Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Cell, Vol.

93, 337–348, May 1, 1998, Copyright 1998 by Cell Press

Prion Protein Biology Review

Stanley B. Prusiner,*† Michael R. Scott,* and seem to be composed exclusively of a modified


Stephen J. DeArmond,‡* and Fred E. Cohen†§k isoform of PrP, designated PrPSc. The normal cellular
* Department of Neurology PrP, denoted PrPC, is converted into PrPSc through a
† Department of Biochemistry and Biophysics process whereby a portion of its a-helical and coil struc-
‡ Department of Pathology ture is refolded into b sheet (Pan et al., 1993). This struc-
§ Department of Molecular and Cellular Pharmacology tural transition is accompanied by profound changes in
k Department of Medicine the physicochemical properties of the PrP. While PrPC
University of California is soluble in nondenaturing detergents, PrPSc is not. PrPC
San Francisco, California 94143 is readily digested by proteases, whereas PrPSc is par-
tially resistant (Oesch et al., 1985). Because prions ap-
Introduction pear to be composed entirely of a protein that adopts
an abnormal conformation, it is not unreasonable to
It is interesting to contemplate how the course of scien- think of prions as infectious proteins (Pan et al., 1993;
tific investigation might have proceeded had studies on Telling et al., 1996). But we hasten to add that we still
the transmissibility of inherited prion diseases not been cannot eliminate the possibility of a small ligand bound
performed until after the molecular genetic lesion had to PrPSc as an essential component of the infectious
been identified (Meggendorfer, 1930; Roos et al., 1973; prion particle.
Hsiao et al., 1989). Had the prion protein (PrP) gene been In a more broad view, prions are elements that impart
identified in families with prion disease by positional and propagate variability through multiple conformers
cloning or through the purification and sequencing of of a normal cellular protein. The species of a particular
PrP in amyloid plaques before brain extracts were prion is encoded by the sequence of the chromosomal
shown to be transmissible, the prion concept might have PrP gene of the mammal in which it last replicated. In
been more readily accepted (Prusiner, 1998). contrast to pathogens with a nucleic acid genome that
But that is not the course of events that led to our encode strain-specific properties in genes, prions seem
current understanding of prions. Creutzfeldt-Jakob dis- to encipher these properties in the tertiary structure
ease (CJD) remained a curious, rare neurodegenerative of PrPSc (Bessen and Marsh, 1994; Telling et al., 1996;
disease of unknown etiology for more than three score Prusiner, 1997).
years. Only the transmission of CJD to apes by inocula- The discovery that mutations of the PrP gene caused
tion of brain extracts from patients who had died of CJD dominantly inherited prion diseases in humans linked
initiated a path of scientific investigation that was to the genetic and infectious forms of prion diseases and
demystify that fascinating area of biomedicine (Gibbs presented another hurdle for investigators who contin-
et al., 1968). Not unexpectedly, once CJD was shown ued to argue that prion diseases are caused by viruses.
to be an infectious disease, relatively little attention was More than 20 mutations of the PrP gene are now known
paid to the familial form of the disease since most cases to cause the inherited human prion diseases, and signifi-
are sporadic and familial clusters account for a minority. cant genetic linkage has been established for five of
In this review, we focus on the prion particles that these mutations (Hsiao et al., 1989) (for review, see Prus-
cause scrapie, bovine spongiform encephalopathy (BSE), iner, 1997). The prion concept readily explains how a
and CJD. Some of the fascinating studies that led to disease can be manifest as a heritable as well as an
our knowledge of prions are not discussed here but infectious illness. Moreover, the hallmark common to
have recently been reviewed elsewhere (Prusiner, 1998). all of the prion diseases whether sporadic, dominantly
Prion diseases may present as genetic, infectious, or inherited, or acquired by infection is that they involve
sporadic disorders, all of which involve modification of the aberrant metabolism of the prion protein (Prusiner,
the prion protein (PrP), a constituent of normal mamma- 1991).
lian cells. CJD generally presents as progressive demen- Although PrPSc is the only known component of the
tia while scrapie of sheep and BSE are generally mani- infectious prion particles, these unique pathogens share
fest as ataxic illnesses (Table 1) (Wells et al., 1987). In several phenotypic traits with other infectious entities
CJD, scrapie, and BSE, as well as all of the other disor- such as viruses. Because some features of the diseases
ders frequently referred to as prion diseases (Table 1), caused by prions and viruses are similar, some scien-
spongiform degeneration and astrocytic gliosis are tists have difficulty accepting the existence of prions
found upon microscopic examination of the CNS. The despite a wealth of scientific data supporting this con-
degree of spongiform degeneration is quite variable cept (Chesebro and Caughey, 1993; Manuelidis and
while the extent of reactive gliosis correlates with the Fritch, 1996; Lasmézas et al., 1997; Chesebro, 1998).
degree of neuron loss (Masters and Richardson, 1978).
Molecular Genetics of Prion Diseases
The Prion Particle Once a PrP cDNA probe became available, molecular
Perhaps, the best current working definition of a prion genetic studies were undertaken to determine whether
is a proteinaceous infectious particle that lacks nucleic the PrP gene controls scrapie incubation times in mice.
acid (Prusiner, 1997). A wealth of data supports the Independent of the enriching of brain fractions for
contention that scrapie prions are devoid of nucleic acid scrapie infectivity that led to the discovery of PrPSc, the
Cell
338

Table 1. The Prion Diseases

Disease Host Mechanism of Pathogenesis


Kuru Fore people Infection through ritualistic cannibalism
iCJD Humans Infection from prion-contaminated HGH, dura mater
grafts, etc.
vCJD Humans Infection from bovine prions?
fCJD Humans Germline mutations in PrP gene
GSS Humans Germline mutations in PrP gene
FFI Humans Germline mutation in PrP gene (D178N, M129)
sCJD Humans Somatic mutation or spontaneous conversion of PrP C
into PrP Sc?
FSI Humans Somatic mutation or spontaneous conversion of PrP C
into PrP Sc?
Scrapie Sheep Infection in genetically susceptible sheep
BSE Cattle Infection with prion-contaminated MBM
TME Mink Infection with prions from sheep or cattle
CWD Mule deer, elk Unknown
FSE Cats Infection with prion-contaminated beef
Exotic ungulate encephalopathy Greater kudu, nyala, oryx Infection with prion-contaminated MBM

Abbreviations: BSE, bovine spongiform encephalopathy; CJD, Creutzfeldt-Jakob disease; sCJD, sporadic CJD; fCJD, familial CJD; iCJD,
iatrogenic CJD; vCJD, (new) variant CJD; CWD, chronic wasting disease; FFI, fatal familial insomnia; FSE, feline spongiform encephalopathy;
FSI, fatal sporadic insomnia; GSS, Gerstmann-Sträussler-Scheinker disease; HGH, human growth hormone; MBM, meat and bone meal; TME,
transmissible mink encephalopathy.

PrP gene was shown to be genetically linked to a locus process (Prusiner et al., 1990). The length of the incuba-
controlling the incubation time (Carlson et al., 1986). tion time after inoculation with SHa prions was inversely
Subsequently, mutation of the PrP gene was shown to proportional to the level of SHaPrPC in the brains of
be genetically linked to the development of familial prion Tg(SHaPrP) mice (Prusiner et al., 1990). Bioassays of
disease (Hsiao et al., 1989). At the same time, expression brain extracts from clinically ill Tg(SHaPrP) mice inocu-
of a Syrian hamster (SHa) PrP transgene in mice was lated with mouse (Mo) prions revealed that only Mo
shown to render the animals highly susceptible to SHa prions but no SHa prions were produced. Conversely,
prions, which demonstrated that expression of a foreign inoculation of Tg(SHaPrP) mice with SHa prions led to
PrP gene could abrogate the species barrier (Scott et the synthesis of only SHa prions. Thus, the rate of PrPSc
al., 1989). Later, PrP-deficient (Prnp 0/0) mice were found synthesis appears to be a function of the level of PrPC
to be resistant to prion infection and failed to replicate expression in Tg mice; however, the level to which PrPSc
prions, as expected (Büeler et al., 1993; Prusiner et al., accumulates appears to be independent of PrPC con-
1993). The results of these studies indicated PrP must centration (Prusiner et al., 1990).
play a central role in the transmission and pathogenesis PrP-Deficient Mice
of prion disease, but equally important, they established The development and lifespan of two lines of Prnp0/0
that the abnormal isoform is an essential component of mice were indistinguishable from controls (Büeler et al.,
the prion particle (Prusiner, 1991). 1992; Manson et al., 1994) while another line exhibited
PrP Gene Dosage Controls Length ataxia and Purkinje cell degeneration at z70 weeks of
age (Sakaguchi et al., 1996). In the former two lines with
of Incubation Time
Scrapie incubation times in mice were used to distin- normal development, altered sleep–wake cycles (Tobler
et al., 1996) and synaptic behavior in brain slices have
guish prion strains and to identify a gene controlling its
been reported (Collinge et al., 1994), but the synaptic
length (Dickinson et al., 1968; Scott et al., 1997). This
changes could not be confirmed by others (Lledo et al.,
gene was initially called Sinc based on genetic crosses
1996).
between C57Bl and VM mice, which exhibited short and
Prnp0/0 mice are resistant to prions (Büeler et al., 1993;
long incubation times, respectively (Dickinson et al., Prusiner et al., 1993). Prnp0/0 mice were sacrificed 5, 60,
1968). Because the distribution of VM mice was re- 120, and 315 days after inoculation with RML prions and
stricted, we searched for another mouse with long incu- brain extracts bioassayed in CD-1 Swiss mice. Except
bation times. I/Ln mice proved to be a suitable substitute for residual infectivity from the inoculum detected at 5
for VM mice; eventually, I/Ln and VM mice were found days after inoculation, no infectivity was detected in the
to be derived from a common ancestor. Subsequently, brains of Prnp0/0 mice (Prusiner et al., 1993). One group
the PrP gene was shown to control the length of the of investigators found that Prnp0/0 mice inoculated with
scrapie incubation time in mice (Carlson et al., 1994; RML prions and sacrificed 20 weeks later had 103.6 ID50
Moore et al., 1998). units/ml of homogenate by bioassay (Büeler et al., 1993).
Overexpression of wtPrP Transgenes Others have used this report to argue that prion infectiv-
Mice were constructed expressing different levels of the ity replicates in the absence of PrP (Chesebro and
wild-type (wt) SHaPrP transgene (Scott et al., 1989). Caughey, 1993; Lasmézas et al., 1997). Neither we nor
Inoculation of these Tg(SHaPrP) mice with SHa prions the authors of the initial report could confirm the finding
demonstrated abrogation of the species barrier resulting of prion replication in Prnp0/0 mice (Prusiner et al., 1993;
in abbreviated incubation times due to a nonstochastic Sailer et al., 1994).
Review
339

Prion Protein Structure


Once cDNA probes for PrP became available, the PrP
gene was found to be constitutively expressed in adult,
uninfected brain (Chesebro et al., 1985; Oesch et al.,
1985). This finding eliminated the possibility that PrPSc
stimulated production of more of itself by initiating tran-
scription of the PrP gene as proposed nearly two de-
cades earlier (Griffith, 1967). Determination of the struc-
ture of the PrP gene eliminated a second possible
mechanism that might explain the appearance of PrPSc
in brains already synthesizing PrPC. Since the entire pro-
tein coding region was contained within a single exon,
there was no possibility that the two PrP isoforms were
the products of alternatively spliced mRNAs (Basler et
al., 1986). Next, a posttranslational chemical modifica-
tion that distinguishes PrPSc from PrPC was considered
but none was found in an exhaustive study (Stahl et al.,
1993), and we considered it likely that PrPC and PrPSc
differed only in their conformations, a hypothesis also
proposed earlier (Griffith, 1967).
When the secondary structures of the PrP isoforms
were compared by optical spectroscopy, they were
found to be markedly different (Pan et al., 1993). Fourier
transform infrared (FTIR) and circular dichroism (CD)
spectroscopy studies showed that PrPC contains about
40% a helix and little b sheet while PrPSc is composed
of about 30% a helix and 45% b sheet (Pan et al., 1993).
That the two PrP isoforms have the same amino acid
sequence runs counter to the widely accepted view that
Figure 1. Species Variations and Mutations of the Prion Protein
the amino acid sequence specifies only one biologically Gene
active conformation of a protein (Anfinsen, 1973). (A) Species variations. The x-axis represents the human PrP se-
Prior to comparative studies on the structures of PrPC quence, with the five octarepeats and H1–H4 regions of putative
and PrPSc, metabolic labeling studies showed that the secondary structure shown as well as the three a helices A, B, and
acquisition of PrPSc protease resistance is a posttransla- C and the two b strands S1 and S2. Vertical bars above the axis
tional process (Borchelt et al., 1990). In a search for indicate the number of species that differ from the human sequence
at each position. Below the axis, the length of the bars indicates the
chemical differences that would distinguish PrPSc from
number of alternative amino acids at each position in the alignment.
PrPC, we identified ethanolamine in hydrolysates of PrP (B) Mutations causing inherited human prion disease and polymor-
27–30, which signaled the possibility that PrP might con- phisms in human, mouse, and sheep. Above the line of the human
tain a glycosylphosphatidyl inositol (GPI) anchor (Stahl sequence are mutations that cause prion disease. Below the lines
et al., 1987). Both PrP isoforms were found to carry GPI are polymorphisms, some but not all of which are known to influence
anchors and PrPC was found on the surface of cells the onset as well as the phenotype of disease. Data were compiled
by Paul Bamborough and Fred E. Cohen. Reprinted with permission
where it could be released by cleavage of the anchor.
from Science 278, pp. 245–251 (copyright 1997 American Associa-
Subsequent studies showed that PrPSc formation occurs tion for the Advancement of Science).
after PrPC reaches the cell surface (Caughey and Ray-
mond, 1991) and is localized to caveolae-like domains
(Gorodinsky and Harris, 1995; Taraboulos et al., 1995). these recombinant PrPs correspond to the secreted
Computational Models and Optical Spectroscopy form of PrP that was also identified in the cell-free trans-
Modeling studies and subsequent nuclear magnetic res- lation studies. This contention is supported by studies
onance (NMR) investigations of a synthetic PrP peptide with recombinant antibody fragments (Fabs) showing
containing residues 90–145 suggested that PrPC might that GPI-anchored PrPC on the surface of cells exhibits
contain an a helix within this region (Figure 1) (Huang an immunoreactivity similar to that of recombinant PrP
et al., 1994). This peptide contains the residues, 113– prepared with an a-helical conformation (Peretz et al.,
128, that are most highly conserved among all species 1997).
studied (Figure 1A) and that correspond to a transmem- Models of PrPSc suggest that formation of the disease-
brane region of PrP which was delineated in cell-free causing isoform involves refolding of a region corre-
translation studies. A transmembrane form of PrP was sponding roughly to residues 108–144 into b sheets (Hu-
found in brains of patients with Gerstmann-Sträussler- ang et al., 1996); the single disulfide bond joining the
Scheinker disease (GSS) caused by the A117V mutation COOH-terminal helices would remain intact since the
and in Tg mice overexpressing either mutant or wtPrP disulfide is required for PrPSc formation (Figure 2D) (Mur-
(Hegde et al., 1998). That no evidence for an a helix in amoto et al., 1996). Deletion of each of several regions
this region has been found in NMR studies of recombi- of putative secondary structure in PrP, except for the
nant PrP in an aqueous environment (Riek et al., 1996; NH 2-terminal 66 amino acids (residues 23–88) and a 36–
Donne et al., 1997; James et al., 1997) suggests that amino acid stretch (Mo residues 141–176), prevented
Cell
340

Figure 2. Structures of Prion Proteins


(A) NMR structure of Syrian hamster (SHa) recombinant (r) PrP(90–231). Presumably, the structure of the a-helical form of rPrP(90–231)
resembles that of PrP C. rPrP(90–231) is viewed from the interface where PrPSc is thought to bind to PrPC. Color scheme: a helices A (residues
144–157), B (172–193), and C (200–227) in pink; disulfide between Cys-179 and Cys-214 in yellow; conserved hydrophobic region composed
of residues 113–126 in red; loops in gray; residues 129–134 in green encompassing strand S1 and residues 159–165 in blue encompassing
strand S2; the arrows span residues 129–131 and 161–163, as these show a closer resemblance to b sheet (James et al., 1997).
(B) NMR structure of rPrP(90–231) is viewed from the interface where protein X is thought to bind to PrPC. Protein X appears to bind to the
side chains of residues that form a discontinuous epitope: some amino acids are in the loop composed of residues 165–171 and at the end
of helix B (Gln-168 and Gln-172 with a low density van der Waals rendering) while others are on the surface of helix C (Thr-215 and Gln-219
with a high density van der Waals rendering) (Kaneko et al., 1997b).
(C) Schematic diagram showing the flexibility of the polypeptide chain for PrP(29–231) (Donne et al., 1997). The structure of the portion of the
protein representing residues 90–231 was taken from the coordinates of PrP(90–231) (James et al., 1997). The remainder of the sequence
was hand-built for illustration purposes only. The color scale corresponds to the heteronuclear { 1H}-15N NOE data: red for the lowest (most
negative) values, where the polypeptide is most flexible, to blue for the highest (most positive) values in the most structured and rigid regions
of the protein. Reprinted with permission from Proc. Natl. Acad. Sci. USA 94, pp. 13452–13457 (copyright 1997 National Academy of Sciences).
(D) Plausible model for the tertiary structure of human PrPSc (Huang et al., 1996). Color scheme: S1 b strands are 108–113 and 116–122 in
red; S2 b strands are 128–135 and 138–144 in green; a helices H3 (residues 178–191) and H4 (residues 202–218) in gray; loop (residues
142–177) in yellow. Four residues implicated in the species barrier are shown in ball-and-stick form (Asn-108, Met-112, Met-129, Ala-133).
Reprinted with permission from Science 278, pp. 245–251 (copyright 1997 American Association for the Advancement of Science).

formation of PrPSc as measured in scrapie-infected cul- of putative secondary structure in PrP and, as such,
tured neuroblastoma cells (Muramoto et al., 1996). With appear to destabilize the structure of PrPC (Huang et al.,
a-PrP Fabs selected from phage display libraries and 1994; Riek et al., 1996).
two monoclonal antibodies (MAbs) derived from hybrid- NMR Structure of Recombinant PrP
omas, a major conformational change that occurs during The NMR structure of recombinant SHaPrP(90–231) was
conversion of PrPC into PrPSc has been localized to resi- determined after the protein was purified and refolded
dues 90–112 (Peretz et al., 1997). Studies with an a-PrP (Figure 2A). Residues 90–112 are not shown since
IgM MAb, which was reported to immunoprecipitate marked conformational heterogeneity was found in this
PrP Sc selectively (Korth et al., 1997), support this conclu- region while residues 113–126 constitute the conserved
sion. While these results indicate that PrPSc formation hydrophobic region that also displays some structural
involves a conformational change at the NH 2 terminus, plasticity (James et al., 1997). Although some features
mutations causing inherited prion diseases have been of the structure of rPrP(90–231) are similar to those re-
found throughout the protein (Figure 1B). Interestingly, ported earlier for the smaller recombinant MoPrP(121–
all of the known point mutations in PrP with biological 231) fragment (Riek et al., 1996), substantial differences
significance occur either within or adjacent to regions were found. For example, the loop at the NH2 terminus
Review
341

Figure 3. Schematic Diagram Showing Tem-


plate-Assisted PrPSc Formation
In the initial step, PrPC binds to protein X to
form the PrP */protein X complex. Next, PrP Sc
binds to PrP* that has already formed a com-
plex with protein X. When PrP* is transformed
into a nascent molecule of PrP Sc, protein X is released and a dimer of PrPSc remains. The inactivation target size of an infectious prion suggests
that it is composed of a dimer of PrP Sc (Bellinger-Kawahara et al., 1988). In the model depicted here, a fraction of infectious PrP Sc dimers
dissociate into uninfectious monomers as the replication cycle proceeds while most of the dimers accumulate in accord with the increase in
prion titer that occurs during the incubation period. The precise stoichiometry of the replication process remains uncertain.

of helix B is defined in rPrP(90–231) but is disordered exposure of PrPC to GdnHCl converts it into a PrP*-like
in MoPrP(121–231); in addition, helix C is composed of molecule. Whether this PrP* -like protein is converted
residues 200–227 in rPrP(90–231) but extends only from into PrPSc is unclear. Although the PrP*-like protein
200–217 in MoPrP(121–231). The loop and the COOH- bound to PrPSc is protease-resistant and insoluble, this
terminal portion of helix C are particularly important as protease-resistant PrP has not been reisolated in order
described below (Figure 2B). Whether the differences to assess whether or not it was converted into PrPSc. It
between the two recombinant PrP fragments are due is noteworthy that recombinant PrP can be refolded into
to (i) their different lengths, (ii) species-specific differ- either a-helical or b-sheet forms but none have been
ences in sequences, or (iii) the conditions used for solv- found to possess prion infectivity as judged by bioassay.
ing the structures remains to be determined. Inherited and Sporadic Prion Diseases
Recent NMR studies of full-length MoPrP(23–231) and For inherited and sporadic prion diseases, the major
SHaPrP(29–231) have shown that the NH2 termini are question is how the first PrPSc molecules are formed.
highly flexible and lack identifiable secondary structure Once these are formed, replication presumably follows
under the experimental conditions employed (Figure 2C) the mechanism outlined for infectious disease. Several
(Donne et al., 1997). Studies of SHaPrP(29–231) indicate lines of evidence suggest that PrPSc is more stable than
transient interactions between the COOH-terminal end PrPC and that a kinetic barrier precludes the formation
of helix B and the highly flexible, NH2-terminal ran- of PrPSc under normal conditions. In the case of the
dom-coil containing the octareapeats (residues 29–125) initiation of inherited prion diseases, the barrier to PrPSc
(Donne et al., 1997). formation must be lower for the mutant (DPrPC) than the
wild-type and thus DPrP* can spontaneously rearrange
Prion Replication to form DPrPSc. While the known mutations would appear
In an uninfected cell, PrP C with the wild-type sequence to be destabilizing to the structure of PrPC, we lack
useful information about the structure of the transition
exists in equilibrium in its monomeric a-helical, prote-
state for either the mutant or wild-type sequences. Stud-
ase-sensitive state or bound to protein X (Figure 3). We
ies of PrP in the brains of patients who were heterozy-
denote the conformation of PrPC that is bound to pro-
gous for the E200K mutation revealed DPrPSc(E200K)
tein X as PrP * (Cohen et al., 1994); this conformation is
molecules that were both detergent-insoluble and re-
likely to be different from that determined under aque-
sistant to limited proteolysis while most wtPrP was
ous conditions for monomeric recombinant PrP. The
detergent-insoluble but protease-sensitive (Gabizon
PrP* /protein X complex will bind PrPSc, thereby creating
et al., 1996). These results suggest that in familial (f)
a replication-competent assembly. Order of addition ex-
CJD(E200K), insoluble wtPrP might represent a form of
periments demonstrate that for PrPC, protein X binding PrP* (Gabizon et al., 1996). In studies with CHO cells,
precedes productive PrPSc interactions (Kaneko et al., expression of DPrP(E200K) was found to be accompa-
1997b). A conformational change takes place wherein nied with the posttranslational acquisition of resistance
PrP, in a shape competent for binding to protein X and to limited proteolysis (Lehmann and Harris, 1996), but
PrPSc , represents the initial phase in the formation of whether such cell lines expressing DPrP(E200K) pro-
infectious PrPSc. duce infectious prions is unknown. It is noteworthy that
Several lines of evidence argue that the smallest infec- levels of proteinase K used in the studies where
tious prion particle is an oligomer of PrPSc , perhaps as DPrP(E200K) was expressed in CHO cells were lower
small as a dimer (Bellinger-Kawahara et al., 1988). Upon by a factor of 10–100 compared to digestions of PrPSc
purification, PrPSc tends to aggregate into insoluble derived from brain or ScN2a cells. Whether these alter-
multimers that can be dispersed into liposomes (Gabi- ations in the properties of DPrP(E200K) in CHO cells
zon et al., 1988). Insolubility does not seem to be a provide evidence for DPrP* or such changes lie outside
prerequisite for PrPSc formation or prion infectivity, the pathway of DPrPSc(E200K) formation remains to be
as suggested by some investigators (Gajdusek, 1988; determined.
Caughey et al., 1995); a protease-resistant PrPSc that is Initiation of sporadic disease may follow from a so-
soluble in 1% Sarkosyl was generated in ScN2a cells matic mutation and thus follow a path similar to that for
by expression of a PrP deletion mutant consisting of germline mutations in inherited disease. In this situation,
106 amino acid residues (Muramoto et al., 1996). the mutant PrPSc must be capable of co-opting wtPrPC,
In attempts to form PrPSc in vitro, PrPC has been ex- a process known to be possible for some mutations
posed to 3 M guanidinium HCl and then diluted 10-fold (e.g., E200K, D178N) but less likely for others (e.g.,
prior to binding to PrPSc (Kocisko et al., 1994; Kaneko P102L) (Telling et al., 1995, 1996). Alternatively, the acti-
et al., 1997a). Based on these results, we presume that vation barrier separating wtPrP C from PrPSc could be
Cell
342

crossed on rare occasions when viewed in the context Table 2. Evidence for Protein X from Transmission Studies of
of a population. Most individuals would be spared while Human Prionsa
presentations in the elderly with an incidence of z1 per
Inoculum Host MoPrP Incubation Time
million would be seen. Gene [days 6 SEM] (n/no)
Mechanism of Prion Propagation?
From the foregoing formalism, we can ask “What is the sCJD Tg(HuPrP) Prnp1/1 721 (1/10)
sCJD Tg(HuPrP)Prnp0/0 Prnp0/0 263 6 2 (6/6)
rate-limiting step in prion formation?” First, we must
sCJD Tg(MHu2M) Prnp1/1 238 6 3 (8/8)
consider the impact of the concentration of PrPSc in the sCJD Tg(MHu2M)Prnp0/0 Prnp0/0 191 6 3 (10/10)
inoculum, which is inversely proportional to the length a
Data with inoculum RG from Telling et al., 1995.
of the incubation time. Second, we must consider the
sequence of PrPSc that forms an interface with PrPC.
When the sequences of the two isoforms are identical,
the shortest incubation times are observed. Third, we PrPSc is sufficient must be the conversion of PrPC to PrP*
must consider the strain-specific conformation of PrPSc. since a dominant negative derived from a single point
Some prion strains exhibit longer incubation times than mutation could gate only a kinetically critical step in a
others; the mechanism underlying this phenomenon is cellular process. In the template-directed model, the
not understood. From these considerations, there exists conversion of PrPC to PrP* is a first order process. By
a set of conditions under which initial PrPSc concentra- contrast, NP processes follow higher order kinetics
tions can be rate-limiting. These effects presumably re- ([monomer]m, where m is the number of monomers in
late to the stability of the PrPSc, its targeting to the correct the nucleus). The experimental implications of these rate
cells and subcellular compartments, and its ability to be relationships are apparent in transgenic studies; if first
cleared. Once infection in a cell is initiated and endoge- order kinetics operate, halving the gene dose (hemizy-
nous PrPSc production is operative, then the following gotes) should double the incubation time while doubling
discussion of PrPSc formation seems most applicable. the dose of a transgene array should halve the time to
If the assembly of PrPSc into a specific dimeric or disease. This quantitative behavior has been observed
multimeric arrangement were difficult, then a nucle- in several studies in mice with altered levels of PrP
ation–polymerization (NP) formalism would be relevant. expression (Prusiner et al., 1990, 1993; Büeler et al.,
In NP processes, nucleation is the rate-limiting step and 1994; Carlson et al., 1994). The existence of prion strains
elongation or polymerization is facile. These conditions that are conformational isoforms of PrPSc with distinct
are frequently observed in peptide models of aggrega- structures, incubation times, and neurohistopathology
tion phenomena (Caughey et al., 1995); however, studies must also be considered in an analysis of the kinetics
with Tg mice expressing foreign PrP genes suggest that of PrPSc accumulation. Since the rate-limiting step in
a different process is occurring. From investigations PrPSc formation cannot involve the unique template pro-
with mice expressing both the SHaPrP transgene and vided by a strain, differential rates of intercellular spread,
the endogenous MoPrP gene, it is clear that PrPSc pro- cellular uptake, and clearance seem most likely to ac-
vides a template for directing prion replication where count for the variation in incubation times. This is consis-
we define a template as a catalyst that leaves its im- tent with the different patterns of protease sensitivity
print on the product of the reaction (Prusiner et al., and glycosylation for distinct prion strains (Bessen and
1990). Inoculation of these mice with SHaPrPSc leads to Marsh, 1994; Collinge et al., 1996; Telling et al., 1996;
the production of nascent SHaPrPSc and not MoPrPSc. Somerville et al., 1997).
Conversely, inoculation of the Tg(SHaPrP) mice with However, we hasten to add that NP models can pro-
MoPrPSc results in MoPrPSc formation and not SHaPrPSc. vide a useful description of other biologic phenomena.
Even stronger evidence for templating has emerged from Under conditions when the monomer is relatively rare
studies of prion strains passaged in Tg(MHu2M)Prnp0/0 and/or the conformational change is facile (e.g., short
mice expressing a chimeric Hu/MoPrP gene as de- peptides), the NP model will dominate. However, when
scribed in more detail below (Telling et al., 1996; Prusi- the monomer is sufficiently abundant and/or the confor-
ner, 1997). Even though the conformational templates mational conversion is difficult to accomplish, the tem-
were initially generated with PrPSc molecules having dif- plate assistance formalism provides a more likely de-
ferent sequences in patients with inherited prion dis- scription of the process.
eases, these templates are sufficient to direct replication Evidence for Protein X
of distinct PrPSc molecules when the amino acid se- Protein X was postulated to explain the results on the
quences of the substrate PrPs are identical. If the forma- transmission of human (Hu) prions to Tg mice (Table 2)
tion of this template were rate-limiting, then an NP model (Telling et al., 1994, 1995). Mice expressing both Mo and
could apply. However, studies of PrPSc formation in HuPrP were resistant to Hu prions while those express-
ScN2a cells point to a distinct rate-limiting step. ing only HuPrP were susceptible. These results argue
Cell biologic and transgenetic investigations argue for that MoPrPC inhibited transmission of Hu prions, i.e.,
the existence of a chaperone-like molecule, referred to the formation of nascent HuPrP Sc. In contrast to the
as protein X, that is required for PrPSc formation (Telling foregoing studies, mice expressing both MoPrP and chi-
et al., 1995). As described below, mutagenesis experi- meric MHu2M PrP were susceptible to Hu prions and
ments have created dominant negative forms of DPrPC mice expressing MHu2MPrP alone were only slightly
that inhibit the formation of wtPrPSc by binding protein X more susceptible. These findings contend that MoPrPC
(Kaneko et al., 1997b). This implies that the rate-limiting has only a minimal effect on the formation of chimeric
step in vivo in prion replication under conditions where MHu2MPrPSc.
Review
343

When the data on Hu prion transmission to Tg mice F1(C57Bl 3 VM) inbred mice began with isolates from
were considered together, they suggested that MoPrPC sheep with scrapie. The prototypic strains called Me7
prevented the conversion of HuPrPC into PrPSc by bind- and 22A gave incubation times of z150 and z400 days
ing to another Mo protein but had little effect on the in C57Bl mice, respectively (Dickinson et al., 1968). The
conversion of MHu2M into PrPSc. We interpreted these PrPs of C57Bl and I/Ln (and later VM) mice differ at two
results in terms of MoPrPC binding to this Mo protein residues and control incubation times (Carlson et al.,
with a higher affinity than does HuPrP C. We postulated 1994; Moore et al., 1998).
that MoPrPC had little effect on the formation of PrPSc Until recently, support for the hypothesis that the ter-
from MHu2M (Table 2) because MoPrP and MHu2M tiary structure of PrPSc enciphers strain-specific informa-
share the same amino acid sequence at the COOH termi- tion (Prusiner, 1991) was minimal except for the DY strain
nus. This also suggested that MoPrPC only weakly inhib- isolated from mink with transmissible encephalopathy
ited transmission of SHa prions to Tg(SHaPrP) mice be- (Bessen and Marsh, 1994). PrPSc in DY prions showed
cause SHaPrP is more closely related to MoPrP than is diminished resistance to proteinase K digestion as well
HuPrP. as an anomalous site of cleavage. The DY strain pre-
Using scrapie-infected mouse neuroblastoma cells sented a puzzling anomaly since other prion strains ex-
transfected with chimeric Hu/Mo PrP genes, we ex- hibiting similar incubation times did not show this al-
tended our studies of protein X. Substitution of a Hu tered susceptibility to proteinase K digestion of PrP Sc
residue at position 214 or 218 prevented PrPSc formation (Scott et al., 1997). Also notable was the generation of
(Figure 2B) (Kaneko et al., 1997b). The side chains of new strains during passage of prions through animals
these residues protrude from the same surface of the with different PrP genes (Scott et al., 1997).
COOH-terminal a helix forming a discontinuous epitope PrP Sc Conformation Enciphers Variation in Prions
with residues 167 and 171 in an adjacent loop. Substitu- Persuasive evidence that strain-specific information is
tion of a basic residue at positions 167, 171, or 218 enciphered in the tertiary structure of PrPSc comes from
prevented PrPSc formation; these mutant PrPs appear transmission of two different inherited human prion
to act as “dominant negatives” by binding protein X diseases to mice expressing a chimeric MHu2M PrP
and rendering it unavailable for prion propagation. Our transgene (Telling et al., 1996). In fatal familial insomnia
findings seem to explain the protective effects of basic (FFI), the protease-resistant fragment of PrPSc after
polymorphic residues in PrP of humans and sheep deglycosylation has an Mr of 19 kDa; whereas in
(Hunter et al., 1993; Westaway et al., 1994; Shibuya et fCJD(E200K) and most sporadic prion diseases, it is 21
al., 1998). kDa (Monari et al., 1994). This difference in molecular
Is Protein X a Molecular Chaperone? size was shown to be due to different sites of proteolytic
Since PrP undergoes a profound structural transition cleavage at the NH 2 termini of the two human PrPSc
during prion propagation, it seems likely that other pro- molecules reflecting different tertiary structures (Monari
teins such as chaperones participate in this process. et al., 1994). These distinct conformations were not un-
Whether protein X functions as a classical molecular expected since the amino acid sequences of the PrPs
chaperone or participates in PrP binding as part of its differ.
normal function but can also facilitate pathogenic as- Extracts from the brains of FFI patients transmitted
pects of PrP biology is unknown. Interestingly, scrapie- disease into mice expressing a chimeric MHu2M PrP
infected cells in culture display marked differences in
gene about 200 days after inoculation and induced for-
the induction of heat-shock proteins (Tatzelt et al., 1995),
mation of the 19 kDa PrPSc ; whereas fCJD(E200K) and
and Hsp70 mRNA has been reported to increase in
sCJD produced the 21 kDa PrPSc in mice expressing
scrapie of mice (Kenward et al., 1994). While attempts
the same transgene (Telling et al., 1996). On second
to isolate specific proteins that bind to PrP have been
passage, Tg(MHu2M) mice inoculated with FFI prions
disappointing (Oesch et al., 1990), PrP has been shown
showed an incubation time of z130 days and a 19 kDa
to interact with Bcl-2, Hsp60, and the laminin receptor
PrP Sc while those inoculated with fCJD(E200K) prions
protein by two-hybrid analysis in yeast (Kurschner and
exhibited an incubation time of z170 days and a 21 kDa
Morgan, 1996; Rieger et al., 1997). Although these stud-
PrPSc (Prusiner, 1997). The experimental data demon-
ies are suggestive, no molecular chaperone involved in
prion formation in mammalian cells has been identified. strate that MHu2MPrPSc can exist in two different confor-
mations based on the sizes of the protease-resistant
fragments; yet, the amino acid sequence of MHu2MPrPSc
Strains of Prions is invariant.
The existence of prion strains raises the question of how The results of our studies argue that PrPSc acts as a
heritable biological information can be enciphered in template for the conversion of PrPC into nascent PrPSc .
any molecule other than nucleic acid (Dickinson et al., Imparting the size of the protease-resistant fragment
1968). Strains or varieties of prions have been defined of PrPSc through conformational templating provides a
by incubation times and the distribution of neuronal vac- mechanism for both the generation and propagation of
uolation (Dickinson et al., 1968; Fraser and Dickinson, prion strains.
1968). Subsequently, the patterns of PrPSc deposition Interestingly, the protease-resistant fragment of PrPSc
were found to correlate with vacuolation profiles, and after deglycosylation with an Mr of 19 kDa has been
these patterns were also used to characterize strains of found in a patient who died after developing a clinical
prions (DeArmond et al., 1987, 1997; Bruce et al., 1989). disease similar to FFI. Since both PrP alleles encoded
The typing of prion strains in C57Bl, VM, and the wild-type sequence and a Met at position 129, we
Cell
344

labeled this case fatal sporadic insomnia (FSI). At au-


topsy, the spongiform degeneration, reactive astroglio-
sis, and PrPSc deposition were confined to the thalamus
(Mastrianni et al., 1997). These findings argue that the
clinicopathologic phenotype is determined by the con-
formation of PrPSc in accord with the results of the trans-
mission of human prions from patients with FFI to Tg
mice (Telling et al., 1996).
Mechanism of Selective Neuronal Targeting?
In addition to incubation times, neuropathologic profiles
of spongiform change have been used to characterize
prion strains (Fraser and Dickinson, 1968). However,
recent studies with PrP transgenes argue that such pro-
files are not an intrinsic feature of strains (Carp et al.,
1997; DeArmond et al., 1997). The mechanism by which
prion strains modify the pattern of spongiform degener-
ation was perplexing since earlier investigations had
shown that PrPSc deposition precedes neuronal vacuola-
tion and reactive gliosis (DeArmond et al., 1987). When
FFI prions were inoculated into Tg(MHu2M) mice, PrPSc
was confined largely to the thalamus (Figure 4A) as is the
case for FFI in humans (Telling et al., 1996). In contrast,
fCJD(E200K) prions inoculated into Tg(MHu2M) mice
produced widespread deposition of PrPSc throughout
the cortical mantel and many of the deep structures of
the CNS (Figure 4B) as is seen in fCJD(E200K) of hu-
mans. To examine whether the diverse patterns of PrPSc
deposition are influenced by Asn-linked glycosylation
of PrPC, we constructed Tg mice expressing PrPs mu-
tated at one or both of the Asn-linked glycosylation
consensus sites (DeArmond et al., 1997). These muta-
tions resulted in aberrant neuroanatomic topologies of
PrPC within the CNS, whereas pathologic point muta-
tions adjacent to the consensus sites did not alter the
distribution of PrPC. Tg mice with mutation of the second
PrP glycosylation site exhibited prion incubation times
of .500 days and unusual patterns of PrPSc deposition.
These findings raise the possibility that glycosylation
can modify the conformation of PrP and affect either
the turnover of PrPC or the clearance of PrPSc. Regional Figure 4. Regional Distribution of PrP Sc Deposition in
differences in the rate of deposition or clearance would Tg(MHu2M)Prnp0/0 Mice Inoculated with Prions from Humans Who
Died of Inherited Prion Diseases
result in specific patterns of PrPSc accumulation.
Histoblot of PrP Sc deposition in a coronal section of a
Tg(MHu2M)Prnp0/0 mouse through the hippocampus and thalamus
Yeast and Other Prions (Telling et al., 1996).
Although prions were originally defined in the context (A) The Tg mouse was inoculated with brain extract prepared from
of an infectious mammalian pathogen (Prusiner, 1982), a patient who died of FFI.
it is now becoming widely accepted that prions are ele- (B) The Tg mouse was inoculated with extract from a patient with
ments that impart and propagate variability through mul- fCJD(E200K). Cryostat sections were mounted on nitrocellulose and
treated with proteinase K to eliminate PrP C (Taraboulos et al., 1992).
tiple conformers of a normal cellular protein. Such a
To enhance the antigenicity of PrPSc , the histoblots were exposed
mechanism must surely not be restricted to a single to 3 M guanidinium isothiocyanate before immunostaining using
class of transmissible pathogens. Indeed, it is likely that a-PrP 3F4 MAb (Kascsak et al., 1987).
the original definition will need to be extended to encom- (C) Labeled diagram of a coronal section of the hippocampus/thala-
pass other situations where a similar mechanism of in- mus region. NC, neocortex; Hp, hippocampus; Hb, habenula; Th,
formation transfer occurs. thalamus; vpl, ventral posterior lateral thalamic nucleus; Hy, hypo-
Two notable prion-like determinants, [URE3] and thalamus; Am, amygdala.
[PSI], have already been described in yeast and one in
another fungus denoted [Het-s* ] (Wickner, 1994; Cher-
noff et al., 1995; Coustou et al., 1997). Studies of candi- molecular chaperone Hsp104; however, no homolog of
date prion proteins in yeast may prove particularly help- Hsp104 has been found in mammals (Chernoff et al.,
ful in the dissection of some of the events that feature 1995). The NH2-terminal prion domains of Ure2p and
in PrPSc formation. Interestingly, different strains of yeast Sup35 that are responsible for the [URE3] and [PSI]
prions have been identified (Derkatch et al., 1996). Con- phenotypes in yeast have been identified. In contrast to
version to the prion-like [PSI] state in yeast requires the PrP, which is a GPI-anchored membrane protein, both
Review
345

Ure2p and Sup35 are cytosolic proteins (Wickner, 1997). the PrP gene (Collinge et al., 1994; Lledo et al., 1996;
When the prion domains of these yeast proteins were Sakaguchi et al., 1996; Tobler et al., 1996).
expressed in E. coli, the proteins were found to polymer- Whether gene therapy for the human prion diseases
ize into fibrils with properties similar to those of proteo- will prove feasible using the dominant negative ap-
lytically trimmed PrP and other amyloids (Paushkin et proach described above for prion-resistant animals de-
al., 1997). pends on the availability of efficient vectors for delivery
Whether prions explain some other examples of ac- of the transgene to the CNS.
quired inheritance in lower organisms is unclear (Land-
man, 1991). For example, studies on the inheritance of Concluding Remarks
positional order and cellular handedness on the surface Although the study of prions has taken several unex-
of small organisms have demonstrated the epigenetic pected directions over the past three decades, a novel
nature of these phenomena but the mechanism remains and fascinating story of prion biology is emerging. In-
unclear (Frankel, 1990). vestigations of prions have elucidated a previously un-
known mechanism of disease in humans and animals.
While learning the details of the structures of PrPs and
Therapeutic Approaches to Prion Diseases deciphering the mechanism of PrPC transformation into
It seems likely that it will be possible to design effective PrPSc will be important, the fundamental principles of
therapeutics for prion diseases as our understanding of prion biology have become reasonably clear. Though
prion propagation increases. Because people at risk for some investigators prefer to view the composition of
inherited prion diseases can now be identified decades the infectious prion particle as unresolved (Manuelidis
before neurologic dysfunction is evident, the develop- and Fritch, 1996; Lasmézas et al., 1997; Chesebro,
ment of an effective therapy for these fully penetrant 1998), such a perspective denies an enlarging body of
disorders is imperative. Although we have no way of data, none of which refutes the prion concept. Moreover,
predicting the number of individuals who may develop the discovery of prion-like phenomena mediated by pro-
neurologic dysfunction from bovine prions in the future, teins unrelated to PrP in yeast and other fungi serves
seeking an effective therapy now seems most prudent not only to strengthen the prion concept but also to
(Prusiner, 1997). Interfering with the conversion of PrPC widen it.
into PrPSc seems to be the most attractive therapeutic The discovery that prion diseases in humans are
target. Reasonable strategies are either stabilizing the uniquely both genetic and infectious greatly strength-
structure of PrPC via the formation of a PrPC-drug com- ened and extended the prion concept. To date, 20 differ-
plex or modifying the action of protein X, which may ent mutations in the human PrP gene, all resulting in
function as a molecular chaperone (Figure 2B). Whether nonconservative substitutions, have been found either
it is more efficacious to design a drug that binds to PrPC to be linked genetically to or to segregate with the inher-
at the protein X–binding site or one that mimics the ited prion diseases (Figure 1B). Yet, the transmissible
structure of PrPC with basic polymorphic residues that prion particle is composed largely, if not exclusively, of
seem to prevent scrapie and CJD remains to be deter- an abnormal isoform of the prion protein designated
mined (Kaneko et al., 1997b; Shibuya et al., 1998). Since PrPSc (Prusiner, 1991).
PrPSc formation seems limited to caveolae-like domains Aberrant PrP Metabolism
(Gorodinsky and Harris, 1995; Taraboulos et al., 1995), The hallmark of all prion diseases—whether sporadic,
drugs designed to inhibit this process need not pene- dominantly inherited, or acquired by infection—is that
trate the cytosol of cells but they do need to be able to they involve the aberrant metabolism and resulting ac-
enter the CNS. Alternatively, drugs that destabilize the cumulation of the prion protein (Table 1) (Prusiner, 1991).
structure of PrPSc might also prove useful. The conversion of PrPC into PrPSc involves a conforma-
The production of domestic animals that do not repli- tion change whereby the a-helical content diminishes
cate prions may also be important with respect to pre- and the amount of b sheet increases (Pan et al., 1993).
venting prion disease. Sheep encoding the R/R polymor- These findings provide a reasonable mechanism to ex-
phism at position 171 seem to be resistant to scrapie plain the conundrum presented by the three different
(Hunter et al., 1993; Westaway et al., 1994); presumably, manifestations of prion disease.
this was the genetic basis of James Parry’s scrapie Understanding how PrPC unfolds and refolds into
eradication program in Great Britain 30 years ago (Parry, PrPSc will be of paramount importance in transferring
1962). A more effective approach using dominant nega- advances in the prion diseases to studies of other de-
tives for producing prion-resistant domestic animals, generative illnesses. The mechanism by which PrPSc is
including sheep and cattle, is probably the expression formed must involve a templating process whereby ex-
of PrP transgenes encoding R171 as well as additional isting PrPSc directs the refolding of PrPC into a nascent
basic residues at the protein X–binding site (Figure 2B) PrPSc with the same conformation. Not only will a knowl-
(Kaneko et al., 1997b). Such an approach can be readily edge of PrPSc formation help in the rational design of
evaluated in Tg mice, and once shown to be effective, drugs that interrupt the pathogenesis of prion diseases,
it could be instituted by artificial insemination of sperm but it may also open new approaches to deciphering
from males homozygous for the transgene. Less practi- the causes of and to developing effective therapies
cal is the production of PrP-deficient cattle and sheep. for the more common neurodegenerative diseases in-
Although such animals would not be susceptible to prion cluding Alzheimer’s disease, Parkinson’s disease, and
disease (Büeler et al., 1993; Prusiner et al., 1993), they amyotrophic lateral sclerosis (ALS). Indeed, the ex-
might suffer some deleterious effects from ablation of panding list of prion diseases and their novel modes of
Cell
346

transmission and pathogenesis (Table 1), as well as the by grants from the National Institute of Aging and the National Insti-
unprecedented mechanisms of prion propagation and tute of Neurologic Diseases and Stroke of the National Institutes
of Health, International Human Frontiers of Science Program, and
information transfer, indicate that much more attention
American Health Assistance Foundation, as well as by gifts from
to these fatal disorders of protein conformation is ur- the Sherman Fairchild Foundation, Keck Foundation, G. Harold and
gently needed. Leila Y. Mathers Foundation, Bernard Osher Foundation, John D.
But prions may have even wider implications than French Foundation, and Centeon.
those noted for the common neurodegenerative dis-
eases. If we think of prion diseases as disorders of References
protein conformation and do not require the diseases
Alper, T., Cramp, W.A., Haig, D.A., and Clarke, M.C. (1967). Does
to be transmissible, then what we have learned from the the agent of scrapie replicate without nucleic acid? Nature 214,
study of prions may reach far beyond these common 764–766.
illnesses. Anfinsen, C.B. (1973). Principles that govern the folding of protein
Conformational Diversity chains. Science 181, 223–230.
The discovery that proteins may have multiple biologi- Basler, K., Oesch, B., Scott, M., Westaway, D., Wälchli, M., Groth,
cally active conformations may prove no less important D.F., McKinley, M.P., Prusiner, S.B., and Weissmann, C. (1986).
than the implications of prions for diseases. How many Scrapie and cellular PrP isoforms are encoded by the same chromo-
different tertiary structures can PrPSc adopt? This query somal gene. Cell 46, 417–428.
not only addresses the issue of the limits of prion diver- Bellinger-Kawahara, C.G., Kempner, E., Groth, D.F., Gabizon, R.,
sity but also applies to proteins as they normally function and Prusiner, S.B. (1988). Scrapie prion liposomes and rods exhibit
target sizes of 55,000 Da. Virology 164, 537–541.
within the cell or act to affect homeostasis in multicellu-
Bessen, R.A., and Marsh, R.F. (1994). Distinct PrP properties sug-
lar organisms. The expanding list of chaperones that
gest the molecular basis of strain variation in transmissible mink
assist in the folding and unfolding of proteins promises encephalopathy. J. Virol. 68, 7859–7868.
much new knowledge about this process. For example,
Borchelt, D.R., Scott, M., Taraboulos, A., Stahl, N., and Prusiner,
it is now clear that proproteases can carry their own S.B. (1990). Scrapie and cellular prion proteins differ in their kinetics
chaperone activity where the pro portion of the protein of synthesis and topology in cultured cells. J. Cell Biol. 110, 743–752.
functions as a chaperone in cis to guide the folding of Bruce, M.E., McBride, P.A., and Farquhar, C.F. (1989). Precise tar-
the proteolytically active portion before it is cleaved geting of the pathology of the sialoglycoprotein, PrP, and vacuolar
(Shinde et al., 1997). Such a mechanism might well fea- degeneration in mouse scrapie. Neurosci. Lett. 102, 1–6.
ture in the maturation of polypeptide hormones. Interest- Büeler, H., Fischer, M., Lang, Y., Bluethmann, H., Lipp, H.-P., DeAr-
ingly, mutation of the chaperone portion of prosubtilisin mond, S.J., Prusiner, S.B., Aguet, M., and Weissmann, C. (1992).
resulted in the folding of a subtilisin protease with differ- Normal development and behaviour of mice lacking the neuronal
cell-surface PrP protein. Nature 356, 577–582.
ent properties than the one folded by the wild-type chap-
Büeler, H., Aguzzi, A., Sailer, A., Greiner, R.-A., Autenried, P., Aguet,
erone. Such chaperones have also been shown to work
M., and Weissmann, C. (1993). Mice devoid of PrP are resistant to
in trans (Shinde et al., 1997). Besides transient metabolic scrapie. Cell 73, 1339–1347.
regulation within the cell and hormonal regulation of Büeler, H., Raeber, A., Sailer, A., Fischer, M., Aguzzi, A., and Weiss-
multicellular organisms, it is not unreasonable to sug- mann, C. (1994). High prion and PrP Sc levels but delayed onset of
gest that assembly of proteins into multimeric structures disease in scrapie-inoculated mice heterozygous for a disrupted
such as intermediate filaments might be controlled at PrP gene. Mol. Med. 1, 19–30.
least in part by alternative conformations of proteins. Carlson, G.A., Kingsbury, D.T., Goodman, P.A., Coleman, S., Mar-
Such regulation of multimeric protein assemblies might shall, S.T., DeArmond, S.J., Westaway, D., and Prusiner, S.B. (1986).
occur in either the proteins that form the multimers or Linkage of prion protein and scrapie incubation time genes. Cell 46,
503–511.
the proteins that function to facilitate the assembly pro-
Carlson, G.A., Ebeling, C., Yang, S.-L., Telling, G., Torchia, M., Groth,
cess. Additionally, apoptosis during development and
D., Westaway, D., DeArmond, S.J., and Prusiner, S.B. (1994). Prion
throughout adult life might also be regulated at least in isolate specified allotypic interactions between the cellular and
part by alternative tertiary structures of proteins. scrapie prion proteins in congenic and transgenic mice. Proc. Natl.
Future Studies Acad. Sci. USA 91, 5690–5694.
The wealth of data establishing the essential role of PrP Carp, R.I., Meeker, H., and Sersen, E. (1997). Scrapie strains retain
in the transmission of prions and the pathogenesis of their distinctive characteristics following passages of homogenates
prion diseases has provoked consideration of how many from different brain regions and spleen. J. Gen. Virol. 78, 283–290.
biological processes are controlled by changes in pro- Caughey, B., and Raymond, G.J. (1991). The scrapie-associated
tein conformation. The extreme radiation-resistance of form of PrP is made from a cell surface precursor that is both
protease- and phospholipase-sensitive. J. Biol. Chem. 266, 18217–
the scrapie infectivity suggested that the pathogen 18223.
causing this disease and related illnesses would be dif-
Caughey, B., Kocisko, D.A., Raymond, G.J., and Lansbury, P.T., Jr.
ferent from viruses, viroids, and bacteria (Alper et al., (1995). Aggregates of scrapie-associated prion protein induce the
1967). Indeed, an unprecedented mechanism of disease cell-free conversion of protease-sensitive prion protein to the prote-
has been revealed where an aberrant conformational ase-resistant state. Chem. Biol. 2, 807–817.
change in a protein is propagated. The future of this Chernoff, Y.O., Lindquist, S.L., Ono, B., Inge-Vechtomov, S.G., and
emerging area of biology should prove even more inter- Liebman, S.W. (1995). Role of the chaperone protein Hsp104 in
esting and productive as many new discoveries emerge. propagation of the yeast prion-like factor [psi1]. Science 268,
880–884.
Acknowledgments Chesebro, B. (1998). Prion diseases: BSE and prions: uncertainties
about the agent. Science 279, 42–43.
We thank G. Carlson, N. Nathanson, and J. Safar for carefully re- Chesebro, B., and Caughey, B. (1993). Scrapie agent replication
viewing sections of this manuscript. This research was supported without the prion protein? Curr. Biol. 3, 696–698.
Review
347

Chesebro, B., Race, R., Wehrly, K., Nishio, J., Bloom, M., Lechner, S.B., and Cohen, F.E. (1994). Proposed three-dimensional structure
D., Bergstrom, S., Robbins, K., Mayer, L., Keith, J.M., et al. (1985). for the cellular prion protein. Proc. Natl. Acad. Sci. USA 91, 7139–
Identification of scrapie prion protein-specific mRNA in scrapie- 7143.
infected and uninfected brain. Nature 315, 331–333. Huang, Z., Prusiner, S.B., and Cohen, F.E. (1996). Scrapie prions: a
Cohen, F.E., Pan, K.-M., Huang, Z., Baldwin, M., Fletterick, R.J., and three-dimensional model of an infectious fragment. Fold. Des. 1,
Prusiner, S.B. (1994). Structural clues to prion replication. Science 13–19.
264, 530–531. Hunter, N., Goldmann, W., Benson, G., Foster, J.D., and Hope, J.
Collinge, J., Whittington, M.A., Sidle, K.C., Smith, C.J., Palmer, M.S., (1993). Swaledale sheep affected by natural scrapie differ signifi-
Clarke, A.R., and Jefferys, J.G.R. (1994). Prion protein is necessary cantly in PrP genotype frequencies from healthy sheep and those
for normal synaptic function. Nature 370, 295–297. selected for reduced incidence of scrapie. J. Gen. Virol. 74, 1025–
Collinge, J., Sidle, K.C.L., Meads, J., Ironside, J., and Hill, A.F. (1996). 1031.
Molecular analysis of prion strain variation and the aetiology of “new James, T.L., Liu, H., Ulyanov, N.B., Farr-Jones, S., Zhang, H., Donne,
variant” CJD. Nature 383, 685–690. D. G., Kaneko, K., Groth, D., Mehlhorn, I., Prusiner, S.B., and Cohen,
Coustou, V., Deleu, C., Saupe, S., and Begueret, J. (1997). The pro- F.E. (1997). Solution structure of a 142-residue recombinant prion
tein product of the het-s heterokaryon incompatibility gene of the protein corresponding to the infectious fragment of the scrapie iso-
fungus Podospora anserina behaves as a prion analog. Proc. Natl. form. Proc. Natl. Acad. Sci. USA 94, 10086–10091.
Acad. Sci. USA 94, 9773–9778. Kaneko, K., Wille, H., Mehlhorn, I., Zhang, H., Ball, H., Cohen, F.E.,
DeArmond, S.J., Mobley, W.C., DeMott, D.L., Barry, R.A., Beckstead, Baldwin, M.A., and Prusiner, S.B. (1997a). Molecular properties of
J.H., and Prusiner, S.B. (1987). Changes in the localization of brain complexes formed between the prion protein and synthetic pep-
prion proteins during scrapie infection. Neurology 37, 1271–1280. tides. J. Mol. Biol. 270, 574–586.
DeArmond, S.J., Sánchez, H., Yehiely, F., Qiu, Y., Ninchak-Casey, Kaneko, K., Zulianello, L., Scott, M., Cooper, C.M., Wallace, A.C.,
A., Daggett, V., Camerino, A.P., Cayetano, J., Rogers, M., Groth, D., James, T.L., Cohen, F.E., and Prusiner, S.B. (1997b). Evidence for
Torchia, M., Tremblay, P., Scott, M.R., Cohen, F.E., and Prusiner, protein X binding to a discontinuous epitope on the cellular prion
S.B. (1997). Selective neuronal targeting in prion disease. Neuron protein during scrapie prion propagation. Proc. Natl. Acad. Sci. USA
19, 1337–1348. 94, 10069–10074.
Derkatch, I.L., Chernoff, Y.O., Kushnirov, V.V., Inge-Vechtomov, Kascsak, R.J., Rubenstein, R., Merz, P.A., Tonna-DeMasi, M., Fer-
S.G., and Liebman, S.W. (1996). Genesis and variability of [PSI] prion sko, R., Carp, R.I., Wisniewski, H.M., and Diringer, H. (1987). Mouse
factors in Saccharomyces cerevisiae. Genetics 144, 1375–1386. polyclonal and monoclonal antibody to scrapie-associated fibril pro-
Dickinson, A.G., Meikle, V.M. H., and Fraser, H. (1968). Identification teins. J. Virol. 61, 3688–3693.
of a gene which controls the incubation period of some strains of Kenward, N., Hope, J., Landon, M., and Mayer, R.J. (1994). Expres-
scrapie agent in mice. J. Comp. Pathol. 78, 293–299. sion of polyubiquitin and heat-shock protein 70 genes increases in
Donne, D.G., Viles, J.H., Groth, D., Mehlhorn, I., James, T.L., Cohen, the later stages of disease progression in scrapie-infected mouse
F.E., Prusiner, S.B., Wright, P.E., and Dyson, H.J. (1997). Structure brain. J. Neurochem. 62, 1870–1877.
of the recombinant full-length hamster prion protein PrP(29–231): Kocisko, D.A., Come, J.H., Priola, S.A., Chesebro, B., Raymond,
the N terminus is highly flexible. Proc. Natl. Acad. Sci. USA 94, G.J., Lansbury, P.T., Jr., and Caughey, B. (1994). Cell-free formation
13452–13457. of protease-resistant prion protein. Nature 370, 471–474.
Frankel, J. (1990). Positional order and cellular handedness. J. Cell Korth, C., Stierli, B., Streit, P., Moser, M., Schaller, O., Fischer, R.,
Sci. 97, 205–211. Schulz-Schaeffer, W., Kretzschmar, H., Raeber, A., Braun, U., et
al. (1997). Prion (PrP Sc)-specific epitope defined by a monoclonal
Fraser, H., and Dickinson, A.G. (1968). The sequential development
of the brain lesions of scrapie in three strains of mice. J. Comp. antibody. Nature 389, 74–77.
Pathol. 78, 301–311. Kurschner, C., and Morgan, J.I. (1996). Analysis of interaction sites
Gabizon, R., McKinley, M.P., Groth, D.F., and Prusiner, S.B. (1988). in homo- and heteromeric complexes containing Bcl-2 family mem-
Immunoaffinity purification and neutralization of scrapie prion in- bers and the cellular prion protein. Mol. Brain Res. 37, 249–258.
fectivity. Proc. Natl. Acad. Sci. USA 85, 6617–6621. Landman, O.E. (1991). The inheritance of acquired characteristics.
Gabizon, R., Telling, G., Meiner, Z., Halimi, M., Kahana, I., and Prusi- Annu. Rev. Genetics 25, 1–20.
ner, S.B. (1996). Insoluble wild-type and protease-resistant mutant Lasmézas, C.I., Deslys, J.-P., Robain, O., Jaegly, A., Beringue, V.,
prion protein in brains of patients with inherited prion disease. Nat. Peyrin, J.-M., Fournier, J.-G., Hauw, J.-J., Rossier, J., and Dormont,
Med. 2, 59–64. D. (1997). Transmission of the BSE agent to mice in the absence of
Gajdusek, D.C. (1988). Transmissible and non-transmissible amy- detectable abnormal prion protein. Science 275, 402–405.
loidoses: autocatalytic post-translational conversion of host precur- Lehmann, S., and Harris, D.A. (1996). Two mutant prion proteins
sor proteins to b-pleated sheet configurations. J. Neuroimmunol. expressed in cultured cells acquire biochemical properties reminis-
20, 95–110. cent of the scrapie isoform. Proc. Natl. Acad. Sci. USA 93, 5610–
Gibbs, C.J., Jr., Gajdusek, D.C., Asher, D.M., Alpers, M.P., Beck, E., 5614.
Daniel, P.M., and Matthews, W.B. (1968). Creutzfeldt-Jakob disease Lledo, P.-M., Tremblay, P., DeArmond, S.J., Prusiner, S.B., and Ni-
(spongiform encephalopathy): transmission to the chimpanzee. Sci- coll, R.A. (1996). Mice deficient for prion protein exhibit normal neu-
ence 161, 388–389. ronal excitability and synaptic transmission in the hippocampus.
Gorodinsky, A., and Harris, D.A. (1995). Glycolipid-anchored pro- Proc. Natl. Acad. Sci. USA 93, 2403–2407.
teins in neuroblastoma cells form detergent-resistant complexes Manson, J.C., Clarke, A.R., Hooper, M.L., Aitchison, L., McConnell,
without caveolin. J. Cell Biol. 129, 619–627. I., and Hope, J. (1994). 129/Ola mice carrying a null mutation in PrP
Griffith, J.S. (1967). Self-replication and scrapie. Nature 215, 1043– that abolishes mRNA production are developmentally normal. Mol.
1044. Neurobiol. 8, 121–127.
Hegde, R.S., Mastrianni, J.A., Scott, M.R., DeFea, K.A., Tremblay, Manuelidis, L., and Fritch, W. (1996). Infectivity and host responses
P., Torchia, M., DeArmond, S.J., Prusiner, S.B., and Lingappa, V.R. in Creutzfeldt-Jakob disease. Virology 216, 46–59.
(1998). A transmembrane form of the prion protein in neurodegener- Masters, C.L., and Richardson, E.P., Jr. (1978). Subacute spongiform
ative disease. Science 279, 827–834. encephalopathy Creutzfeldt-Jakob disease—the nature and pro-
Hsiao, K., Baker, H.F., Crow, T.J., Poulter, M., Owen, F., Terwilliger, gression of spongiform change. Brain 101, 333–344.
J.D., Westaway, D., Ott, J., and Prusiner, S.B. (1989). Linkage of a Mastrianni, J., Nixon, F., Layzer, R., DeArmond, S.J., and Prusiner,
prion protein missense variant to Gerstmann-Sträussler syndrome. S.B. (1997). Fatal sporadic insomnia: fatal familial insomnia pheno-
Nature 338, 342–345. type without a mutation of the prion protein gene. Neurology [Suppl.]
Huang, Z., Gabriel, J.-M., Baldwin, M.A., Fletterick, R.J., Prusiner, 48, A296.
Cell
348

Meggendorfer, F. (1930). Klinische und genealogische Beobach- Scott, M., Foster, D., Mirenda, C., Serban, D., Coufal, F., Wälchli,
tungen bei einem Fall von spastischer Pseudosklerose Jakobs. Z. M., Torchia, M., Groth, D., Carlson, G., DeArmond, S.J., et al. (1989).
Gesamte Neurol. Psychiatr. 128, 337–341. Transgenic mice expressing hamster prion protein produce species-
Monari, L., Chen, S.G., Brown, P., Parchi, P., Petersen, R.B., Mikol, specific scrapie infectivity and amyloid plaques. Cell 59, 847–857.
J., Gray, F., Cortelli, P., Montagna, P., Ghetti, B., et al. (1994). Fatal Scott, M.R., Groth, D., Tatzelt, J., Torchia, M., Tremblay, P., DeAr-
familial insomnia and familial Creutzfeldt-Jakob disease: different mond, S.J., and Prusiner, S.B. (1997). Propagation of prion strains
prion proteins determined by a DNA polymorphism. Proc. Natl. Acad. through specific conformers of the prion protein. J. Virol. 71, 9032–
Sci. USA 91, 2839–2842. 9044.
Moore, R.C., Hope, J., McBride, P.A., McConnell, I., Selfridge, J., Shibuya, S., Higuchi, J., Shin, R.-W., Tateishi, J., and Kitamoto, T.
Melton, D.W., and Manson, J.C. (1998). Mice with gene targetted (1998). Protective prion protein polymorphisms against sporadic
prion protein alterations show that Prn-p, Sinc and Prni are congru- Creutzfeldt-Jakob disease. Lancet 351, 419.
ent. Nat. Genet. 18, 118–125. Shinde, U.P., Liu, J.J., and Inouye, M. (1997). Protein memory
Muramoto, T., Scott, M., Cohen, F., and Prusiner, S.B. (1996). Re- through altered folding mediated by intramolecular chaperones. Na-
combinant scrapie-like prion protein of 106 amino acids is soluble. ture 389, 520–522.
Proc. Natl. Acad. Sci. USA 93, 15457–15462. Somerville, R.A., Chong, A., Mulqueen, O.U., Birkett, C.R., Wood,
Oesch, B., Westaway, D., Wälchli, M., McKinley, M.P., Kent, S.B.H., S.C.E.R., and Hope, J. (1997). Biochemical typing of scrapie strains.
Aebersold, R., Barry, R.A., Tempst, P., Teplow, D.B., Hood, L.E., et Nature 386, 564.
al. (1985). A cellular gene encodes scrapie PrP 27–30 protein. Cell Stahl, N., Borchelt, D.R., Hsiao, K., and Prusiner, S.B. (1987). Scrapie
40, 735–746. prion protein contains a phosphatidylinositol glycolipid. Cell 51,
Oesch, B., Teplow, D.B., Stahl, N., Serban, D., Hood, L.E., and Prusi- 229–240.
ner, S.B. (1990). Identification of cellular proteins binding to the Stahl, N., Baldwin, M.A., Teplow, D.B., Hood, L., Gibson, B.W., Burl-
scrapie prion protein. Biochemistry 29, 5848–5855. ingame, A.L., and Prusiner, S.B. (1993). Structural analysis of the
Pan, K.-M., Baldwin, M., Nguyen, J., Gasset, M., Serban, A., Groth, scrapie prion protein using mass spectrometry and amino acid se-
D., Mehlhorn, I., Huang, Z., Fletterick, R.J., Cohen, F.E., and Prusiner, quencing. Biochemistry 32, 1991–2002.
S.B. (1993). Conversion of a-helices into b-sheets features in the Taraboulos, A., Jendroska, K., Serban, D., Yang, S.-L., DeArmond,
formation of the scrapie prion proteins. Proc. Natl. Acad. Sci. USA S.J., and Prusiner, S.B. (1992). Regional mapping of prion proteins
90, 10962–10966. in brains. Proc. Natl. Acad. Sci. USA 89, 7620–7624.
Parry, H.B. (1962). Scrapie: a transmissible and hereditary disease Taraboulos, A., Scott, M., Semenov, A., Avrahami, D., Laszlo, L.,
of sheep. Heredity 17, 75–105. and Prusiner, S.B. (1995). Cholesterol depletion and modification
Paushkin, S.V., Kushnirov, V.V., Smirnov, V.N., and Ter-Avanesyan, of COOH-terminal targeting sequence of the prion protein inhibits
M.D. (1997). In vitro propagation of the prion-like state of yeast formation of the scrapie isoform. J. Cell Biol. 129, 121–132.
Sup35 protein. Science 277, 381–383. Tatzelt, J., Zuo, J., Voellmy, R., Scott, M., Hartl, U., Prusiner, S.B.,
Peretz, D., Williamson, R.A., Matsunaga, Y., Serban, H., Pinilla, C., and Welch, W.J. (1995). Scrapie prions selectively modify the stress
Bastidas, R., Rozenshteyn, R., James, T.L., Houghten, R.A., Cohen, response in neuroblastoma cells. Proc. Natl. Acad. Sci. USA 92,
F.E., et al. (1997). A conformational transition at the N terminus of 2944–2948.
the prion protein features in formation of the scrapie isoform. J. Telling, G.C., Scott, M., Hsiao, K.K., Foster, D., Yang, S.-L., Torchia,
Mol. Biol. 273, 614–622. M., Sidle, K.C.L., Collinge, J., DeArmond, S.J., and Prusiner, S.B.
Prusiner, S.B. (1982). Novel proteinaceous infectious particles cause (1994). Transmission of Creutzfeldt-Jakob disease from humans to
scrapie. Science 216, 136–144. transgenic mice expressing chimeric human-mouse prion protein.
Prusiner, S.B. (1991). Molecular biology of prion diseases. Science Proc. Natl. Acad. Sci. USA 91, 9936–9940.
252, 1515–1522. Telling, G.C., Scott, M., Mastrianni, J., Gabizon, R., Torchia, M.,
Prusiner, S.B. (1997). Prion diseases and the BSE crisis. Science Cohen, F.E., DeArmond, S.J., and Prusiner, S.B. (1995). Prion propa-
gation in mice expressing human and chimeric PrP transgenes impli-
278, 245–251.
cates the interaction of cellular PrP with another protein. Cell 83,
Prusiner, S.B. (1998). Prions. Les Prix Nobel, in press.
79–90.
Prusiner, S.B., Scott, M., Foster, D., Pan, K.-M., Groth, D., Mirenda, Telling, G.C., Parchi, P., DeArmond, S.J., Cortelli, P., Montagna, P.,
C., Torchia, M., Yang, S.-L., Serban, D., Carlson, G.A., et al. (1990). Gabizon, R., Mastrianni, J., Lugaresi, E., Gambetti, P., and Prusiner,
Transgenetic studies implicate interactions between homologous S.B. (1996). Evidence for the conformation of the pathologic isoform
PrP isoforms in scrapie prion replication. Cell 63, 673–686.
of the prion protein enciphering and propagating prion diversity.
Prusiner, S.B., Groth, D., Serban, A., Koehler, R., Foster, D., Torchia, Science 274, 2079–2082.
M., Burton, D., Yang, S.-L., and DeArmond, S.J. (1993). Ablation of Tobler, I., Gaus, S.E., Deboer, T., Achermann, P., Fischer, M., Rül-
the prion protein (PrP) gene in mice prevents scrapie and facilitates icke, T., Moser, M., Oesch, B., McBride, P.A., and Manson, J.C.
production of anti-PrP antibodies. Proc. Natl. Acad. Sci. USA 90, (1996). Altered circadian activity rhythms and sleep in mice devoid
10608–10612. of prion protein. Nature 380, 639–642.
Rieger, R., Edenhofer, F., Lasmézas, C.I., and Weiss, S. (1997). The Wells, G.A.H., Scott, A.C., Johnson, C.T., Gunning, R.F., Hancock,
human 37-kDa laminin receptor precursor interacts with the prion R.D., Jeffrey, M., Dawson, M., and Bradley, R. (1987). A novel pro-
protein in eukaryotic cells. Nat. Med. 3, 1383–1388. gressive spongiform encephalopathy in cattle. Vet. Rec. 121,
Riek, R., Hornemann, S., Wider, G., Billeter, M., Glockshuber, R., 419–420.
and Wüthrich, K. (1996). NMR structure of the mouse prion protein Westaway, D., Zuliani, V., Cooper, C.M., Da Costa, M., Neuman, S.,
domain PrP(121–231). Nature 382, 180–182. Jenny, A.L., Detwiler, L., and Prusiner, S.B. (1994). Homozygosity
Roos, R., Gajdusek, D.C., and Gibbs, C.J., Jr. (1973). The clinical for prion protein alleles encoding glutamine-171 renders sheep sus-
characteristics of transmissible Creutzfeldt-Jakob disease. Brain ceptible to natural scrapie. Genes Dev. 8, 959–969.
96, 1–20. Wickner, R.B. (1994). [URE3] as an altered URE2 protein: evidence
Sailer, A., Büeler, H., Fischer, M., Aguzzi, A., and Weissmann, C. for a prion analog in Saccharomyces cerevisiae. Science 264,
(1994). No propagation of prions in mice devoid of PrP. Cell 77, 566–569.
967–968. Wickner, R.B. (1997). A new prion controls fungal cell fusion incom-
Sakaguchi, S., Katamine, S., Nishida, N., Moriuchi, R., Shigematsu, patibility. Proc. Natl. Acad. Sci. USA 94, 10012–10014.
K., Sugimoto, T., Nakatani, A., Kataoka, Y., Houtani, T., Shirabe, S.,
et al. (1996). Loss of cerebellar Purkinje cells in aged mice homozy-
gous for a disrupted PrP gene. Nature 380, 528–531.

You might also like