Transmission and spreading of tauopathy in transgenic mouse brain. Nat. Cell Biol. 2009, 11, 909–913. [Google Scholar] [CrossRef]
  181. Sydow, A.; Mandelkow, E.M. ‘Prion-Like’ Propagation of Mouse and Human Tau Aggregates in an Inducible Mouse Model of Tauopathy. Neurodegener. Dis. 2010, 7, 28–31. [Google Scholar] [CrossRef] [PubMed]
  182. Guo, J.L.; Lee, V.M.-Y. Seeding of normal Tau by pathological Tau conformers drives pathogenesis of Alzheimer-like tangles. J. Biol. Chem. 2011, 286, 15317–15331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Clavaguera, F.; Akatsu, H.; Fraser, G.; Crowther, R.A.; Frank, S.; Hench, J.; Probst, A.; Winkler, D.T.; Reichwald, J.; Staufenbiel, M.; et al. Brain homogenates from human tauopathies induce tau inclusions in mouse brain. Proc. Natl. Acad. Sci. USA 2013, 110, 9535–9540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Takahashi, R.H.; Yokotsuka, M.; Tobiume, M.; Sato, Y.; Hasegawa, H.; Nagao, T.; Gouras, G.K. Accumulation of cellular prion protein within β-amyloid oligomer plaques in aged human brains. Brain Pathol. 2021, 31, e12941. [Google Scholar] [CrossRef]
  185. Aulić, S.; Masperone, L.; Narkiewicz, J.; Isopi, E.; Bistaffa, E.; Ambrosetti, E.; Pastore, B.; De Cecco, E.; Scaini, D.; Zago, P. α-Synuclein amyloids hijack prion protein to gain cell entry, facilitate cell-to-cell spreading and block prion replication. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef]
  186. Haas, L.T.; Salazar, S.V.; Kostylev, M.A.; Um, J.W.; Kaufman, A.C.; Strittmatter, S.M. Metabotropic glutamate receptor 5 couples cellular prion protein to intracellular signalling in Alzheimer’s disease. Brain A J. Neurol. 2016, 139, 526–546. [Google Scholar] [CrossRef] [Green Version]
  187. Hachiya, N.; Fułek, M.; Zajączkowska, K.; Kurpas, D.; Trypka, E.; Leszek, J. Cellular Prion Protein and Amyloid–β Oligomers in Alzheimer’s Disease–There Are Connections? Preprints 2021, 2021050032. [Google Scholar] [CrossRef]
  188. Kostylev, M.A.; Tuttle, M.D.; Lee, S.; Klein, L.E.; Takahashi, H.; Cox, T.O.; Gunther, E.C.; Zilm, K.W.; Strittmatter, S.M. Liquid and Hydrogel Phases of PrPC Linked to Conformation Shifts and Triggered by Alzheimer’s Amyloid-β; Oligomers. Mol. Cell 2018, 72, 426–443.e412. [Google Scholar] [CrossRef] [Green Version]
  189. Um, J.W.; Kaufman, A.C.; Kostylev, M.; Heiss, J.K.; Stagi, M.; Takahashi, H. Metabotropic glutamate receptor 5 is a coreceptor for alzheimer Aβ oligomer bound to cellular prion protein. Neuron 2013, 79, 887–902. [Google Scholar] [CrossRef] [Green Version]
  190. Rushworth, J.V.; Griffiths, H.H.; Watt, N.T.; Hooper, N.M. Prion Protein-mediated Toxicity of Amyloid–β Oligomers Requires Lipid Rafts and the Transmembrane LRP1. J. Biol. Chem. 2013, 288, 8935–8951. [Google Scholar] [CrossRef] [Green Version]
  191. Mattei, V.; Manganelli, V.; Martellucci, S.; Capozzi, A.; Mantuano, E.; Longo, A.; Ferri, A.; Garofalo, T.; Sorice, M.; Misasi, R. A multimolecular signaling complex including PrP(C) and LRP1 is strictly dependent on lipid rafts and is essential for the function of tissue plasminogen activator. J. Neurochem. 2020, 152, 468–481. [Google Scholar] [CrossRef] [PubMed]
  192. Laurén, J.; Gimbel, D.A.; Nygaard, H.B.; Gilbert, J.W.; Strittmatter, S.M. Cellular prion protein mediates impairment of synaptic plasticity by amyloid-β oligomers. Nature 2009, 457, 1128–1132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Kessels, H.W.; Nguyen, L.N.; Nabavi, S.; Malinow, R. The prion protein as a receptor for amyloid-β. Nature 2010, 466, E3–E5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Calella, A.M.; Farinelli, M.; Nuvolone, M.; Mirante, O.; Moos, R.; Falsig, J.; Mansuy, I.M.; Aguzzi, A. Prion protein and A-β-related synaptic toxicity impairment. EMBO Mol. Med. 2010, 2, 306–314. [Google Scholar] [CrossRef] [Green Version]
  195. Mengel, D.; Hong, W.; Corbett, G.T.; Liu, W.; DeSousa, A.; Solforosi, L.; Fang, C.; Frosch, M.P.; Collinge, J.; Harris, D.A.; et al. PrP-grafted antibodies bind certain amyloid β-protein aggregates, but do not prevent toxicity. Brain Res. 2019, 1710, 125–135. [Google Scholar] [CrossRef]
  196. Chen, S.; Yadav, S.P.; Surewicz, W.K. Interaction between human prion protein and amyloid-β (Aβ) oligomers: Role of N-terminal residues. J. Biol. Chem. 2010, 285, 26377–26383. [Google Scholar] [CrossRef] [Green Version]
  197. De Cecco, E.; Legname, G. The role of the prion protein in the internalization of α-synuclein amyloids. Prion 2018, 12, 23–27. [Google Scholar] [CrossRef] [Green Version]
  198. Gomes, L.A.; Hipp, S.A.; Upadhaya, A.R.; Balakrishnan, K.; Ospitalieri, S.; Koper, M.J.; Largo-Barrientos, P.; Uytterhoeven, V.; Reichwald, J.; Rabe, S. Aβ-induced acceleration of Alzheimer-related τ-pathology spreading and its association with prion protein. Acta Neuropathol. 2019, 138, 913–941. [Google Scholar] [CrossRef]
  199. Rösener, N.S.; Gremer, L.; Wördehoff, M.M.; Kupreichyk, T.; Etzkorn, M.; Neudecker, P.; Hoyer, W. Clustering of human prion protein and α-synuclein oligomers requires the prion protein N-terminus. Commun. Biol. 2020, 3, 365. [Google Scholar] [CrossRef]
  200. De Cecco, E.; Celauro, L.; Vanni, S.; Grandolfo, M.; Bistaffa, E.; Moda, F.; Aguzzi, A.; Legname, G. The uptake of tau amyloid fibrils is facilitated by the cellular prion protein and hampers prion propagation in cultured cells. J. Neurochem. 2020, 155, 577–591. [Google Scholar] [CrossRef]
  201. Resenberger, U.K.; Harmeier, A.; Woerner, A.C.; Goodman, J.L.; Muller, V.; Krishnan, R. The cellular prion protein mediates neurotoxic signalling of β-sheet-rich conformers independent of prion replication. EMBO J. 2011, 30, 2057–2070. [Google Scholar] [CrossRef] [PubMed]
  202. Nieznanski, K.; Choi, J.K.; Chen, S.; Surewicz, K.; Surewicz, W.K. Soluble prion protein inhibits amyloid-β (Aβ) fibrillization and toxicity. J. Biol. Chem. 2012, 287, 33104–33108. [Google Scholar] [CrossRef] [Green Version]
  203. Fluharty, B.R.; Biasini, E.; Stravalaci, M.; Sclip, A.; Diomede, L.; Balducci, C.; La Vitola, P.; Messa, M.; Colombo, L.; Forloni, G.; et al. An N-terminal Fragment of the Prion Protein Binds to Amyloid-β Oligomers and Inhibits Their Neurotoxicity in Vivo. J. Biol. Chem. 2013, 288, 7857–7866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Nieznanska, H.; Bandyszewska, M.; Surewicz, K.; Zajkowski, T.; Surewicz, W.K.; Nieznanski, K. Identification of prion protein-derived peptides of potential use in Alzheimer’s disease therapy. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 2143–2153. [Google Scholar] [CrossRef] [PubMed]
  205. Béland, M.; Bédard, M.; Tremblay, G.; Lavigne, P.; Roucou, X. Aβ induces its own prion protein N-terminal fragment (PrPN1)–mediated neutralization in amorphous aggregates. Neurobiol. Aging 2014, 35, 1537–1548. [Google Scholar] [CrossRef]
  206. Heiseke, A.; Schöbel, S.; Lichtenthaler, S.F.; Vorberg, I.; Groschup, M.H.; Kretzschmar, H.; Schätzl, H.M.; Nunziante, M. The Novel Sorting Nexin SNX33 Interferes with Cellular PrPSc Formation by Modulation of PrPc Shedding. Traffic 2008, 9, 1116–1129. [Google Scholar] [CrossRef] [PubMed]
  207. Kanaani, J.; Prusiner, S.B.; Diacovo, J.; Baekkeskov, S.; Legname, G. Recombinant prion protein induces rapid polarization and development of synapses in embryonic rat hippocampal neurons in vitro: Prion protein enhances neuronal polarization. J. Neurochem. 2005, 95, 1373–1386. [Google Scholar] [CrossRef]
  208. Bove-Fenderson, E.; Urano, R.; Straub, J.E.; Harris, D.A. Cellular prion protein targets amyloid-β fibril ends via its C-terminal domain to prevent elongation. J. Biol. Chem. 2017, 292, 16858–16871. [Google Scholar] [CrossRef] [Green Version]
  209. Roberts, T.K.; Eugenin, E.A.; Morgello, S.; Clements, J.E.; Zink, M.C.; Berman, J.W. PrP(C), the cellular isoform of the human prion protein, is a novel biomarker of HIV-associated neurocognitive impairment and mediates neuroinflammation. Am. J. Pathol. 2010, 177, 1848–1860. [Google Scholar] [CrossRef]
  210. Linsenmeier, L.; Mohammadi, B.; Wetzel, S.; Puig, B.; Jackson, W.S.; Hartmann, A.; Uchiyama, K.; Sakaguchi, S.; Endres, K.; Tatzelt, J.; et al. Structural and mechanistic aspects influencing the ADAM10-mediated shedding of the prion protein. Mol. Neurodegener. 2018, 13, 18. [Google Scholar] [CrossRef] [Green Version]
  211. Legname, G.; Scialò, C. On the role of the cellular prion protein in the uptake and signaling of pathological aggregates in neurodegenerative diseases. Prion 2020, 14, 257–270. [Google Scholar] [CrossRef] [PubMed]
  212. Scott-McKean, J.J.; Surewicz, K.; Choi, J.-K.; Ruffin, V.A.; Salameh, A.I.; Nieznanski, K.; Costa, A.C.; Surewicz, W.K. Soluble prion protein and its N-terminal fragment prevent impairment of synaptic plasticity by Aβ oligomers: Implications for novel therapeutic strategy in Alzheimer’s disease. Neurobiol. Dis. 2016, 91, 124–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Vella, L.J.; Greenwood, D.L.V.; Cappai, R.; Scheerlinck, J.-P.Y.; Hill, A.F. Enrichment of prion protein in exosomes derived from ovine cerebral spinal fluid. Vet. Immunol. Immunopathol. 2008, 124, 385–393. [Google Scholar] [CrossRef] [PubMed]
  214. Falker, C.; Hartmann, A.; Guett, I.; Dohler, F.; Altmeppen, H.; Betzel, C.; Schubert, R.; Thurm, D.; Wegwitz, F.; Joshi, P.; et al. Exosomal cellular prion protein drives fibrillization of amyloid beta and counteracts amyloid beta-mediated neurotoxicity. J. Neurochem. 2016, 137, 88–100. [Google Scholar] [CrossRef]
  215. Yuyama, K.; Sun, H.; Sakai, S.; Mitsutake, S.; Okada, M.; Tahara, H.; Furukawa, J.-i.; Fujitani, N.; Shinohara, Y.; Igarashi, Y. Decreased amyloid-β pathologies by intracerebral loading of glycosphingolipid-enriched exosomes in Alzheimer model mice. J. Biol. Chem. 2014, 289, 24488–24498. [Google Scholar] [CrossRef] [Green Version]
  216. Yuyama, K.; Sun, H.; Usuki, S.; Sakai, S.; Hanamatsu, H.; Mioka, T.; Kimura, N.; Okada, M.; Tahara, H.; Furukawa, J.-i. A potential function for neuronal exosomes: Sequestering intracerebral amyloid-β peptide. FEBS Lett. 2015, 589, 84–88. [Google Scholar] [CrossRef]
  217. An, K.; Klyubin, I.; Kim, Y.; Jung, J.H.; Mably, A.J.; T O’Dowd, S.; Lynch, T.; Kanmert, D.; Lemere, C.A.; Finan, G.M. Exosomes neutralize synaptic-plasticity-disrupting activity of Aβ assemblies in vivo. Mol. Brain 2013, 6, 1–13. [Google Scholar] [CrossRef] [Green Version]
  218. Cervenakova, L.; Saá, P.; Yakovleva, O.; Vasilyeva, I.; de Castro, J.; Brown, P.; Dodd, R. Are prions transported by plasma exosomes? Transfus. Apher. Sci. 2016, 55, 70–83. [Google Scholar] [CrossRef]
  219. Vella, L.; Sharples, R.; Lawson, V.; Masters, C.; Cappai, R.; Hill, A. Packaging of prions into exosomes is associated with a novel pathway of PrP processing. J. Pathol. A J. Pathol. Soc. Great Br. Irel. 2007, 211, 582–590. [Google Scholar] [CrossRef]
  220. Alves, R.N.; Iglesia, R.P.; Prado, M.B.; Melo Escobar, M.I.; Boccacino, J.M.; Fernandes, C.F.d.L.; Coelho, B.P.; Fortes, A.C.; Lopes, M.H. A New Take on Prion Protein Dynamics in Cellular Trafficking. Int. J. Mol. Sci. 2020, 21, 7763. [Google Scholar] [CrossRef]
  221. Leblanc, P.; Arellano-Anaya, Z.E.; Bernard, E.; Gallay, L.; Provansal, M.; Lehmann, S.; Schaeffer, L.; Raposo, G.; Vilette, D. Isolation of Exosomes and Microvesicles from Cell Culture Systems to Study Prion Transmission. In Exosomes and Microvesicles: Methods and Protocols; Hill, A.F., Ed.; Springer: New York, NY, USA, 2017; pp. 153–176. [Google Scholar]
  222. Fevrier, B.; Vilette, D.; Archer, F.; Loew, D.; Faigle, W.; Vidal, M.; Laude, H.; Raposo, G. Cells release prions in association with exosomes. Proc. Natl. Acad. Sci. USA 2004, 101, 9683–9688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Rösener, N.S.; Gremer, L.; Reinartz, E.; König, A.; Brener, O.; Heise, H.; Hoyer, W.; Neudecker, P.; Willbold, D. A d-enantiomeric peptide interferes with heteroassociation of amyloid-β oligomers and prion protein. J. Biol. Chem. 2018, 293, 15748–15764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Schematic presentation of PrPC with associated cleavages. Mature PrPC is approximately 210 amino acids long. The flexible unstructured N-terminal part (residues 23–120) contains the octapeptide repeat region (OR, purple) whereas the highly structured C-terminal part (residues 121–231) is composed of three α-helices (green), two β-sheets (orange), a disulfide bond, two N-glycans (CHO; positions 181 and 197) and a C-terminal GPI anchor. PrP can undergo four cleavages: α-cleavage (cleavage site position 111/112); β-cleavage (cleavage site position 89/90); γ-cleavage (cleavage site presumably between positions 170–120); and shedding (near the C-terminus of PrP). Cleavages result in released (N1, N2, N3, shed PrP) and attached (C1, C2, C3) fragments of PrPC.
Figure 1. Schematic presentation of PrPC with associated cleavages. Mature PrPC is approximately 210 amino acids long. The flexible unstructured N-terminal part (residues 23–120) contains the octapeptide repeat region (OR, purple) whereas the highly structured C-terminal part (residues 121–231) is composed of three α-helices (green), two β-sheets (orange), a disulfide bond, two N-glycans (CHO; positions 181 and 197) and a C-terminal GPI anchor. PrP can undergo four cleavages: α-cleavage (cleavage site position 111/112); β-cleavage (cleavage site position 89/90); γ-cleavage (cleavage site presumably between positions 170–120); and shedding (near the C-terminus of PrP). Cleavages result in released (N1, N2, N3, shed PrP) and attached (C1, C2, C3) fragments of PrPC.
Ijms 23 01232 g001
Figure 2. Proteins, signaling pathways and interactions that may be affected by PrP and/or PrP fragments. This scheme presents various proteins, signaling pathways and interactions that reportedly involve PrP and/or its fragments. In ischemic stroke, PrP species were found to be involved in modulating neuroprotection, neurite outgrowth, neurogenesis and angiogenesis. In neurodegenerative diseases, released PrP fragments may act protectively whereas anchored PrP regulates oligomer-induced toxicity. PrP and its derivatives are also involved in Adgrg6-induced myelination homeostasis (orange) and may be involved in microglia communication and differentiation as well as regulating intercellular communication through EVs and sEVs, etc. Several of the proposed interplays are regulated by a direct interaction with PrP species whereas others are regulated indirectly. Protective pathways and interactions are colored blue whereas green color presents harmful outcomes.
Figure 2. Proteins, signaling pathways and interactions that may be affected by PrP and/or PrP fragments. This scheme presents various proteins, signaling pathways and interactions that reportedly involve PrP and/or its fragments. In ischemic stroke, PrP species were found to be involved in modulating neuroprotection, neurite outgrowth, neurogenesis and angiogenesis. In neurodegenerative diseases, released PrP fragments may act protectively whereas anchored PrP regulates oligomer-induced toxicity. PrP and its derivatives are also involved in Adgrg6-induced myelination homeostasis (orange) and may be involved in microglia communication and differentiation as well as regulating intercellular communication through EVs and sEVs, etc. Several of the proposed interplays are regulated by a direct interaction with PrP species whereas others are regulated indirectly. Protective pathways and interactions are colored blue whereas green color presents harmful outcomes.
Ijms 23 01232 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kovač, V.; Čurin Šerbec, V. Prion Protein: The Molecule of Many Forms and Faces. Int. J. Mol. Sci. 2022, 23, 1232. https://doi.org/10.3390/ijms23031232

AMA Style

Kovač V, Čurin Šerbec V. Prion Protein: The Molecule of Many Forms and Faces. International Journal of Molecular Sciences. 2022; 23(3):1232. https://doi.org/10.3390/ijms23031232

Chicago/Turabian Style

Kovač, Valerija, and Vladka Čurin Šerbec. 2022. "Prion Protein: The Molecule of Many Forms and Faces" International Journal of Molecular Sciences 23, no. 3: 1232. https://doi.org/10.3390/ijms23031232

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