Major prion protein

Last updated
PRNP
PRNP .png
Available structures
PDB Ortholog search: PDBe RCSB
Identifiers
Aliases PRNP , ASCR, AltPrP, CD230, CJD, GSS, KURU, PRIP, PrP, PrP27-30, PrP33-35C, PrPc, p27-30, prion protein
External IDs OMIM: 176640 MGI: 97769 HomoloGene: 7904 GeneCards: PRNP
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_001278256
NM_011170

RefSeq (protein)

NP_001265185
NP_035300

Location (UCSC) Chr 20: 4.69 – 4.7 Mb Chr 2: 131.75 – 131.78 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse

Major prion protein (PrP) is encoded in the human body by the PRNP gene also known as CD230 (cluster of differentiation 230). [5] [6] [7] [8] Expression of the protein is most predominant in the nervous system but occurs in many other tissues throughout the body. [9] [10] [11]

The protein can exist in multiple isoforms: the normal PrPC form, and the protease-resistant form designated PrPRes such as the disease-causing PrPSc (scrapie) and an isoform located in mitochondria. The misfolded version PrPSc is associated with a variety of cognitive disorders and neurodegenerative diseases such as in animals: ovine scrapie, bovine spongiform encephalopathy (BSE, mad cow disease), feline spongiform encephalopathy, transmissible mink encephalopathy (TME), exotic ungulate encephalopathy, chronic wasting disease (CWD) which affects deer; and in humans: Creutzfeldt–Jakob disease (CJD), fatal familial insomnia (FFI), Gerstmann–Sträussler–Scheinker syndrome (GSS), kuru, and variant Creutzfeldt–Jakob disease (vCJD). Similarities exist between kuru, thought to be due to human ingestion of diseased individuals, and vCJD, thought to be due to human ingestion of BSE-tainted cattle products.

Gene

Chromosome 20 Location of PRNP-gene in chromosome 20.svg
Chromosome 20

The human PRNP gene is located on the short (p) arm of chromosome 20 between the end (terminus) of the arm and position 13, from base pair 4,615,068 to base pair 4,630,233.

Structure

PrP is highly conserved through mammals, lending credence to application of conclusions from test animals such as mice. [12] Comparison between primates is especially similar, ranging from 92.9-99.6% similarity in amino acid sequences. The human protein structure consists of a globular domain with three α-helices and a two-strand antiparallel β-sheet, an NH2-terminal tail, and a short COOH-terminal tail. [13] A glycophosphatidylinositol (GPI) membrane anchor at the COOH-terminal tethers PrP to cell membranes, and this proves to be integral to the transmission of conformational change; secreted PrP lacking the anchor component is unaffected by the infectious isoform. [14]

The primary sequence of PrP is 253 amino acids long before post-translational modification. Signal sequences in the amino- and carboxy- terminal ends are removed posttranslationally, resulting in a mature length of 208 amino acids. For human and golden hamster PrP, two glycosylated sites exist on helices 2 and 3 at Asn181 and Asn197. Murine PrP has glycosylation sites as Asn180 and Asn196. A disulfide bond exists between Cys179 of the second helix and Cys214 of the third helix (human PrPC numbering).

PrP messenger RNA contains a pseudoknot structure (prion pseudoknot), which is thought to be involved in regulation of PrP protein translation. [15]

Ligand-binding

The mechanism for conformational conversion to the scrapie isoform is speculated to be an elusive ligand-protein, but, so far, no such compound has been identified. However, a large body of research has developed on candidates and their interaction with the PrPC. [16]

Copper, zinc, manganese, and nickel are confirmed PrP ligands that bind to its octarepeat region. [17] Ligand binding causes a conformational change with unknown effect. Heavy metal binding at PrP has been linked to resistance to oxidative stress arising from heavy metal toxicity. [17] [18]

PrPC (normal cellular) isoform

Although the precise function of PrP is not yet known, it is possibly involved in the transport of ionic copper to cells from the surrounding environment. Researchers have also proposed roles for PrP in cell signaling or in the formation of synapses. [19] PrPC attaches to the outer surface of the cell membrane by a glycosylphosphatidylinositol anchor at its C-terminal Ser231.

Prion protein contains five octapeptide repeats with sequence PHGGGWGQ (though the first repeat has the slightly-modified, histidine-deficient sequence PQGGGGWGQ). This is thought to generate a copper-binding domain via nitrogen atoms in the histidine imidazole side-chains and deprotonated amide nitrogens from the 2nd and 3rd glycines in the repeat. The ability to bind copper is, therefore, pH-dependent. NMR shows copper binding results in a conformational change at the N-terminus.

PrPSc (scrapie) isoform

PrPSc is a conformational isoform of PrPC, but this orientation tends to accumulate in compact, protease-resistant aggregates within neural tissue. [20] The abnormal PrPSc isoform has a different secondary and tertiary structure from PrPC, but identical primary sequence. Whereas PrPC has largely alpha helical and disordered domains [21] , PrPSc has no alpha helix and an amyloid fibril core composed of a stack of PrP molecules glued together by parallel in-register intermolecular beta sheets. [22] [23] [24] This refolding renders the PrPSc isoform extremely resistant to proteolysis.

The propagation of PrPSc is a topic of great interest, as its accumulation is a pathological cause of neurodegeneration. Based on the progressive nature of spongiform encephalopathies, the predominant hypothesis posits that the change from normal PrPC is caused by the presence and interaction with PrPSc. [25] Strong support for this is taken from studies in which PRNP-knockout mice are resistant to the introduction of PrPSc. [26] Despite widespread acceptance of the conformation conversion hypothesis, some studies mitigate claims for a direct link between PrPSc and cytotoxicity. [27]

Polymorphisms at sites 136, 154, and 171 are associated with varying susceptibility to ovine scrapie. (These ovine sites correspond to human sites 133, 151, and 168.) Polymorphisms of the PrP-VRQ form and PrP-ARQ form are associated with increased susceptibility, whereas PrP-ARR is associated with resistance. The National Scrapie Plan of the UK aims to breed out these scrapie polymorphisms by increasing the frequency of the resistant allele. [28] However, PrP-ARR polymorphisms are susceptible to atypical scrapie, so this may prove unfruitful.

Function

Nervous system

The strong association to neurodegenerative diseases raises many questions of the function of PrP in the brain. A common approach is using PrP-knockout and transgenic mice to investigate deficiencies and differences. [29] Initial attempts produced two strains of PrP-null mice that show no physiological or developmental differences when subjected to an array of tests. However, more recent strains have shown significant cognitive abnormalities. [16]

As the null mice age, a marked loss of Purkinje cells in the cerebellum results in decreased motor coordination. However, this effect is not a direct result of PrP's absence, and rather arises from increased Doppel gene expression. [30] Other observed differences include reduced stress response and increased exploration of novel environments. [31] [32]

Circadian rhythm is altered in null mice. [11] Fatal familial insomnia is thought to be the result of a point mutation in PRNP at codon 178, which corroborates PrP's involvement in sleep-wake cycles. [33] In addition, circadian regulation has been demonstrated in PrP mRNA, which cycles regularly with day-night. [34]

Memory

While null mice exhibit normal learning ability and short-term memory, long-term memory consolidation deficits have been demonstrated. As with ataxia, this is attributable to Doppel gene expression. However, spatial learning, a predominantly hippocampal-function, is decreased in the null mice and can be recovered with the reinstatement of PrP in neurons; this indicates that loss of PrP function is the cause. [35] [36] The interaction of hippocampal PrP with laminin (LN) is pivotal in memory processing and is likely modulated by the kinases PKA and ERK1/2. [37] [38]

Further support for PrP's role in memory formation is derived from several population studies. A test of healthy young humans showed increased long-term memory ability associated with an MM or MV genotype when compared to VV. [39] Down syndrome patients with a single valine substitution have been linked to earlier cognitive decline. [40] Several polymorphisms in PRNP have been linked with cognitive impairment in the elderly as well as earlier cognitive decline. [41] [42] [43] All of these studies investigated differences in codon 129, indicating its importance in the overall functionality of PrP, in particular with regard to memory.

Neurons and synapses

PrP is present in both the pre- and post-synaptic compartments, with the greatest concentration in the pre-synaptic portion. [44] Considering this and PrP's suite of behavioral influences, the neural cell functions and interactions are of particular interest. Based on the copper ligand, one proposed function casts PrP as a copper buffer for the synaptic cleft. In this role, the protein could serve as either a copper homeostasis mechanism, a calcium modulator, or a sensor for copper or oxidative stress. [45] Loss of PrP function has been linked to long-term potentiation (LTP). This effect can be positive or negative and is due to changes in neuronal excitability and synaptic transmission in the hippocampus. [46] [47]

Some research indicates PrP involvement in neuronal development, differentiation, and neurite outgrowth. The PrP-activated signal transduction pathway is associated with axon and dendritic outgrowth with a series of kinases. [27] [48]

Immune system

Though most attention is focused on PrP's presence in the nervous system, it is also abundant in immune system tissue. PrP immune cells include hematopoietic stem cells, mature lymphoid and myeloid compartments, and certain lymphocytes; also, it has been detected in natural killer cells, platelets, and monocytes. T cell activation is accompanied by a strong up-regulation of PrP, though it is not requisite. The lack of immunoresponse to transmissible spongiform encephalopathies (TSE), neurodegenerative diseases caused by prions, could stem from the tolerance for PrPSc. [49]

Muscles, liver, and pituitary

PrP-null mice provide clues to a role in muscular physiology when subjected to a forced swimming test, which showed reduced locomotor activity. Aging mice with an overexpression of PRNP showed significant degradation of muscle tissue.

Though present, very low levels of PrP exist in the liver and could be associated with liver fibrosis. Presence in the pituitary has been shown to affect neuroendocrine function in amphibians, but little is known concerning mammalian pituitary PrP. [16]

Cellular

Varying expression of PrP through the cell cycle has led to speculation on involvement in development. A wide range of studies has been conducted investigating the role in cell proliferation, differentiation, death, and survival. [16] Engagement of PrP has been linked to activation of signal transduction.

Modulation of signal transduction pathways has been demonstrated in cross-linking with antibodies and ligand-binding (hop/STI1 or copper). [16] Given the diversity of interactions, effects, and distribution, PrP has been proposed as dynamic surface protein functioning in signaling pathways. Specific sites along the protein bind other proteins, biomolecules, and metals. These interfaces allow specific sets of cells to communicate based on level of expression and the surrounding microenvironment. The anchoring on a GPI raft in the lipid bilayer supports claims of an extracellular scaffolding function. [16]

Diseases caused by PrP misfolding

More than 20 mutations in the PRNP gene have been identified in people with inherited prion diseases, which include the following: [50] [51]

The conversion of PrPC to PrPSc conformation is the mechanism of transmission of fatal, neurodegenerative transmissible spongiform encephalopathies (TSE). This can arise from genetic factors, infection from external source, or spontaneously for reasons unknown. Accumulation of PrPSc corresponds with progression of neurodegeneration and is the proposed cause. Some PRNP mutations lead to a change in single amino acids (the building-blocks of proteins) in the prion protein. Others insert additional amino acids into the protein or cause an abnormally short protein to be made. These mutations cause the cell to make prion proteins with an abnormal structure. The abnormal protein PrPSc accumulates in the brain and destroys nerve cells, which leads to the mental and behavioral features of prion diseases.

Several other changes in the PRNP gene (called polymorphisms) do not cause prion diseases but may affect a person's risk of developing these diseases or alter the course of the disorders. An allele that codes for a PRNP variant, G127V, provides resistance to kuru. [54]

In addition, some prion diseases can be transmitted from external sources of PrPSc. [55]

Alzheimer's disease

PrPC protein is one of several cellular receptors of soluble amyloid beta (Aβ) oligomers, which are canonically implicated in causing Alzheimer's disease. [56] These oligomers are composed of smaller Aβ plaques, and are the most damaging to the integrity of a neuron. [56] The precise mechanism of soluble Aβ oligomers directly inducing neurotoxicity is unknown, and experimental deletion of PRNP in animals has yielded several conflicting findings. When Aβ oligomers were injected into the cerebral ventricles of a mouse model of Alzheimer's, PRNP deletion did not offer protection, only anti-PrPC antibodies prevented long-term memory and spatial learning deficits. [57] [58] This would suggest either an unequal relation between PRNP and Aβ oligomer-mediated neurodegeneration or a site-specific relational significance. In the case of direct injection of Aβ oligomers into the hippocampus, PRNP-knockout mice were found to be indistinguishable from control with respect to both neuronal death rates and measurements of synaptic plasticity. [56] [58] It was further found that Aβ-oligomers bind to PrPC at the postsynaptic density, indirectly overactivating the NMDA receptor via the Fyn enzyme, resulting in excitotoxicity. [57] Soluble Aβ oligomers also bind to PrPC at the dendritic spines, forming a complex with Fyn and excessively activating tau, another protein implicated in Alzheimer's. [57] As the gene FYN codes for the enzyme Fyn, FYN-knockout mice display neither excitotoxic events nor dendritic spine shrinkage when injected with Aβ oligomers. [57] In mammals, the full functional significance of PRNP remains unclear, as PRNP deletion has been prophylactically implemented by the cattle industry without apparent harm. [56] In mice, this same deletion phenotypically varies between Alzheimer's mouse lines, as hAPPJ20 mice and TgCRND8 mice show a slight increase in epileptic activity, contributing to conflicting results when examining Alzheimer's survival rates. [56] Of note, the deletion of PRNP in both APPswe and SEN1dE9, two other transgenic models of Alzheimer's, attenuated the epilepsy-induced death phenotype seen in a subset of these animals. [56] Taken collectively, recent evidence suggests PRNP may be important for conducing the neurotoxic effects of soluble Aβ-oligomers and the emergent disease state of Alzheimer's. [56] [57] [58]

In humans, the methionine/valine polymorphism at codon 129 of PRNP (rs1799990) is most closely associated with Alzheimer's disease. [59] Variant V allele carriers (VV and MV) show a 13% decreased risk with respect to developing Alzheimer's compared to the methionine homozygote (MM). However, the protective effects of variant V carriers have been found exclusively in Caucasians. The decreased risk in V allele carriers is further limited to late-onset Alzheimer's disease only (≥ 65 years). [59] PRNP can also functionally interact with polymorphisms in two other genes implicated in Alzheimer's, PSEN1 and APOE, to compound risk for both Alzheimer's and sporadic Creutzfeldt–Jakob disease. [56] A point mutation on codon 102 of PRNP at least in part contributed to three separate patients' atypical frontotemporal dementia within the same family, suggesting a new phenotype for Gerstmann–Sträussler–Scheinker syndrome. [56] [60] The same study proposed sequencing PRNP in cases of ambiguously diagnosed dementia, as the various forms of dementia can prove challenging to differentially diagnose. [60]

Research

In 2006 the production of cattle lacking PrPC form of the major prion protein (PrP) protein was reported which were resistant to prion propagation with no apparent developmental abnormalities. Besides the study of bovine products free of prion proteins another use could be so that human pharmaceuticals can be made in their blood without the danger that those products might get contaminated with the infectious agent that causes mad cow. [61] [62]

Interactions

A strong interaction exists between PrP and the cochaperone Hop (Hsp70/Hsp90 organizing protein; also called STI1 (Stress-induced protein 1)). [63] [64]

Related Research Articles

<span class="mw-page-title-main">Creutzfeldt–Jakob disease</span> Degenerative neurological disorder

Creutzfeldt–Jakob disease (CJD), also known as subacute spongiform encephalopathy or neurocognitive disorder due to prion disease, is a fatal degenerative brain disorder. Early symptoms include memory problems, behavioral changes, poor coordination, and visual disturbances. Later symptoms include dementia, involuntary movements, blindness, weakness, and coma. About 70% of people die within a year of diagnosis. The name Creutzfeldt–Jakob disease was introduced by Walther Spielmeyer in 1922, after the German neurologists Hans Gerhard Creutzfeldt and Alfons Maria Jakob.

<span class="mw-page-title-main">Prion</span> Pathogenic type of misfolded protein

A prion is a misfolded protein that can induce misfolding of normal variants of the same protein and trigger cellular death. Prions cause prion diseases known as transmissible spongiform encephalopathies (TSEs) that are transmissible, fatal neurodegenerative diseases in humans and animals. The proteins may misfold sporadically, due to genetic mutations, or by exposure to an already misfolded protein. The consequent abnormal three-dimensional structure confers on them the ability to cause misfolding of other proteins.

<span class="mw-page-title-main">Fatal insomnia</span> Prion disease of the human brain

Fatal insomnia is an extremely rare neurodegenerative prion disease that results in trouble sleeping as its hallmark symptom. The majority of cases are familial, stemming from a mutation in the PRNP gene, with the remainder of cases occurring sporadically. The problems with sleeping typically start out gradually and worsen over time. Eventually, the patient will succumb to total insomnia, most often leading to other symptoms such as speech problems, coordination problems, and dementia. It results in death within a few months to a few years and has no known cure.

<span class="mw-page-title-main">Scrapie</span> Degenerative disease that affects sheep and goats

Scrapie is a fatal, degenerative disease affecting the nervous systems of sheep and goats. It is one of several transmissible spongiform encephalopathies (TSEs), and as such it is thought to be caused by a prion. Scrapie has been known since at least 1732 and does not appear to be transmissible to humans. However, it has been found to be experimentally transmissible to humanised transgenic mice and non-human primates.

Transmissible spongiform encephalopathies (TSEs) also known as prion diseases, are a group of progressive, incurable, and fatal conditions that are associated with prions and affect the brain and nervous system of many animals, including humans, cattle, and sheep. According to the most widespread hypothesis, they are transmitted by prions, though some other data suggest an involvement of a Spiroplasma infection. Mental and physical abilities deteriorate and many tiny holes appear in the cortex causing it to appear like a sponge when brain tissue obtained at autopsy is examined under a microscope. The disorders cause impairment of brain function, including memory changes, personality changes and problems with movement that worsen chronically.

<span class="mw-page-title-main">Chronic wasting disease</span> Prion disease affecting the deer family

Chronic wasting disease (CWD), sometimes called zombie deer disease, is a transmissible spongiform encephalopathy (TSE) affecting deer. TSEs are a family of diseases thought to be caused by misfolded proteins called prions and include similar diseases such as BSE in cattle, Creutzfeldt–Jakob disease (CJD) in humans and scrapie in sheep. Natural infection causing CWD affects members of the deer family. In the United States, CWD affects mule deer, white-tailed deer, red deer, sika deer, elk, caribou, and moose. Experimental transmission of CWD to other species such as squirrel monkeys and genetically modified mice has been shown.

<span class="mw-page-title-main">Gerstmann–Sträussler–Scheinker syndrome</span> Human neurodegenerative disease

Gerstmann–Sträussler–Scheinker syndrome (GSS) is an extremely rare, always fatal neurodegenerative disease that affects patients from 20 to 60 years in age. It is exclusively heritable, and is found in only a few families all over the world. It is, however, classified with the transmissible spongiform encephalopathies (TSE) due to the causative role played by PRNP, the human prion protein. GSS was first reported by the Austrian physicians Josef Gerstmann, Ernst Sträussler and Ilya Scheinker in 1936.

<span class="mw-page-title-main">Amyloid beta</span> Group of peptides

Amyloid beta denotes peptides of 36–43 amino acids that are the main component of the amyloid plaques found in the brains of people with Alzheimer's disease. The peptides derive from the amyloid-beta precursor protein (APP), which is cleaved by beta secretase and gamma secretase to yield Aβ in a cholesterol-dependent process and substrate presentation. Aβ molecules can aggregate to form flexible soluble oligomers which may exist in several forms. It is now believed that certain misfolded oligomers can induce other Aβ molecules to also take the misfolded oligomeric form, leading to a chain reaction akin to a prion infection. The oligomers are toxic to nerve cells. The other protein implicated in Alzheimer's disease, tau protein, also forms such prion-like misfolded oligomers, and there is some evidence that misfolded Aβ can induce tau to misfold.

<span class="mw-page-title-main">Amyloid plaques</span> Extracellular deposits of the amyloid beta protein

Amyloid plaques are extracellular deposits of the amyloid beta (Aβ) protein mainly in the grey matter of the brain. Degenerative neuronal elements and an abundance of microglia and astrocytes can be associated with amyloid plaques. Some plaques occur in the brain as a result of aging, but large numbers of plaques and neurofibrillary tangles are characteristic features of Alzheimer's disease. The plaques are highly variable in shape and size; in tissue sections immunostained for Aβ, they comprise a log-normal size distribution curve, with an average plaque area of 400-450 square micrometers (µm²). The smallest plaques, which often consist of diffuse deposits of Aβ, are particularly numerous. Plaques form when Aβ misfolds and aggregates into oligomers and longer polymers, the latter of which are characteristic of amyloid.

Protein misfolding cyclic amplification (PMCA) is an amplification technique to multiply misfolded prions originally developed by Soto and colleagues. It is a test for spongiform encephalopathies like chronic wasting disease (CWD) or bovine spongiform encephalopathy (BSE).

Karen K. Hsiao Ashe is a professor at the Department of Neurology and Neuroscience at the University of Minnesota (UMN) Medical School, where she holds the Edmund Wallace and Anne Marie Tulloch Chairs in Neurology and Neuroscience. She is the founding director of the N. Bud Grossman Center for Memory Research and Care, and her specific research interest is memory loss resulting from Alzheimer's disease and related dementias. Her research has included the development of an animal model of Alzheimer's.

The biochemistry of Alzheimer's disease, the most common cause of dementia, is not yet very well understood. Alzheimer's disease (AD) has been identified as a proteopathy: a protein misfolding disease due to the accumulation of abnormally folded amyloid beta (Aβ) protein in the brain. Amyloid beta is a short peptide that is an abnormal proteolytic byproduct of the transmembrane protein amyloid-beta precursor protein (APP), whose function is unclear but thought to be involved in neuronal development. The presenilins are components of proteolytic complex involved in APP processing and degradation.

Laura Manuelidis is a physician and neuropathologist at Yale University.

<span class="mw-page-title-main">Proteinopathy</span> Medical condition

In medicine, proteinopathy, or proteopathy, protein conformational disorder, or protein misfolding disease, is a class of diseases in which certain proteins become structurally abnormal, and thereby disrupt the function of cells, tissues and organs of the body. Often the proteins fail to fold into their normal configuration; in this misfolded state, the proteins can become toxic in some way or they can lose their normal function. The proteinopathies include such diseases as Creutzfeldt–Jakob disease and other prion diseases, Alzheimer's disease, Parkinson's disease, amyloidosis, multiple system atrophy, and a wide range of other disorders. The term proteopathy was first proposed in 2000 by Lary Walker and Harry LeVine.

<span class="mw-page-title-main">Kuru (disease)</span> Rare neurodegenerative disease caused by prions

Kuru is a rare, incurable, and fatal neurodegenerative disorder that was formerly common among the Fore people of Papua New Guinea. Kuru is a form of transmissible spongiform encephalopathy (TSE) caused by the transmission of abnormally folded proteins (prions), which leads to symptoms such as tremors and loss of coordination from neurodegeneration.

<span class="mw-page-title-main">PRND</span> Protein-coding gene in the species Homo sapiens

Prion protein 2 (dublet), also known as PRND, or Doppel protein, is a protein which in humans is encoded by the PRND gene.

<span class="mw-page-title-main">Bovine spongiform encephalopathy</span> Fatal neurodegenerative disease of cattle

Bovine spongiform encephalopathy (BSE), commonly known as mad cow disease, is an incurable and invariably fatal neurodegenerative disease of cattle. Symptoms include abnormal behavior, trouble walking, and weight loss. Later in the course of the disease the cow becomes unable to function normally. There is conflicting information about the time between infection and onset of symptoms. In 2002, the World Health Organization (WHO) suggested it to be approximately four to five years. Time from onset of symptoms to death is generally weeks to months. Spread to humans is believed to result in variant Creutzfeldt–Jakob disease (vCJD). As of 2018, a total of 231 cases of vCJD had been reported globally.

<span class="mw-page-title-main">Surround optical-fiber immunoassay</span>

Surround optical-fiber immunoassay (SOFIA) is an ultrasensitive, in vitro diagnostic platform incorporating a surround optical-fiber assembly that captures fluorescence emissions from an entire sample. The technology's defining characteristics are its extremely high limit of detection, sensitivity, and dynamic range. SOFIA's sensitivity is measured at the attogram level (10−18 g), making it about one billion times more sensitive than conventional diagnostic techniques. Based on its enhanced dynamic range, SOFIA is able to discriminate levels of analyte in a sample over 10 orders of magnitude, facilitating accurate titering.

Frank O. Bastian is an American medical doctor and neuropathologist, who previously worked at Louisiana State University, moved to a university in New Orleans in 2019. He specializes in the transmissible spongiform encephalopathies (TSEs), which include, but are not limited to, Bovine spongiform encephalopathy (BSE) "Mad cow disease" in cattle, scrapie in sheep and goats, and Creutzfeldt–Jakob disease (CJD) in humans.

Michael Coulthart is a Canadian microbiologist who is employed as the head of the Canadian Creutzfeldt–Jakob Disease Surveillance System (CJDSS) within the Public Health Agency of Canada (PHAC), which terms CJD a zoonotic and infectious disease. In 2006, a working group named "classic CJD" as well as Variant Creutzfeldt–Jakob disease as two notifiable diseases. It is unknown whether PHAC tracks in an official capacity other transmissible spongiform encephalopathies (TSE), but Coulthart is on the Advisory Committee of the Center for Infectious Disease Research and Policy for Chronic Wasting Disease of cervidae.

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000171867 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000079037 - Ensembl, May 2017
  3. "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. Kretzschmar HA, Stowring LE, Westaway D, Stubblebine WH, Prusiner SB, Dearmond SJ (August 1986). "Molecular cloning of a human prion protein cDNA". DNA. 5 (4): 315–24. doi:10.1089/dna.1986.5.315. PMID   3755672.
  6. Sparkes RS, Simon M, Cohn VH, Fournier RE, Lem J, Klisak I, Heinzmann C, Blatt C, Lucero M, Mohandas T (October 1986). "Assignment of the human and mouse prion protein genes to homologous chromosomes". Proc. Natl. Acad. Sci. U.S.A. 83 (19): 7358–62. Bibcode:1986PNAS...83.7358S. doi: 10.1073/pnas.83.19.7358 . PMC   386716 . PMID   3094007.
  7. Liao YC, Lebo RV, Clawson GA, Smuckler EA (July 1986). "Human prion protein cDNA: molecular cloning, chromosomal mapping, and biological implications". Science. 233 (4761): 364–7. Bibcode:1986Sci...233..364L. doi:10.1126/science.3014653. PMID   3014653.
  8. Robakis NK, Devine-Gage EA, Jenkins EC, Kascsak RJ, Brown WT, Krawczun MS, Silverman WP (October 1986). "Localization of a human gene homologous to the PrP gene on the p arm of chromosome 20 and detection of PrP-related antigens in normal human brain". Biochem. Biophys. Res. Commun. 140 (2): 758–65. doi:10.1016/0006-291X(86)90796-5. PMID   2877664.
  9. Prusiner SB (2001). "Shattuck lecture--neurodegenerative diseases and prions". N Engl J Med. 344 (20): 1516–26. doi: 10.1056/NEJM200105173442006 . PMID   11357156.
  10. Weissmann C (2004). "The state of the prion". Nat Rev Microbiol. 2 (11): 861–71. doi:10.1038/nrmicro1025. PMID   15494743. S2CID   20992257.
  11. 1 2 Zomosa-Signoret V, Arnaud JD, Fontes P, Alvarez-Martinez MT, Liautard JP (2008). "Physiological role of the cellular prion protein" (PDF). Vet. Res. 39 (4): 9. doi: 10.1051/vetres:2007048 . PMID   18073096.
  12. Damberger FF, Christen B, Pérez DR, Hornemann S, Wüthrich K (October 2011). "Cellular prion protein conformation and function". Proc. Natl. Acad. Sci. U.S.A. 108 (42): 17308–13. Bibcode:2011PNAS..10817308D. doi: 10.1073/pnas.1106325108 . PMC   3198368 . PMID   21987789.
  13. Schätzl HM, Da Costa M, Taylor L, Cohen FE, Prusiner SB (January 1995). "Prion protein gene variation among primates". J. Mol. Biol. 245 (4): 362–74. doi:10.1006/jmbi.1994.0030. PMID   7837269.
  14. Chesebro B, Trifilo M, Race R, Meade-White K, Teng C, LaCasse R, Raymond L, Favara C, Baron G, Priola S, Caughey B, Masliah E, Oldstone M (June 2005). "Anchorless prion protein results in infectious amyloid disease without clinical scrapie". Science. 308 (5727): 1435–9. Bibcode:2005Sci...308.1435C. CiteSeerX   10.1.1.401.781 . doi:10.1126/science.1110837. PMID   15933194. S2CID   10064966.
  15. Barrette I, Poisson G, Gendron P, Major F (2001). "Pseudoknots in prion protein mRNAs confirmed by comparative sequence analysis and pattern searching". Nucleic Acids Res. 29 (3): 753–758. doi:10.1093/nar/29.3.753. PMC   30388 . PMID   11160898.
  16. 1 2 3 4 5 6 Linden R, Martins VR, Prado MA, Cammarota M, Izquierdo I, Brentani RR (April 2008). "Physiology of the prion protein". Physiol. Rev. 88 (2): 673–728. doi:10.1152/physrev.00007.2007. PMID   18391177.
  17. 1 2 Prčina M, Kontseková E, Novák M (2015). "Prion protein prevents heavy metals overloading of cells and thus protects them against their toxicity". Acta Virol. 59 (2): 179–84. doi: 10.4149/av_2015_02_179 . PMID   26104335.
  18. Brown DR, Clive C, Haswell SJ (January 2001). "Antioxidant activity related to copper binding of native prion protein". J. Neurochem. 76 (1): 69–76. doi: 10.1046/j.1471-4159.2001.00009.x . PMID   11145979. S2CID   45647133.
  19. Kanaani J, Prusiner SB, Diacovo J, Baekkeskov S, Legname G (December 2005). "Recombinant prion protein induces rapid polarization and development of synapses in embryonic rat hippocampal neurons in vitro". Journal of Neurochemistry. 95 (5): 1373–86. doi: 10.1111/j.1471-4159.2005.03469.x . PMID   16313516. S2CID   24329326.
  20. Ross CA, Poirier MA (July 2004). "Protein aggregation and neurodegenerative disease". Nat. Med. 10 (7): S10–7. doi:10.1038/nm1066. PMID   15272267. S2CID   205383483.
  21. Riek R, Hornemann S, Wider G, Glockshuber R, Wüthrich K (1997-08-18). "NMR characterization of the full-length recombinant murine prion protein, m PrP(23–231)". FEBS Letters. 413 (2): 282–288. doi:10.1016/S0014-5793(97)00920-4. ISSN   0014-5793. PMID   9280298. S2CID   39791520.
  22. Kraus A, Hoyt F, Schwartz CL, Hansen B, Artikis E, Hughson AG, Raymond GJ, Race B, Baron GS, Caughey B (November 2021). "High-resolution structure and strain comparison of infectious mammalian prions". Molecular Cell. 81 (21): 4540–4551.e6. doi:10.1016/j.molcel.2021.08.011. PMID   34433091.
  23. Manka SW, Zhang W, Wenborn A, Betts J, Joiner S, Saibil HR, Collinge J, Wadsworth JD (2022-07-13). "2.7 Å cryo-EM structure of ex vivo RML prion fibrils". Nature Communications. 13 (1): 4004. Bibcode:2022NatCo..13.4004M. doi:10.1038/s41467-022-30457-7. ISSN   2041-1723. PMC   9279362 . PMID   35831275.
  24. Hoyt F, Standke HG, Artikis E, Schwartz CL, Hansen B, Li K, Hughson AG, Manca M, Thomas OR, Raymond GJ, Race B, Baron GS, Caughey B, Kraus A (2022-07-13). "Cryo-EM structure of anchorless RML prion reveals variations in shared motifs between distinct strains". Nature Communications. 13 (1): 4005. Bibcode:2022NatCo..13.4005H. doi:10.1038/s41467-022-30458-6. ISSN   2041-1723. PMC   9279418 . PMID   35831291.
  25. Sandberg MK, Al-Doujaily H, Sharps B, Clarke AR, Collinge J (February 2011). "Prion propagation and toxicity in vivo occur in two distinct mechanistic phases". Nature. 470 (7335): 540–2. Bibcode:2011Natur.470..540S. doi:10.1038/nature09768. PMID   21350487. S2CID   4399936.
  26. Büeler H, Aguzzi A, Sailer A, Greiner RA, Autenried P, Aguet M, Weissmann C (July 1993). "Mice devoid of PrP are resistant to scrapie". Cell. 73 (7): 1339–47. doi: 10.1016/0092-8674(93)90360-3 . PMID   8100741.
  27. 1 2 Aguzzi A, Baumann F, Bremer J (2008). "The prion's elusive reason for being". Annu. Rev. Neurosci. 31: 439–77. doi:10.1146/annurev.neuro.31.060407.125620. PMID   18558863.
  28. Atkinson M (October 2001). "National scrapie plan". The Veterinary Record. 149 (15): 462. PMID   11688751.
  29. Weissmann C, Flechsig E (2003). "PrP knock-out and PrP transgenic mice in prion research". Br. Med. Bull. 66: 43–60. doi: 10.1093/bmb/66.1.43 . PMID   14522848.
  30. Katamine S, Nishida N, Sugimoto T, Noda T, Sakaguchi S, Shigematsu K, Kataoka Y, Nakatani A, Hasegawa S, Moriuchi R, Miyamoto T (December 1998). "Impaired motor coordination in mice lacking prion protein". Cell. Mol. Neurobiol. 18 (6): 731–42. doi:10.1023/A:1020234321879. PMID   9876879. S2CID   23409873.
  31. Nico PB, de-Paris F, Vinadé ER, Amaral OB, Rockenbach I, Soares BL, Guarnieri R, Wichert-Ana L, Calvo F, Walz R, Izquierdo I, Sakamoto AC, Brentani R, Martins VR, Bianchin MM (July 2005). "Altered behavioural response to acute stress in mice lacking cellular prion protein". Behav. Brain Res. 162 (2): 173–81. doi:10.1016/j.bbr.2005.02.003. PMID   15970215. S2CID   37511702.
  32. Roesler R, Walz R, Quevedo J, de-Paris F, Zanata SM, Graner E, Izquierdo I, Martins VR, Brentani RR (August 1999). "Normal inhibitory avoidance learning and anxiety, but increased locomotor activity in mice devoid of PrP(C)". Brain Res. Mol. Brain Res. 71 (2): 349–53. doi:10.1016/S0169-328X(99)00193-X. PMID   10521590.
  33. Medori R, Tritschler HJ, LeBlanc A, Villare F, Manetto V, Chen HY, Xue R, Leal S, Montagna P, Cortelli P (February 1992). "Fatal familial insomnia, a prion disease with a mutation at codon 178 of the prion protein gene". N. Engl. J. Med. 326 (7): 444–9. doi:10.1056/NEJM199202133260704. PMC   6151859 . PMID   1346338.
  34. Cagampang FR, Whatley SA, Mitchell AL, Powell JF, Campbell IC, Coen CW (1999). "Circadian regulation of prion protein messenger RNA in the rat forebrain: a widespread and synchronous rhythm". Neuroscience. 91 (4): 1201–4. doi:10.1016/S0306-4522(99)00092-5. PMID   10391428. S2CID   42892475.
  35. Criado JR, Sánchez-Alavez M, Conti B, Giacchino JL, Wills DN, Henriksen SJ, et al. (2005). "Mice devoid of prion protein have cognitive deficits that are rescued by reconstitution of PrP in neurons". Neurobiology of Disease. 19 (1–2): 255–265. doi:10.1016/j.nbd.2005.01.001. PMID   15837581. S2CID   2618712.
  36. Balducci C, Beeg M, Stravalaci M, Bastone A, Sclip A, Biasini E, et al. (February 2010). "Synthetic amyloid-beta oligomers impair long-term memory independently of cellular prion protein". Proceedings of the National Academy of Sciences of the United States of America. 107 (5): 2295–2300. Bibcode:2010PNAS..107.2295B. doi: 10.1073/pnas.0911829107 . PMC   2836680 . PMID   20133875.
  37. Coitinho AS, Freitas AR, Lopes MH, Hajj GN, Roesler R, Walz R, et al. (December 2006). "The interaction between prion protein and laminin modulates memory consolidation". The European Journal of Neuroscience. 24 (11): 3255–3264. doi:10.1111/j.1460-9568.2006.05156.x. PMID   17156386. S2CID   17164351.
  38. Shorter J, Lindquist S (June 2005). "Prions as adaptive conduits of memory and inheritance". Nature Reviews. Genetics. 6 (6): 435–450. doi:10.1038/nrg1616. PMID   15931169. S2CID   5575951.
  39. Papassotiropoulos A, Wollmer MA, Aguzzi A, Hock C, Nitsch RM, de Quervain DJ (August 2005). "The prion gene is associated with human long-term memory" (PDF). Hum. Mol. Genet. 14 (15): 2241–6. doi: 10.1093/hmg/ddi228 . PMID   15987701.
  40. Del Bo R, Comi GP, Giorda R, Crimi M, Locatelli F, Martinelli-Boneschi F, Pozzoli U, Castelli E, Bresolin N, Scarlato G (June 2003). "The 129 codon polymorphism of the prion protein gene influences earlier cognitive performance in Down syndrome subjects". J. Neurol. 250 (6): 688–92. doi:10.1007/s00415-003-1057-5. PMID   12796830. S2CID   21049364.
  41. Berr C, Richard F, Dufouil C, Amant C, Alperovitch A, Amouyel P (September 1998). "Polymorphism of the prion protein is associated with cognitive impairment in the elderly: the EVA study". Neurology. 51 (3): 734–7. doi:10.1212/wnl.51.3.734. PMID   9748018. S2CID   11352163.
  42. Croes EA, Dermaut B, Houwing-Duistermaat JJ, Van den Broeck M, Cruts M, Breteler MM, Hofman A, van Broeckhoven C, van Duijn CM (August 2003). "Early cognitive decline is associated with prion protein codon 129 polymorphism". Ann. Neurol. 54 (2): 275–6. doi:10.1002/ana.10658. PMID   12891686. S2CID   31538672.
  43. Kachiwala SJ, Harris SE, Wright AF, Hayward C, Starr JM, Whalley LJ, Deary IJ (September 2005). "Genetic influences on oxidative stress and their association with normal cognitive ageing". Neurosci. Lett. 386 (2): 116–20. doi:10.1016/j.neulet.2005.05.067. PMID   16023289. S2CID   23642220.
  44. Herms J, Tings T, Gall S, Madlung A, Giese A, Siebert H, Schürmann P, Windl O, Brose N, Kretzschmar H (October 1999). "Evidence of presynaptic location and function of the prion protein". J. Neurosci. 19 (20): 8866–75. doi: 10.1523/JNEUROSCI.19-20-08866.1999 . PMC   6782778 . PMID   10516306.
  45. Kardos J, Kovács I, Hajós F, Kálmán M, Simonyi M (August 1989). "Nerve endings from rat brain tissue release copper upon depolarization. A possible role in regulating neuronal excitability". Neurosci. Lett. 103 (2): 139–44. doi:10.1016/0304-3940(89)90565-X. PMID   2549468. S2CID   24917999.
  46. Bailey CH, Kandel ER, Si K (September 2004). "The persistence of long-term memory: a molecular approach to self-sustaining changes in learning-induced synaptic growth". Neuron. 44 (1): 49–57. doi: 10.1016/j.neuron.2004.09.017 . PMID   15450159. S2CID   2637074.
  47. Barco A, Bailey CH, Kandel ER (June 2006). "Common molecular mechanisms in explicit and implicit memory". J. Neurochem. 97 (6): 1520–33. doi: 10.1111/j.1471-4159.2006.03870.x . PMID   16805766. S2CID   26307975.
  48. Laurén J, Gimbel DA, Nygaard HB, Gilbert JW, Strittmatter SM (February 2009). "Cellular prion protein mediates impairment of synaptic plasticity by amyloid-beta oligomers". Nature. 457 (7233): 1128–32. Bibcode:2009Natur.457.1128L. doi:10.1038/nature07761. PMC   2748841 . PMID   19242475.
  49. Isaacs JD, Jackson GS, Altmann DM (October 2006). "The role of the cellular prion protein in the immune system". Clin. Exp. Immunol. 146 (1): 1–8. doi:10.1111/j.1365-2249.2006.03194.x. PMC   1809729 . PMID   16968391.
  50. Castilla J, Hetz C, Soto C (June 2004). "Molecular mechanisms of neurotoxicity of pathological prion protein". Current Molecular Medicine. 4 (4): 397–403. doi:10.2174/1566524043360654. PMID   15354870.
  51. Kovács GG, Trabattoni G, Hainfellner JA, Ironside JW, Knight RS, Budka H (November 2002). "Mutations of the prion protein gene phenotypic spectrum". Journal of Neurology. 249 (11): 1567–1582. doi:10.1007/s00415-002-0896-9. PMID   12420099. S2CID   22688729.
  52. Collins S, McLean CA, Masters CL (September 2001). "Gerstmann-Sträussler-Scheinker syndrome,fatal familial insomnia, and kuru: a review of these less common human transmissible spongiform encephalopathies". Journal of Clinical Neuroscience. 8 (5): 387–397. doi:10.1054/jocn.2001.0919. PMID   11535002. S2CID   31976428.
  53. Montagna P, Gambetti P, Cortelli P, Lugaresi E (March 2003). "Familial and sporadic fatal insomnia". The Lancet. Neurology. 2 (3): 167–176. doi:10.1016/S1474-4422(03)00323-5. PMID   12849238. S2CID   20822956.
  54. Mead S, Whitfield J, Poulter M, Shah P, Uphill J, Campbell T, et al. (November 2009). "A novel protective prion protein variant that colocalizes with kuru exposure". The New England Journal of Medicine. 361 (21): 2056–2065. doi: 10.1056/NEJMoa0809716 . PMID   19923577.
  55. Hwang D, Lee IY, Yoo H, Gehlenborg N, Cho JH, Petritis B, et al. (2009). "A systems approach to prion disease". Molecular Systems Biology. 5 (1): 252. doi:10.1038/msb.2009.10. PMC   2671916 . PMID   19308092.
  56. 1 2 3 4 5 6 7 8 9 Laurén J (2014). "Cellular prion protein as a therapeutic target in Alzheimer's disease". Journal of Alzheimer's Disease. 38 (2): 227–244. doi:10.3233/JAD-130950. PMID   23948943.
  57. 1 2 3 4 5 Zhou J, Liu B (May 2013). "Alzheimer's disease and prion protein". Intractable & Rare Diseases Research. 2 (2): 35–44. doi:10.5582/irdr.2013.v2.2.35. PMC   4204584 . PMID   25343100.
  58. 1 2 3 Laurén J, Gimbel DA, Nygaard HB, Gilbert JW, Strittmatter SM (February 2009). "Cellular prion protein mediates impairment of synaptic plasticity by amyloid-beta oligomers". Nature. 457 (7233): 1128–1132. Bibcode:2009Natur.457.1128L. doi:10.1038/nature07761. PMC   2748841 . PMID   19242475.
  59. 1 2 He J, Li X, Yang J, Huang J, Fu X, Zhang Y, Fan H (March 2013). "The association between the methionine/valine (M/V) polymorphism (rs1799990) in the PRNP gene and the risk of Alzheimer disease: an update by meta-analysis". Journal of the Neurological Sciences. 326 (1–2): 89–95. doi:10.1016/j.jns.2013.01.020. PMID   23399523. S2CID   31070331.
  60. 1 2 Giovagnoli AR, Di Fede G, Aresi A, Reati F, Rossi G, Tagliavini F (December 2008). "Atypical frontotemporal dementia as a new clinical phenotype of Gerstmann-Straussler-Scheinker disease with the PrP-P102L mutation. Description of a previously unreported Italian family". Neurological Sciences. 29 (6): 405–410. doi:10.1007/s10072-008-1025-z. PMID   19030774. S2CID   20553167.
  61. Weiss R (1 January 2007). "Scientists Announce Mad Cow Breakthrough". The Washington Post. Retrieved 1 January 2007.
  62. Richt JA, Kasinathan P, Hamir AN, Castilla J, Sathiyaseelan T, Vargas F, et al. (January 2007). "Production of cattle lacking prion protein". Nature Biotechnology. 25 (1): 132–138. doi:10.1038/nbt1271. PMC   2813193 . PMID   17195841.
  63. Americo TA, Chiarini LB, Linden R (June 2007). "Signaling induced by hop/STI-1 depends on endocytosis". Biochem. Biophys. Res. Commun. 358 (2): 620–5. doi:10.1016/j.bbrc.2007.04.202. PMID   17498662.
  64. Zanata SM, Lopes MH, Mercadante AF, Hajj GN, Chiarini LB, Nomizo R, Freitas AR, Cabral AL, Lee KS, Juliano MA, de Oliveira E, Jachieri SG, Burlingame A, Huang L, Linden R, Brentani RR, Martins VR (Jul 2002). "Stress-inducible protein 1 is a cell surface ligand for cellular prion that triggers neuroprotection". EMBO J. 21 (13): 3307–16. doi:10.1093/emboj/cdf325. PMC   125391 . PMID   12093732.