filippo tesi phd completa - [email protected]/2736/1/filippo_zanetti.pdf ·...

101
1 INDEX ABSTRACT .........................................................................7 RIASSUNTO.......................................................................9 INTRODUCTION .............................................................. 11 Prions and prion diseases........................................................................ 11 The cellular prion protein (PrP C ) ........................................................... 12 The scrapie prion protein (PrP Sc ) and its conversion from PrP C 15 PrP C -PrP Sc conversion: the gain-of-function hypothesis ......... 15 PrP C -PrP Sc conversion: the loss-of-function hypothesis .......... 16 Biological functions of PrP C ..................................................................... 17 Oxidative stress .......................................................................................... 21 ROS mediated oxidative damage ......................................................... 24 PrP C and oxidative stress......................................................................... 25 The Prion Protein and muscular tissue .............................................. 27 The heart and ischemia/reperfusion injury ...................................... 28 Ischemia and reperfusion injury....................................................... 29 Functional and metabolic modifications in ischemic heart ..... 29 Reversible and non-reversible injury.............................................. 30 The reperfusion injury and the ROS damage .............................. 30 Ischemic Preconditioning..................................................................... 32 MATERIALS AND METHODS ............................................. 35 Mouse models .............................................................................................. 35 The Langendorff model and protocols for the perfusion of isolated hearts ............................................................................................. 36 Perfusion Protocols .................................................................................... 37 H 2 O 2 titration ................................................................................................ 39 Estimation of LDH release....................................................................... 39 Preparation of mitochondria from mouse hearts ........................... 41 SDS-PAGE and Western blotting .......................................................... 42 Sample preparation ............................................................................... 42 Deglycosylation with peptide N-glycosidase F ............................ 42 SDS-PAGE.................................................................................................. 42 Western blotting ..................................................................................... 43 Antibodies .................................................................................................. 43 Estimation of tropomyosin oxidation .............................................. 44 In situ superoxide detection .................................................................. 44 Enzymatic activity assays ....................................................................... 46 Superoxide dismutase activity assay ............................................. 46 Catalase activity assay ......................................................................... 46

Upload: hoangdien

Post on 15-Feb-2019

216 views

Category:

Documents


0 download

TRANSCRIPT

1

INDEX

ABSTRACT ......................................................................... 7

RIASSUNTO ....................................................................... 9

INTRODUCTION .............................................................. 11

Prions and prion diseases ........................................................................ 11 The cellular prion protein (PrPC) ........................................................... 12 The scrapie prion protein (PrPSc) and its conversion from PrPC 15 PrPC-PrPSc conversion: the gain-of-function hypothesis ......... 15 PrPC-PrPSc conversion: the loss-of-function hypothesis .......... 16

Biological functions of PrPC ..................................................................... 17 Oxidative stress .......................................................................................... 21 ROS mediated oxidative damage ......................................................... 24 PrPC and oxidative stress ......................................................................... 25 The Prion Protein and muscular tissue .............................................. 27 The heart and ischemia/reperfusion injury ...................................... 28 Ischemia and reperfusion injury ....................................................... 29 Functional and metabolic modifications in ischemic heart ..... 29 Reversible and non-reversible injury .............................................. 30 The reperfusion injury and the ROS damage .............................. 30 Ischemic Preconditioning ..................................................................... 32

MATERIALS AND METHODS ............................................. 35

Mouse models .............................................................................................. 35 The Langendorff model and protocols for the perfusion of isolated hearts ............................................................................................. 36 Perfusion Protocols .................................................................................... 37 H2O2 titration ................................................................................................ 39 Estimation of LDH release ....................................................................... 39 Preparation of mitochondria from mouse hearts ........................... 41 SDS-PAGE and Western blotting .......................................................... 42 Sample preparation ............................................................................... 42 Deglycosylation with peptide N-glycosidase F ............................ 42 SDS-PAGE .................................................................................................. 42 Western blotting ..................................................................................... 43 Antibodies .................................................................................................. 43 Estimation of tropomyosin oxidation .............................................. 44

In situ superoxide detection .................................................................. 44 Enzymatic activity assays ....................................................................... 46 Superoxide dismutase activity assay ............................................. 46 Catalase activity assay ......................................................................... 46

2

AIMS AND RATIONALE.................................................... 47

RESULTS ......................................................................... 49

Evaluation of the myocardial damage induced by I/R protocols in isolated hearts with different PrPC levels ...................................... 50 Hearts isolated from PrPC-OE mice are protected against loss of viability induced by post-ischemic reperfusion ...................... 50 The over-expression of PrPC reduces the degree of oxidative stress caused by post-ischemic reperfusion ................................ 52

PrPC performs anti-oxidant functions in the heart ......................... 55 The absence of PrPC enhances the protective effects of ischemic preconditioning...................................................................... 55 PrPC protects the heart from non-ischemic oxidative injury .. 58

Evaluation of the expression and/or activity of proteins involved the oxidative response, in hearts with different PrPC levels ................................................................................................................ 62 The enzymatic activity of CAT is diminished in PrPC–KO hearts .......................................................................................................... 62 Hearts with different PrPC levels have no difference in superoxide dismutase activities and expression ........................ 63 p66Shc expression is increased in PrPC–KO hearts ..................... 65

The fate of PrPC during and after the ischemic and oxidative challenges ...................................................................................................... 69 PrPC levels are decreased after I/R, but not after ischemia alone, in WT and PrPC-OE hearts ...................................................... 69 PrPC levels are preserved when I/R is preceded by IPC .......... 71 PrPC levels are largely reduced after perfusion with H2O2 ...... 72 Which is the fate of myocardial PrPC during post-ischemic reperfusion, or perfusion with H2O2? .............................................. 73

CONCLUSIONS AND PERSPECTIVES ................................ 77

BIBLIOGRAPHY .............................................................. 85

AKNOWLEDGEMENTS .................................................... 101

3

ABBREVIATIONS

.HO-: hydroxil radical

.O2-: superoxide anion

a.a.: aminoacids

ab: antibody

ADP: adenosine diphosphate

AMP: adenosine monophosphate

Asn: asparagine

ATP: adenosine triphosphate

BCA: bicinchoninic acid (protein assay)

bpm: beats per minute

BSA: bovine serum albumin

BSE: bovine spongiform encephalophathy

CAM: cell adhesion molecules

cAMP: cyclic adenosine monophosphate

CAT: catalase

CJD: Creutzfeldt Jacob Disease

CK: creatine-kinase

CNS: central nervous system

CWD: chronic wasting disease

Cys: cystein

DCB: disulphide cross-bridges

DMSO: dimethyl sulfoxide

DNA: deoxyribonucleic acid

Dpl: Doppel

DTT: dithiothreitol

ECL: Enhanced ChemiLuminescence

ECM: extra cellular matrix

EDTA: ethylenediaminetetraacetic acid

EGTA: ethylene glycol tetraacetic acid

ER: endoplasmic reticulum

Erk: extracellular signal-regulated kinases

FFI: fatal familial insomnia

4

GAGs: glycosamminoglycans

GPI: glycosyl-phosphatidylinositol

GSH: glutathione

GSS: Gerstmann-Sträussler-Scheinker

H2O2: hydrogen peroxide

HE: hydroethidium

His: histidine

HRP: horse radish peroxidase

I/R: ischemia/reperfusion

IPC: ischemic preconditioning

IS: isolation solution

KO: knock-out

LDH: lactate dehydrogenase

Mab: monoclonal antibody

MAPK: mitogen-activated protein kinases

MMPs: matrix metallo-proteases

MPG: N-2-mercaptopropionyl-glycine

mRNA: messenger ribonucleic acid

MW: molecular weight

NAD+: nicotinamide adenine dinucleotide (oxidized form)

NADH: nicotinamide adenine dinucleotide (reduced form)

NEM: N-ethylmaleimide

NMR: Nuclear Magnetic Resonance

NO: nitric oxide

NP-40: nonyl phenoxylpolyethoxylethanol

OE: over-expressing

ON: over-night

OONO-: peroxynitrite

ORF: open reading frame

Pab: polyclonal antibody

PB: perfusion buffer

PBS: phosphate buffer saline

PBS-T: phosphate buffered saline Tween-20

PC: phosphate-creatine

PI3K: phosphoinositide 3-kinases

PK: proteinase K

PKA: protein kinase A

5

PKC: protein kinase C

PM: plasma membrane

PNGase-F: peptide N-glycosidase F

Prnp: prion protein gene

PrP: prion protein

PrPC: cellular prion protein

PrPSC: scrapie prion protein (infective isoform of PrP)

RNA: ribonucleic acid

ROS: reactive oxygen species

RT: room temperature

SB: sample buffer

SDS: sodium dodecyl-sulphate

SDS-PAGE: sodium dodecyl sulfate polyacrylamide gel electrophoresis

SERCA: sarco/endoplasmic reticulum Ca2+-ATPase

SOD: superoxide dismutase

SOD1: Cu/Zn SOD

SOD2: Mn SOD

SP: signal peptide

STI1: stress inducible factor 1

TEMED: tetramethylethylenediamine

Tg: transgenic

Tm: tropomyosin

TNF: tumor necrosis factor

Tris: tris(hydroxymethyl)aminomethane

TSE: transmissible spongiform encephalopathies

Tyr: tyrosine

w/v: weight/volume

WT: wild type

6

7

ABSTRACT

The elusive function of the cellular prion protein (PrPC) hampers the

understanding of the molecular mechanism at the basis of prion diseases, and

the development of suitable therapeutic protocols. Use of cell model systems,

and genetically modified animals, have nevertheless suggested a number of

potential roles for the protein, ranging from protecting against oxidative

stress to cell differentiation. Because we now know that muscle is involved in

PrPC pathophysiology, we have considered intact heart paradigms for the in

situ study of the cell-protecting function of PrPC.

Isolated muscle organs retain the cell native environment and are also more

suitable to experimental designs than whole animals. Accordingly, by taking

advantage of mice expressing different PrPC amounts (wild type (WT), knock-

out (KO) and overexpressors (OE)), the protection of PrPC against cell

oxidative injuries was investigated in isolated hearts subjected to

ischemia/reperfusion and perfusion protocols that involve oxidative stress. In

line with the putative capability of PrPC to antagonize oxidative injury and cell

death mechanisms, our prediction was that hearts from PrPC-KO adult mice

manifest an overt phenotype after ischemic challenge, resulting in

exacerbation of heart oxidative damage. Conversely, PrPC-OE mice should

demonstrate a higher resistance over reactive oxygen species (ROS)

production.

We found that PrPC-OE hearts were more protected from the damage induced

by post-ischemic reperfusion than WT and PrPC-KO hearts, as indicated by

reduced cell death and decreased oxidation of myofibrillar protein and

accumulation of ROS. We then reasoned that, if indeed PrPC acts as an

antioxidant, absence of PrPC should increase the effect of ischemic

preconditioning (IPC), in contrast to the less evident protection in hearts from

PrPC-OE mice. Our data on hearts subjected to IPC nicely fitted with this

prediction, given that IPC led to a strong decrease of damage in PrPC-KO

hearts, an intermediate protection in WT hearts, and no significant effect in

PrPC-OE hearts. We also applied protocols of non-ischemic oxidative injury, by

subjecting isolated hearts to perfusion with hydrogen peroxide. Such

treatment was associated with a significantly larger myocardial cell loss and

8

myofibrillar oxidative damage PrPC-KO hearts, compared to hearts from WT

and PrPC-OE mice.

We then investigated the possible modulation by PrPC of proteins involved in

the oxidative stress response. We performed enzymatic activity assays on

catalase (CAT) and mitochondrial and cytosolic superoxide dismutase (SOD):

we found a decrease in CAT activity in PrPC-KO hearts with respect to PrPC-

expressing counterparts, whereas no major variation in the

activity/expression of SOD was registered among the different PrPC-

genotypes. In addition, we found increased levels of both total and

mitochondrial p66Shc, a protein involved in oxidative stress-mediated

apoptosis, in hearts lacking PrPC. This unprecedented and intriguing finding

demands further investigations in the future.

This data thus supports both the value of the in situ muscle paradigm for

studying the physiologic function of PrPC, and the role of PrPC against

oxidative insults and cell damage.

9

RIASSUNTO

L’esatto ruolo che la proteina prionica cellulare (PrPC) svolge nella fisiologia

della cellula è ancora incerto, e questo impedisce la comprensione dei

meccanismi che stanno alla base delle malattie da prione e lo sviluppo di

opportune strategie terapeutiche. Studi condotti su modelli cellulare e animali

geneticamente modificati, tuttavia, hanno suggerito che PrPC possa svolgere

un ruolo protettivo nei confronti dello stress ossidativo e di segnali di morte

cellulare. PrPC è particolarmente abbondante nel sistema nervoso centrale, ma

è espressa a livelli elevati anche nei tessuti muscolari. Inoltre, recenti

evidenze hanno correlato la proteina alla fisiopatologia muscolare. Per questo

motivo, abbiamo orientato la nostra ricerca sull’utilizzo di cuori isolati e

perfusi, un paradigma sperimentale innovativo per lo studio in situ delle

funzioni protettive della PrPC.

Rispetto alle cellule in coltura, i cuori isolati hanno il vantaggio di mantenere

l’ambiente cellulare di origine e sono inoltre modelli più adatti dell’animale

intero per le manipolazioni sperimentali. Di conseguenza, servendoci di topi

wild-type (WT) e geneticamente modificati, esprimenti differenti quantità di

PrPC (knock-out (KO) e sovra-esprimenti (OE) la proteina), abbiamo verificato

la putativa funzione antiossidante di PrPC servendoci di cuori isolati sottoposti

a protocolli di ischemia/riperfusione (I/R), o di perfusione, che implicano lo

stress ossidativo. La nostra previsione, in linea con la putativa capacità di PrPC

di contrastare l’insulto ossidativo ed i meccanismi di morte cellulare, era che i

cuori espiantati da topi PrPC-KO e da topi PrPC-OE, e sottoposti a protocolli di

I/R, manifestassero, rispettivamente, una maggiore e minore sensibilità al

danno rispetto alla controparte WT.

Quello che abbiamo rilevato è che i cuori PrPC-OE sono più resistenti al danno

indotto dalla riperfusione post-ischemica rispetto a cuori WT e PrPC-KO, come

indicato dalla riduzione di morte cellulare, ossidazione di proteine miofibrillari

ed accumulo di specie reattive dell’ossigeno (ROS). Abbiamo quindi ipotizzato

che, se realmente PrPC agisse come un agente anti-ossidante, l’assenza della

proteina avrebbe potuto aumentare la protezione conferita dall’utilizzo di un

protocollo di pre-condizionamento ischemico (IPC), il cui meccanismo si basa

sulla produzione di piccole quantità di ROS. Questa ipotesi si è dimostrata

10

corretta, dato che il protocollo di IPC svolge un forte ruolo protettivo nei cuori

PrPC-KO, uno intermedio nei WT, e nessun effetto nei cuori PrPC-OE. Abbiamo

inoltre applicato protocolli basati su un tipo di danno ossidativo non

ischemico, perfondendo i cuori isolati con perossido di idrogeno. Tale

trattamento produce una maggiore morte cellulare ed una maggiore

ossidazione delle proteine miofibrillari nei cuori PrPC-KO, paragonati a quelli

WT e PrPC-OE.

Abbiamo infine ipotizzato un possibile ruolo di PrPC nella modulazione

dell’attività/espressione di proteine coinvolte nella risposta agli stimoli

ossidativi. A tal fine, abbiamo testato l’attività, in cuori non perfusi, di alcuni

enzimi scavenger di ROS, tra cui catalasi (CAT) e superossido dismutasi

(SOD) mitocondriale e citosolica. Mentre abbiamo osservato una riduzione

significativa dell’attività di CAT nei cuori PrPC-KO rispetto a quelli esprimenti

PrPC, l’espressione e l’attività delle SOD non sono risultate differenti nei tre

genotipi di PrPC. Da sottolineare, infine, che è stato dimostrato un aumento

dell’espressione di p66Shc, una proteina coinvolta nella mediazione di segnali

pro-apoptotici, nei cuori privi PrPC. Tale osservazione, assolutamente inedita,

meriterà ulteriori approfondimenti futuri.

I nostri risultati supportano dunque sia il valore del nuovo modello

sperimentale in situ per lo studio della funzione fisiologica di PrPC, sia il

coinvolgimento della proteina nelle difese contro lo stress ossidativo ed il

danno cellulare.

11

INTRODUCTION

The prion protein (PrP) was discovered while trying to identify an elusive

etiological agent of a group of rare fatal neurodegenerative disease, anatomo-

pathologically defined transmissible spongiform encephalopathies (TSEs).

Such etiological agent, later termed prion, was found in patients and animals

affected by TSEs. Prions were found in β-amyloid aggregates that were mainly

composed by an aberrant conformer (PrPSc) of the cellular prion protein

(PrPC). PrPC is a highly conserved cell surface sialo-glycoprotein,

physiologically expressed – in a non-aggregated form – in all mammalian

tissues, particularly in the central nervous system (CNS). While the

implication of PrPSc in the onset and transmission of TSEs is now well

recognised, the mechanisms of prion-associated neurodegeneration and the

physiologic role of PrPC are still largely elusive.

Prions and prion diseases

TSEs can be of infectious, genetic, or sporadic nature and are characterized

by neurodegeneration and protein aggregation (Prusiner, 1998). These

diseases include Creutzfeldt-Jakob disease (CJD), Gerstmann-Sträussler-

Scheinker (GSS), fatal familial insomnia (FFI), and kuru in humans, scrapie in

sheep, chronic wasting disease (CWD) in cervids and the bovine spongiform

encephalopathy (BSE), also known as “mad cow disease”. Human TSEs can

affect subjects at distinct age groups, with a variety of motor or cognitive

symptoms, and although their prevalence is relatively low (one case per

million per year in western countries), they are still incurable and invariably

fatal (Knight and Will, 2004).

In 1967, J.S. Griffith proposed the idea that a sole protein, without the action

of nucleic acid, could “replicate”, thus spreading biological information in other

organisms (Griffith, 1967). This proposal was confirmed by several studies

demonstrating that the transmissible agent resisted doses of radiation that

easily inactivated both viruses and bacteria (Alper, 1967), and the profile of

sensitivity of the infectious agent to various chemicals differed from both

viruses and viroids, suggesting that it might not depend on nucleic acids to

12

propagate (Bellinger-Kawahara et al., 1987). On the basis of those

observations, Stanley Prusiner demonstrated that a protein unusually

resistant to proteolysis was required for infectivity of diseased brain extracts

(Prusiner et al., 1984), whereas no compelling evidence supported the need

for other components, especially nucleic acids. This and other findings

brought him to reconsider the Griffith’s hypothesis of the sole protein in the

“prion hypothesis”, were the newly coined term prion (the acronym for

“proteinaceous infectious particle”) indicates this novel pathogen (Prusiner

1998). This hypothesis contributed the assignment of the Nobel Prize in

Medicine to Prusiner in 1997. According to this hypothesis, the TSE

pathogenesis would not be determined by a common infectious agent

(bacteria, virus), but it would be caused by a conformational conversion of

PrPC into the aberrant isoform PrPSc, where the suffix “Sc” stands for scrapie,

the first “prion disease” to be historically (18th century) identified (Fig. 3). At

the basis of Prusiner’s hypothesis is the PrPSc putative capacity to catalyze the

pathological structural conversion of the physiologically expressed PrPC. PrPSc

could this way accumulate in the nervous tissue - due to its high resistance to

degradation - through an auto-catalytic process not mediated by nucleic

acids.

In the light of these observations, prions are unique elements in the world of

proteins, able to transmit a biological function, a property known only for

nucleic acids. This hypothesis was subsequently supported by the discovery of

prions in yeast and fungi, acting as heritable protein-based genetic elements

that cause biologically important phenotypic changes without any underlying

nucleic acid modification (Uptain and Lindquist, 2002).

The cellular prion protein (PrPC)

PrPC (Figs. 1 and 2) is a sialo-glycoprotein of about 210 aminoacids (a.a.) in

mammals, having a molecular weight of 35-36 kDa. It is probably present in

all vertebrates, which express the protein in all tissues, and particularly in the

CNS.

The prion protein gene (Prnp) was identified in 1986 (Basler et al., 1986). It is

well conserved among mammalian species, and in humans it is localized in

the short branch of chromosome 20, in the position 20p12.1 (Sparkes et al.,

1986). The gene is composed by three exons and no alternative splicing is

13

present, the open reading frame (ORF) being contained in the third exon only.

For this reason the origin of the two PrP isoforms (PrPC and PrPSc) from an

alternative splicing event was excluded. In humans the ORF codify for a 253

a.a. long polypeptide that is subsequently processed in the endoplasmatic

reticulum (ER). In the ER, the N-terminal signal peptide (a.a. 1-22) and the

sequence for a glycosyl-phosphatidylinositol (GPI) anchor docking in the C-

terminus (a.a. 231-253) are removed, and the N-glycosilation process on two

asparagine residues (Asn181 and Asn197) begins. In the Golgi apparatus,

glycans are processed by the removal of mannose residues and the addition

of complex oligosaccharidic chains. The mature protein then moves along the

secretory pathway, to be eventually delivered to the plasma membrane (PM).

PrPC is located extracellularly, bound to the external leaflet of the PM through

the GPI moiety. Like other GPI-anchored proteins, PrPC is enriched in

sphingolipid- and cholesterol-abundant membrane microdomains, known as

detergent-resistant membranes, or rafts (Taylor and Hooper, 2006),

putatively considered centres for signal transduction events (Kabouridis,

2006). PrPC half-life is of about six hours (Caughey et al., 1989), and during

its turnover it is internalized, to be then either recycled to the PM, or

degraded in acidic compartments (Vey et al., 1996; Peters et al., 2003).

FIGURE 1. Scheme of PrPC structure. The Signal peptide (SP, a.a. 1-22) is removed in the

mature form, as well as a.a. 231-253 (TM2) for the binding of the glycosyl-phosphatidylinositol

anchor (GPI). a.a. 51-91 contain a sequence of eight aminoacids repeated five-six folds. a.a. 104-

126 contain a putative transmembrane region. In the blue boxes α-helix (A, B, C) and β-strands

positions are indicated. B and C α-helix contain the glucidic (CHO) branches bound to the Asn

residues and the disulphide cross-bridge (Cys179-Cys210).

14

PrPC is highly conserved among mammals, with a 89% identity and a 97%

homology between human and murine protein sequences. The tri-dimensional

structure of recombinant PrPs from different species has been resolved by

nuclear magnetic resonance (Riek et al., 1996). It contains an N-terminal

flexible, random coiled sequence of about 100 a.a., and a C-terminal globular

domain of about another 100 a.a.. The globular domain is arranged in three

α-helices, interspersed with an anti-parallel β-pleated sheet formed by two

short β-strands. This structure is stabilized by a single disulfide bond between

cysteine residues 179 and 214 (human sequence). The N-terminal domain of

the protein has not a well-defined secondary structure, and contains five

repetitions of sequences of eight aminoacids (PHGGGWGQ) (octarepeats) that

can coordinate up-to six copper ions (Brown et al., 1997a). A hydrophobic

region, located between the octarepeat region and the first α-helix (a.a. 106-

126) is considered a possible trans-membrane domain, and exerts neurotoxic

functions (Forloni et al., 1993). Notably, despite the low sequence identity

between PrPC in chicken, turtle, frog, or fish, and the mammalian proteins, the

major structural features of PrPC are remarkably preserved in those non-

mammalian species, suggesting evolutionarily conserved functions of the

protein.

FIGURE 2. Tridimensional structure of the prion protein. The three α-helix and the two

short β-strands, composing the structured domain of PrPC, are shown in the figure. As

represented in the figure PrPC is anchored by a GPI extension to the plasma-membrane.

15

The scrapie prion protein (PrPSc) and its conversion from PrPC

PrPC and its aberrant isoform share the same aminoacidic sequence, the same

covalent structure and undergo the same post-translational modifications. The

two isoforms, however, have a different content of secondary structure. The

α-helix and β-strands content of PrPC is about 30% and 3%, respectively.

PrPSc maintains the α-helix portion, while the β-strands percentage is much

higher (45%), due to a remarkable conversion from random coil to β-

structure (Pan et al., 1993; Safar et al., 1993). The different conformation

confers to PrPSc different physico-chemical and biological properties, such as

detergent insolubility and propensity to aggregate, resistance to proteolytic

digestion, the ability to self-propagate in a host-organism, and neurotoxic

potentials (Caughey et al., 1991; Prusiner, 1984). In particular, the presence

of proteinase K (PK)-resistant PrP in tissue extracts is often taken as a proof

of prion infection. The conversion of PrPC into PrPSc can be initiated

spontaneously, as occurring in sporadic or genetic TSEs, or induced by the

challenge of exogenous prions in a host organism, as in the case of the

infectious forms. The mechanisms of prion-induced neurotoxicity are,

however, still debated. (For a recent review on prion properties and the

putative mechanisms of prion toxicity, see Aguzzi and Calella, 2009).

PrPC-PrPSc conversion: the gain-of-function hypothesis

Although several models have been proposed to account for the formation of

PrPSc aggregates, the basic proposal is that, following either infection with

PrPSc or conversion of PrPC into PrPSc associated with certain mutations

thought to destabilize the protein (Cohen et al., 1994), binding of PrPSc to

PrPC leads to further conversion, thus resulting in accumulation of PrPSc at the

expense of the normal PrPC. This hypothesis is consistent with the progressive

nature of all variants of the prion diseases, as well as with the resistance of

Prnp knockout mice to prion infection (Steele et al., 2007; Weissmann and

Flechsig, 2003). It is also thought to underlie the predominant sporadic forms,

in which pathogenesis might start with spontaneous conversion of a fraction

of PrPC by hitherto unknown reasons (Fornai et al., 2006). It is believed that

accumulation of PrPSc is the main pathogenic event leading to

neurodegeneration. PrPSc, as well as the PrPC 106–126 fragment (PrPC 105–

125 in the murine sequence), known as the neurotoxic peptide, induce cell

16

death both in vitro and in vivo (Ettaiche et al., 2000). These data are taken as

evidence that prion diseases are gain-of-function consequences of the

formation of PrPSc (Collins et al., 2004).

PrPC-PrPSc conversion: the loss-of-function hypothesis

Despite compelling evidence for conformational conversion in the course of

the diseases, it is still not clear what leads to the accumulation and

cytotoxicity of the pathological conformer. For example, although it is widely

assumed that accumulation of PrPSc causes neurodegeneration, systematic

examination of the brains of deceased patients revealed no spatial correlation

between neuronal apoptosis and deposits of PrPSc (Chrétien et al., 1999;

Dorandeu et al., 1998). Accumulated PrPSc within PrPC-expressing tissue

grafted into the brains of Prnp-knockout mice does not damage the

neighboring PrPC-null tissue (Brandner et al., 1996), and progressive

accumulation of PrPSc in glial cells around PrPC-null neurons does not induce

cell death in the knockout neurons, also arguing against a direct cytotoxic

effect of PrPSc (Mallucci and Collinge, 2004).

Moreover, subclinical forms of prion diseases have been observed in

experimentally or naturally infected animals that harbor high levels of

infectivity and PrPSc but are asymptomatic during a normal life-span (Race

and Chesebro, 1998; Hill et al., 2000). Conversely, wild-type mice inoculated

with PrPSc of bovine spongiform encephalopathy showed no detectable PK-

resistant PrP in the brain despite the presence of neurological symptoms and

neuronal death (Lasmézas et al., 1997). These conditions were observed not

only in animals but also in humans. FFI or GSS with substitution of valine for

alanine at residue 117 (A117V) revealed striking clinical manifestations but

little or undetectable PK-resistant PrP (Collinge et al., 1990; Medori et al.,

1992). Thus the pervasive gain-of-toxic-function hypothesis is still unproven,

and current models assume that PrPSc propagates at the expense of depletion

of PrPC (Weissmann, 1999), which warrants an examination of the hypothesis

that loss of function of PrPC (Samaia and Brentani, 1998), or of neurochemical

systems associated with PrPC, contributes to the pathogenesis of TSEs. Critical

appraisal of loss-of-function components in prion diseases is, nonetheless,

hampered by the controversies surrounding the physiological functions of

PrPC.

17

FIGURE 3. Ribbon drawing of the NMR structure model of the PrPC and of the

hypotetical structure of PrPSc. The α-helical regions are shown in green, β-strands blue, and

the unstructured regions in yellow. To be noted the conversion from the prevalent α-helical

structure, in PrPC (on the left), to the β-enriched structure, in PrPSc (on the right) (Cohen et al.,

1999).

Biological functions of PrPC

Despite the intimate involvement of PrPC in the onset of TSEs, its function in

cell physiology remains enigmatic. A plausible conceptual obstacle to this issue

is the lack of serious alterations in lifespan, development, or behavior of

genetically modified mice with the targeted (also post-natal) disruption of the

Prnp gene (Büeler et al., 1992; Manson, 1994; Mallucci et al., 2002). Recently,

however, mild vacuolar brain degeneration was observed in PrPC-KO mice with

FVB genotype. These animals show no prion-like clinical manifestation but

sensorimotor deficits are clearly evident long before the vacuolization stage

(Nazor et al., 2007). The sensorimotor phenotype also occurs in a GSS Tg

model, 2-3 months before the disease manifestation, highlighting the

possibility that the prion pathogenic mechanism involves the progressive loss

of PrPC function. To account for the absence of obvious PrPC-KO-associated

phenotypes, another current hypothesis proposes that PrPC deficiency provokes

subtle changes, whose manifestation needs, however, defined cell stress

conditions (reviewed in Steele et al., 2007). This notion has recently been

supported in vivo, in that PrP-less mice show a defective response to

hematopoietic cell depletion (Zhang et al., 2006). This result is particularly

18

relevant since it is the first example in which the combination of stress

conditions and analysis of extra-neuronal cells provides a clear insight on PrPC

function.

The search for the true physiologic role of PrPC is further confounded by the

unrealistic plethora of possible functions that have been ascribed to the

protein. Indeed, the extensive research devoted in last years to this issue, by

means of several cellular and animal models, has resulted in the proposition

that PrPC may play multiple, sometime contrasting, cellular actions. To name a

few of them, the involvement in copper metabolism and in defense

mechanisms against oxidative and apoptotic challenges, and a role in cell

adhesion, migration, proliferation, differentiation and death, possibly by

interacting with extracellular partners, or by taking part in multi-component

signaling complexes at the cell surface. A brief review of the major putative

PrPC’s functions is reported in the following paragraphs. (For comprehensive

reviews of PrP’s structure and functional properties, see Aguzzi et al., 2008;

Linden et al., 2008).

A large body of evidence supports the concept that PrPC is involved in

pathways protecting cells from different injuries. PrPC is able to protect cells

from (apoptotic) death induced by diverse stimuli, including serum deprivation

(Kim et al., 2004), Bax overexpression (Bounhar et al., 2001), TNF-α (Diarra-

Mehrpour et al., 2004), and anisomycin (Zanata et al., 2002). PrPC is also able

to counteract the neurodegeneration induced by N-terminal and N-proximal

PrPC deletion mutants, and by the ectopic expression in the brain of the PrPC

paralogue Doppel (Dpl), a PrPC-like protein lacking the whole unstructured N-

terminus. Indeed, these truncated PrPs induce a specific cerebellar

neurodegeneration, with demyelination and apoptotic neuronal death, only

when expressed in mice with a PrP-null genotype, and these dramatic

phenotypes are partially or totally abrogated by reintroduction of a functional

full-length PrP transgene (Shmerling et al., 1998; Moore et al., 1999; Moore et

al., 2001; Rossi et al., 2001; Radovanovich et al., 2005; Li et al., 2008). The

neuroprotective potentials of PrPC have been further underscored by studies on

ischemic brain injury in rodents. PrPC is up-regulated after cerebral ischemia,

and this correlates with a reduced damage severity (Weise et al., 2004; Shyu

et al., 2005). Accordingly, adenovirus-mediated PrPC overexpression reduces

infarct size and neurological impairment in rat brain (Shyu et al., 2005), while

– conversely – a more severe ischemic brain injury is observed in PrPC-KO mice

19

(McLennan et al., 2004; Spudich et al., 2005; Weise et al., 2006; Mitteregger

et al., 2007; Steele et al., 2009).

PrPC was also implicated in cell adhesion, recognition and differentiation,

possibly through the interaction with cell adhesion molecules (CAMs),

responsible for cell growth and differentiation (Hansen et al., 2008), and other

extra-cellular matrix (ECM) proteins, and the activation of downstream

signalling pathways. A case in point is the interaction, both in cis- and trans-

configurations, with the neuronal adhesion protein N-CAM (Schmitt-Ulms et al.,

2001) that led to neurite outgrowth (Santuccione et al., 2005). N-CAM belongs

to the CAM superfamily, which can not only mediate adhesion of cells, or link

ECM proteins to the cytoskeleton, but also, following homo- or heterophylic

interactions, act as a receptor to transduce signals ultimately resulting in

neurite outgrowth, neuronal survival and synaptic plasticity (Hansen et al.,

2008). Another example is the binding of PrPC to laminin, an heterotrimeric

glycoprotein of the ECM, which induced neuritogenesis together with neurite

adhesion and maintenance (Graner et al., 2000a, b), but also learning and

memory consolidation (Coitinho et al., 2006). Further, it has been described

that PrPC interacts with the mature 67 kDa-receptor (67LR) (and its 37 kDa-

precursor) for laminin, and with glycosamminoglycans (GAGs), each of which is

involved in neuronal differentiation and axon growth (Caughey et al., 1994;

Rieger et al., 1997; Gauczynski et al., 2001; Hundt et al., 2001; Pan et al.,

2002). More recently, Hajj et al. (2007) have reported that the direct

interaction of PrPC with another ECM protein, vitronectin, could accomplish the

same process, and that the absence of PrPC could be functionally compensated

by the overexpression of integrin, another laminin receptor (McKerracher et al.,

1996). Incidentally, the latter finding may provide a plausible explanation for

the absence of clear phenotypes in mammalian PrP-null paradigms. By

exposing primary cultured neurons to recombinant PrPs, others have shown

that homophylic trans-interactions of PrPCs are equally important for neuronal

outgrowth (Chen et al., 2003; Kanaani et al., 2005), including the formation of

synaptic contacts (Kanaani et al., 2005). Finally, it has been demonstrated that

the binding of PrPC with the secreted co-chaperone stress-inducible protein 1

(STI1) stimulated neuritogenesis (Lopes et al., 2005). However, this same

interaction had also a pro-survival effect, as did the interaction of PrPC with its

recombinant form (Chen et al., 2003).

20

More recently, by using zebrafish as an experimental paradigm, a lethal

developmental phenotype linked to the absence of PrPC was unravelled.

Zebrafish expresses two PrPC isoforms (PrP1 and PrP2) that, similarly to

mammalian PrPCs, are glycosylated and attached to the external side of the

plasma membrane through a glycolipid anchor. PrP1 and PrP2 are, however,

expressed in distinct time frames of the zebrafish embryogenesis. Accordingly,

the knockdown of the PrP1, or PrP2, gene very early in embryogenesis

impaired development at different stages (Málaga-Trillo et al., 2009). By

focusing on PrP1, this study showed that the protein was essential for cell

adhesion, and that this event occurred through PrP1 homophilic trans-

interactions and signaling. This comprised activation of the Src-related tyrosine

(Tyr) kinase p59fyn, and, possibly, Ca2+ metabolism, leading to the regulation

of the trafficking of E-cadherin, another member of CAMs superfamily. It was

also reported that overlapping PrP1 functions were performed by PrPCs from

other species, while the murine PrPC was capable to replace PP1 in rescuing, at

least in part, the PrP1-knockdown developmental phenotype. Apart from

providing the long-sought proof for a vital role of PrPC, the demonstration that

a mammalian isoform corrected the lethal zebrafish phenotype strongly

reinforces the notion of a functional interplay of PrPC with CAMS or ECM

proteins, and cell signaling, to promote neuritogenesis and neuronal survival.

The most sensible hypothesis for the multifaceted behaviour of PrPC is that the

protein participates in signal transduction centres at the cell surface, as already

suggested for other GPI-anchored proteins (Simons and Ikonen, 1997).

Accordingly, several putative partners of PrPC have been proposed (recently

reviewed in Aguzzi et al., 2008). If one assumes that these interactions are all

functionally significant, the most immediate interpretation of this “sticky”

behavior entails that PrPC acts as a scaffolding protein in different

ECM/membrane protein complexes. Each complex could then activate a specific

signaling pathway depending on the type and state of the cell, the expression

and glycosylation levels of PrPC, and availability of extra- and/or intra-cellular

signaling partners. In line with this proposition, several intracellular effectors of

PrPC-mediated signalling events have been proposed, including p59fyn,

mitogen-activated kinases (MAPK) Erk1/2, PI3K/Akt, and cAMP-PKA.

For example, it has been shown – by antibody-mediated cross-linking of PrPC –

that activation of the protein converged to Erk1/2 through p59fyn signalling

(Mouillet-Richard et al., 2000; Schneider et al., 2003). A PrPC-dependent

21

activation of p59fyn (Kanaani et al., 2005; Santuccione et al., 2005), and

Erk1/2 (but also of PI3K and cAMP-PKA) (Chen et al., 2003), was evident in

other neuronal cell paradigms, and, consistent with the almost ubiquitous

expression of PrPC, in non-neuronal cells such as Jurkat and T cells (Stuermer

et al., 2004). In addition, it has been proposed that the interaction of PrPC with

STI1 can either lead to neuritogenesis, through the activation of the ERK1/2

pathway, or promote neuronal survival, by impinging on the cAMP/PKA

pathway (Lopes et al., 2005). Interestingly, this is not the only example

reporting that engagement of PrPC activates simultaneously two independent

pathways. In fact, possibly after trans-activating the receptor for the epidermal

growth factor, the antibody-mediated clustering of PrPC was shown to impinge

on both the Erk1/2 pathway, and on a protein (stathmin) involved in

controlling microtubule dynamics (Monnet et al., 2004). It must also be noted

that, in line with the alleged role of PrPC in mediating signal transduction

events, perturbations of the ERK1/2 (Spudich et al., 2005) and Akt (Weise et

al., 2006) signalling pathways have been reported upon ischemic challenge in

PrPC-KO brains with respect to the WT counterparts, with consequent increased

post-ischemic caspase-3 activation, and exacerbation of neuronal damage.

(Spudich et al., 2005; Weise et al., 2006).

In conclusion, regardless of the still uncertain molecular and cellular

mechanisms, a mosaic of experimental data is accumulating that convincingly

assign to PrPC benign functions. This also reinforces the notion that a clear

PrPC-less phenotype, which is probably masked by compensative systems in

normal circumstances, could emerge under specific stress conditions, and that

a loss of function of PrPC may cause, or take part to, prion-induced

neurodegeneration.

Given that the proposed role of PrPC in protecting cells against oxidative injury

is central to our work, the basic principles of oxidative stress, and the putative

anti-oxidant functions of PrPC, will be briefly described in the next sections.

Oxidative stress

Oxidative stress is caused by an imbalance between the production of reactive

oxygen species (ROS) and the ability of a biological system to readily detoxify

the reactive intermediates, and/or easily repair the resulting damage. All

forms of life maintain a reducing environment within their cells. This reducing

22

environment is preserved by enzymes that maintain the reduced state

through a constant input of metabolic energy. Disturbances in this normal

redox state can cause toxic effects through the production of peroxides and

free radicals that damage all components of the cell, including proteins, lipids,

and DNA. Some of the less reactive of these species (such as superoxide) can

be converted by oxido-reduction reactions with transition metals or other

redox cycling compounds (including quinones) into more aggressive radical

species that can cause extensive cellular damage (Valko et al., 2005) (Fig. 4).

SUPEROXIDE ANION

HYDROGEN PEROXIDE

HYDROXYL RADICAL

2H+ 2H+

SUPEROXIDE ANION

HYDROGEN PEROXIDE

HYDROXYL RADICAL

2H+ 2H+

FIGURE 4. Intermediate compounds formation in the reduction of the oxygen into

water.

Most of these oxygen-derived species are produced at a low level by normal

aerobic metabolism and the damage they cause to cells is constantly repaired.

However, under severe levels of oxidative stress, the damage causes ATP

depletion, causing the cell to simply fall apart.

One source of reactive oxygen under normal conditions in humans is the

leakage of activated oxygen from the electron transport chain in mitochondria

during oxidative phosphorylation. In the electron transport chain, electrons

are passed through a series of proteins via redox reactions, with each

acceptor protein along the chain having a greater reduction potential than the

previous. The last destination for an electron along this chain is an oxygen

molecule. Normally, oxygen is reduced to produce water. However, in about

0.1-2% of electrons passing through the chain, oxygen is instead prematurely

and incompletely reduced to give the superoxide radical (�O2-), an event most

well documented for Complex I and Complex III. This radical can act both as

an oxidant and as reducer. As a reducer it reacts with cytochrome c and

metallic ions, while as an oxidant it reacts with cathecolamines and leuco-

flavins. Another important radical is hydrogen peroxide (H2O2). However,

toxicity is seldom mediated by a direct effect of H2O2, except at high

concentrations. Instead, the H2O2 is a precursor of highly oxidizing, tissue-

23

damaging radicals. H2O2 reacts with Fe2+ ions to form the hydroxyl radical

(�OH-), by the Fenton reaction. This intermediate can be originated also by the

reaction of �O2- with H2O2 in the Haber-Weiss reaction. Other sources of �OH

-

are high energy electromagnetic radiations over-exposure, or the Fe-

independent reaction between �O2- and nitric oxide (Fig. 5) (Beckman et al.,

1996).

FIGURE 5. Generation of �OH-.

ROS play important roles in cell signalling, a process termed redox signalling.

Thus, to maintain proper cellular homeostasis, a balance must be struck

between reactive oxygen production and consumption. The best studied

cellular antioxidants are the enzymes superoxide dismutase (SOD), catalase

(CAT), and glutathione peroxidise (Fig. 6). Less well studied (but probably

just as important) enzymatic antioxidants are the peroxiredoxins and the

recently discovered sulfiredoxin. Other enzymes that have antioxidant

properties (though this is not their primary role) include paraoxonase,

glutathione-S transferases, and aldehyde dehydrogenases.

Fenton’s Reaction

Peroxinitrate Reaction

Haber-Weiss Reaction

Fe2+ + H2O2 → Fe3+ + ·OH + OH−

.O2- + H2O2 → OH + HO- + O2

NO+ .O2-→ OONO-

ONOO- + H → ONOOH

ONOOH → �OH + NO2

Fenton’s Reaction

Peroxinitrate Reaction

Haber-Weiss Reaction

Fe2+ + H2O2 → Fe3+ + ·OH + OH−

.O2- + H2O2 → OH + HO- + O2

NO+ .O2-→ OONO-

ONOO- + H → ONOOH

ONOOH → �OH + NO2

24

FIGURE 6. Reactions of two scavengers of ROS: Superoxide dismutase and catalase.

It is now widely accepted that oxidative stress might be important in

neurodegenerative diseases including ALS, Parkinson's disease, Alzheimer's

disease, Huntington's disease, and – probably – also prion diseases. Moreover

it is thought to be linked to certain cardiovascular diseases, since oxidation of

low-density lipoproteins in the vascular endothelium is a precursor to plaque

formation. Oxidative stress also plays a role in the ischemic cascade due to

oxygen reperfusion injury following hypoxia. This cascade includes both

strokes and heart attacks.

ROS mediated oxidative damage

ROS, in particular �OH-, can oxidize all the biological macromolecules. The

major portion of long term effects is inflicted by damage on DNA (Evans et

al., 2004). Iron ions involved in the Fenton reaction show an electrostatic

affinity for DNA. As a consequence the oxygen radicals produced in such

reaction could form nearby DNA, thus provoking its oxidation. DNA oxidation

can lead to helix scission and production of several hydroxilated bases. In this

situation reparation enzymes could not prevent from degradation and these

modifications, in the years, could bring to carcinogenesis and aging. Protein

modification is another destructive aspect of oxidative stress. The principles

that regulate protein oxidation have been established by Garrison and co-

workers (Rodgers et al., 1968; Peterson et al., 1969). Oxidized groups, found

in the modified proteins, include charbonilic, catecholic and hydroperossidic

groups. Oxidized proteins can not be repaired but can be proteolized by a

Superoxide Dismutase Reaction

Catalase Reaction

2 H2O2 → 2 H2O + O2

2 .O2- + 2H → H2O2 + O2

Superoxide Dismutase Reaction

Catalase Reaction

2 H2O2 → 2 H2O + O2

2 .O2- + 2H → H2O2 + O2

25

family of enzymes called proteases, and replaced. Proteases can undergo

oxidation themselves, however, determining an accumulation of damaged

proteins.

FIGURE 7. Scheme of biological ROS reaction.

PrPC and oxidative stress

It is clearly established that oxidative stress can induce apoptosis (Hampton et

al., 1998). The mechanism leading to neuronal demise in TSEs is unknown, but

it is a common opinion that, as is the case for other neurodegenerative

disorders, it may be related – at least in part – to a deregulation of the defence

mechanisms against oxidative stress.

Besides its possible generic neuroprotective and antiapoptotic functions, many

reports ascribe to PrPC a role in the control of copper homeostasis and – most

importantly –anti-oxidant potentials (see Milhavet and Lehmann, 2002 for a

review). Cells selected for the resistance to copper toxicity or oxidative stress

show increased levels of PrPC, associated with increased activity of anti-oxidant

enzymes, such as SOD and glutathione peroxidase (Brown et al., 1997a). In

addition, it has been found that primary neurons from PrPC-KO mice are more

sensitive to the oxidative challenge than WT neurons (Brown et al., 1997b;

White et al., 1999; Brown et al., 2002). This may be related to a decreased

activity of glutathione reductase (White et al., 1999) and Cu/Zn SOD (SOD1)

(Brown et al., 1997b). Accordingly, decreased total SOD activity, together with

BiologicalBiological DamageDamageO2

-

H2O2

RO2.

HO.

O2 ROO.

Enzyme Enzyme inactivationinactivation

PeroxidationPeroxidation

Fenton ReactionFenton ReactionHabeHabe--Weiss ReactionWeiss Reaction

DNA DNA degradationdegradationEnzyme inactivation Enzyme inactivation

RadicalsRadicals

PeroxidationPeroxidation

Fenton ReactionFenton Reaction

BiologicalBiological DamageDamageO2

-

H2O2

RO2.

HO.

O2 ROO.

Enzyme Enzyme inactivationinactivation

PeroxidationPeroxidation

Fenton ReactionFenton ReactionHabeHabe--Weiss ReactionWeiss Reaction

DNA DNA degradationdegradationEnzyme inactivation Enzyme inactivation

RadicalsRadicals

PeroxidationPeroxidation

Fenton ReactionFenton Reaction

26

increased protein oxidation and lipid peroxidation, was reported in different

brain regions from PrPC-KO mice compared to WT brains (Klamt et al., 2001).

Importantly, reduced SOD1 activity and resistance to oxidative stress was also

reported in cultured PrPC-KO skeletal myocytes (Brown et al., 1998), indicating

that the anti-oxidant potentials of PrPC may also be significant for non-neuronal

cell types. In line with this, reduced CAT activity and increased oxidative

damage to proteins and lipids were observed in PrPC-KO cardiac and skeletal

muscles with respect to the WT counterparts. It is also important to mention

the brain metals perturbations and altered anti-oxidant activities, associated

with oxidative damage, reported in prion-infected brains, supporting the

hypothesis that a loss of the anti-oxidant functions of PrPC in the course of

prion pathogenesis is relevant for the neurodegenerative process (Wong et al.,

2001). Although the proposition that PrPC possesses an intrinsic SOD1 activity

(Brown et al., 1999) is strongly disputed (Hutter et al., 2003; Jones et al.,

2005), the possibility that the protein serves to the delivery of copper ions to

intracellular cupro-enzymes, such as SOD1, is intriguing (Brown and Besinger,

1998). Indeed, PrPC is able to bind with high affinity Cu2+ through the His-rich

N-proximal octarepeats domain (see above) (Kramer et al., 2001), and – to a

lesser extent – also through a C-proximal site (Cereghetti et al., 2001), and

Cu2+ binding promotes PrPC internalisation (Pauly and Harris, 1998). An

alternative explanation entails that PrPC might serve as a detoxifying agent

that buffers Cu2+ at the synaptic cleft, where improperly high Cu2+

concentrations are extremely harmful, and PrPC is particularly abundant. In line

with these propositions is the finding that brain microsomes from WT mice

have a 15-fold higher Cu2+ concentration per gram of wet weight than the

PrPC-KO counterpart, whereas serum from PrPC-KO mice contained almost

twice as much copper ions as WT mice (Prince and Gunson, 1998).

Besides enhancing the activity of ROS-scavenging systems, PrPC might as well

control the cellular production/accumulation of ROS. Mitochondria are believed

to be among the major intracellular sources of ROS, and an increase of

superoxide formation has been recently reported in brain mitochondria from

PrPC-KO mice with respect to WT mice (Paterson et al., 2008).

Given that in this study we have selected the cardiac muscle as an

experimental paradigm in which studying the biologic functions of PrPC, we

will now briefly describe some of the evidences that have related PrPC patho-

physiology to muscular tissues, the functional properties of the myocardium,

27

and the injury induced by ischemia/reperfusion (I/R) events, all of which are

instrumental to the understanding of our experimental approach.

The Prion Protein and muscular tissue

Also extra-neural tissues can be affected by prions, e.g. cardiac and skeletal

muscles that naturally express substantial levels of PrPC (Miele et al., 2003;

Massimino et al., 2006). PrPSc accumulates in the skeletal muscle of individuals

(humans and animals) naturally, or experimentally, affected by TSEs (Bosque

et al., 2002; Glatzel et al., 2003; Andreoletti et al., 2004; Thomzig et al.,

2004; Angers et al., 2006; Peden et al., 2006), but also of transgenic (Tg)

mouse models of some inherited TSEs, showing specific muscular pathological

changes. An example is Tg mice expressing the murine homologue of a nine-

octapeptide insertional mutation (PG14), where necrotic fibers and

accumulation of a PrPSc-like form in the skeletal muscle were observed (Chiesa

et al., 2001). Recently, a primary myopathy has been found in a Tg mouse

with muscle-specific 40 fold-overexpression of PrPc, together with abnormal

processing of the protein (Huang et al., 2007). Notably, dilated

cardiomyopathy (Ashwath et al., 2005) and skeletal muscle myositis,

accompanied by PrPSc-rich inclusion bodies (Kovacs et al., 2004), have been

described in two cases of sporadic CJD.

Two different interpretations of TSE-associated myopathies are possible, which,

however, are not necessarily mutually exclusive. The first one envisages that

loss of functional PrPC, due to its continuous conversion into PrPSc, is the major,

or at least a concurrent, cause of muscle damage (Bianchin et al., 2005). In

line with this, stand the higher levels of oxidative stress of both skeletal and

cardiac muscles, and the diminished tolerance to physical exercise of PrPC-KO

mice (Nico et al., 2005; Klamt et al., 2001). On the other hand, given that

myocytes respond to stress conditions with increasing PrPC expression (Sarkozi

et al., 1994; Zanusso et al., 2001), one cannot exclude that, consequent to

injury, availability of higher amounts of PrPC favors the formation and

accumulation of PrPSc in extra-neural organs. Whatever the explanation, the

established capacity of cells to respond to injuries with increasing PrPC levels

(see also Marciano et al., 2004; Shyu et al., 2005a), highlights once again the

possible role of PrPC as major defendant against cell insults, also in non-

neuronal tissues.

28

The heart and ischemia/reperfusion injury

The heart is a muscular organ found in all vertebrates that is responsible for

pumping blood throughout the blood vessels by repeated, rhythmic

contractions. The mammalian heart is derived from embryonic mesoderm

germ-layer cells.

The heart lies in the anterior part of the body cavity, above the diaphragm and

behind the sternum, dorsal to the gut. The heart has four chambers, two

superior atria and two inferior ventricles, communicating by means of two

valves: tricuspid and mitral atrio-ventricular valves. Oxygen-deprived blood

from the vena cava enters the right atrium of the heart and flows into the right

ventricle, from which it is pumped through the pulmonary semilunar valve into

the pulmonary arteries which go to the lungs. Pulmonary veins return the now

oxygen-rich blood to the heart, where it enters the left atrium before flowing

into the left ventricle. Then, oxygen-rich blood from the left ventricle is

pumped out via the aorta, and on to the rest of the body.

The heart is enclosed in a double-walled sac called the pericardium. The heart

is composed of three layers, all of which are rich with blood vessels. The

superficial layer, called the visceral layer, the middle layer, called the

myocardium, and the third layer which is called the endocardium.

The heart is effectively a syncytium, a meshwork of cardiac muscle cells

interconnected by contiguous cytoplasmic bridges. This relates to electrical

stimulation of one cell spreading to neighboring cells. A region of the human

heart called the sinoatrial node, or pacemaker, sets the rate and timing at

which all cardiac muscle cells contract. The impulses also pass to another

region of specialized cardiac muscle tissue, a relay point called the

atrioventricular node, located in the wall between the right artrium and the

right ventricle.

The heart rate (50-80 bpm in humans, ∼500 bpm in mice) is the principal

cause of variation in the cardiac output (blood volume pumped per minute) and

it is regulated by: cardiac muscle automaticity, energetic and chemical factors,

extra-cardiac nervous control; the latter can be inhibitory, through the vagus

nerve fibers and acetylcolin mediator, or stimulatory, through the release by

sympathetic nervous system of catecholamines, such as norepinephrine.

29

Ischemia and reperfusion injury

Ischemic heart deseases arise when there is an imbalance between the

myocardial oxygen demand and blood supply. Ischemia usually progresses

from hypoxia and result in a condition in which the heart is unable to maintain

its rate of cellular oxidation leading to metabolic inbalances. These changes are

initially of a reversible nature (stunning), however, if oxygen is deprived for an

extended period of time, these changes progressively become more severe,

leading to tissue damage and eventually irreversible injury (or infarction).

Furthermore, the severity and progression of ischemia is not solely determined

by the extent of oxygen deprivation but by any other factors including the

relative accumulation of toxic metabolites and ionic imbalances. The reduction

in blood supply during ischemia also limits the removal of these

metabolites/catabolites further contributing to the severe metabolic injury

(Hearse, 1998).

Functional and metabolic modifications in ischemic heart

In the first 8-10 seconds of ischemia, the available oxygen is consumed and

functional and metabolic changes occur (Allen et al., 1990; Allen and Orchard,

1987). These alterations include:

• Contractile activity arrest: the myocardial ATP and phosphate creatine

(PC) supply is able to sustain a short number of contractions. Moreover, in

the early phases, there is an increased accumulation of phosphate and H+

ions, which, linked to a reduced pH, are the causes of the reduced

contractility (Lee et al., 1991);

• Conversion of aerobic metabolism into anaerobic: hypoxia occurring

during ischemia accelerates anaerobic glycolisis;

• Anaerobic glycolisis is able to maintain a sufficient rate of ATP; this is due

to the reduced ATP demand occurring in the first minutes of the ischemic

event. PC is used by the creatine-kinase (CK) to phosphorilate ADP and its

level is rapidly decreased. Ohterwise, the enzyme myokinase can convert

ADP into ATP;

• AMP produced by myokinaseis dephosphorilated by adenosine which is

then degradated to inosin. The diffusion of these nucleotides in the

30

extracellular matrix lead to a reduction in the intracellular nucleotides

pool.

At the beginning of the ischemic period the rate of anaerobic glycolisis is

maximum and determines a high accumulation of lactate (Braasch et al.,

1968; Jennings, 1987). After 60-90 seconds the lactate and reduced

nucleotides accumulation and the diminished pH reduce the rate of the

anaerobic glycolisis. This condition lasts for 40-60 minutes, then the glycolisis

arrests. Other than the lactate accumulation, glycolisis arrest leads to

glucose-6-phosphate, glycerol-3-phosphate and glucose-1-phosphate

accumulation, which leads to an increased osmolarity and edema.

Reversible and non-reversible injury

The salvage of ischemic tissue is most successful when interventions are

initiated as soon as possible (within 15 minutes from the beginning) after

vessel occlusion, resulting in the restoration of blood flow (reperfusion) to the

affected myocardium.

Ischemic damage can be irreversible after 40-60 minutes of ischemia; the ATP

content is reduced to the 90% and glycolisis arrests, reduced co-enzymes

cannot be re-oxidized. In the cell important alterations can be observed in

mitochondria and sarcolemma. Mitochondria undergo swelling, the cristae

appear disorganized and lipids and denaturated protein accumulate in the

matrix. This cellular membrane aberrations are termed “blebs” and are

normally associated to apoptotic cell death. The cause for the blebs formation

is the disorganization and disruption of the cytoscheleton.

The reperfusion injury and the ROS damage

Reperfusion injury refers to damage to tissue caused when blood supply

returns to the tissue after a period of ischemia. The reperfusion injury is

caused by:

• Rapid normalization of tissue pH;

• Rapid normalization of tissue osmolality;

• Re-energisation;

31

• Oxygen radical generation (Hearse, 1991).

The damage includes:

• Lethal arrhytmias and death of the cells previously weakened by ischemia;

• Myocardial stunning, a mechanical disfunction persisting during reperfusion,

even in the absence of histological damage or metabolic disfunctions. It

consists in a setting of the post-ischemic blood pressure at a lower rate

with respect to the pre-ischemic rate. The higher accumulation of Ca2+ ions

contributes to myocardial stunning;

• Loss of cellular proteins (Demaison et al. 1999, Lemasters et al., 1995;

Reimer et al., 1979).

The reduction of the ATP content and pH during ischemia leads to tissue

acidification that activates the Na+/H+ exchanger, in order to restore the

physiological pH. ATP depletion and phosphate ions increase inhibit the Na+/K+-

ATPase, thus leading to an accumulation of Na+ ions in the cell. With reduction

of the Na+ gradient and membrane depolarization, the Na+/Ca2+ exchanger is

turned into its "reverse mode," which leads to cytosolic accumulation of Ca2+;

this accumulation during reperfusion is the cause of activation of degradation

enzymes such as phospholipases, proteases and nucleases, which in turn

damage the myocardial tissue. This effect is demonstrated by the loss of ionic

homeostasis, mitochondrial swelling and de-energization, release of cytosolic

enzymes in the coronary effluent such as the lactate dehydrogenase (LDH) and

the CK, routinely used as markers of myocardial cell death. Once the plasma-

membrane integrity is loss, the cell cannot recover.

Even if the Ca2+ imbalance is an important mediator, ROS generation plays a

major role in I/R damage. Vanden-Hoek and collegues (1997) demonstrated

that the mitochondrial respiratory chain is an important source of ROS in the

act of reperfusion. ROS could directly damage mitochondrial proteins

determining a drop in the mitochondrial potential leading to the activation of

cell death pathways (Levraut et al., 2003). Moreover ROS could affect ionic

pumps at the plasma-membrane level, that together with the ATP deprivation

lead to ionic imbalance.

32

FIGURE 8. Ischemia and Reperfusion damage.

Ischemic Preconditioning

Ischemic preconditioning (IPC) is based on brief episodes of I/R given before a

long I/R episode, which trigger adaptive mechanisms that protect the

myocardium from oxidative injury. In 1986 Murry and coworkers exposed the

myocardium to a ‘preconditioning protocol’ with repetitive short episodes of

regional ischemia. Surprisingly, they found that this protocol induced increased

tolerance to a subsequent long-lasting ischemic period (Sommerschild et al.,

2002). In their study the myocardial infarct size after a 40-min period of

ischemia was reduced from 29% without IPC to 7% with IPC. The cellular basis

for this adaptation is not identified in detail. Different plausible hypotheses

have been suggested and several mechanisms are probably involved.

Hausenloy and Yellon (2006) identified a signalling mechanism based on two

phases: an early phase and a late phase. In the early phase the cellular

memory is related to translocation of PKC from cytosol to cellular membranes.

This causes a more rapid activation of PKC during the prolonged ischemic

period (Downey et al., 1994). After a certain time, PKC will re-translocate to

cytosol, and the memory will disappear. During the late phase of

preconditioning it is believed that cellular memory is related to synthesis of

proteins (Marber et al., 1997). The cells will need some time to produce new

33

proteins, therefore it needs a while before protection occurs. Synthesis of heat-

shock proteins has been also implicated in the mechanisms of IPC (Housenloy

and Yellon, 2006).

34

35

MATERIALS AND METHODS

Mouse models

In this study we have used 3 to 4 month-old male congenic (FVB) mice

expressing different PrPC amounts:

• FVB wild type (WT) (Harlan, Milano).

• Knock-out (KO) for the gene codifying for PrPC on an almost pure FVB

genetic background (F10 mouse line, kindly provided by Dr. G. Mallucci,

MRC Prion Unit, University of Leicester, UK).

• A transgenic mouse line over-expressing (OE) PrPC 3 to 4-folds the natural

levels (TG37 line, kindly provided by prof. J. Collinge, Imperial College,

London, UK).

PrPC-KO mice derived from Zurich I PrPC-KO mice, originally obtained by the

group led by Charles Weissman on a hybrid genotype SV129X/C57-Bl6

(Büeler et al., 1992). These mice have been cross-bred for ten generations in

hemizygosis with FVB WT (PrP+/+) mice. The PrP+/- littermates (N10) have

been inter-bred to generate the PrPC-KO line (F10) with an almost pure

(theoretically >99.9%) FVB genotype (Mallucci et al., 2002). PrPC-OE TG37

transgenic mice have been obtained by transgenesis on PrPC-KO F10 mice

(Mallucci et al., 2002).

All aspects of animal care and experimentation were performed in accordance

with the Guide for the Care and Use of Laboratory Animals published by

European and Italian (D.L. 116/92) laws concerning the care and use of

laboratory animals. The Authors’ Institution has been acknowledged by the

Italian Ministry of Health for the use of mice for experimental purposes, and

the experimental protocols were approved by the Ethical Committee of the

University of Padova.

36

The Langendorff model and protocols for the perfusion of isolated

hearts

Mice were weighed and then anaesthetized with an intraperitoneal injection of

Zoletil 100 (30 mg/kg body weight). Hearts were rapidly excised from the

thorax and placed in a cold (4°C) bicarbonate buffer (see below). Under an

illuminated magnifier, the aortic opening was immediately cannulated and tied

on a stainless steel blunt needle. Perfusion was performed in the non-

recirculating Langendorff mode (Langendorff, 1985). In the Langendorff

preparation, the heart is perfused in a retrograde (reverse) fashion,

maintained at the desired flow rate by a pump, with a nutrient rich,

oxygenated solution. The pressure of the solution causes the aortic valve to

shut and the perfusate is then forced into the coronary vessels, from these to

the right atrium and then to the right ventriculm. The perfusate is then sent

out from the pulmonary arteries (Fig. 1). This may allow the isolated heart to

beat for up to several hours.

The hearts were perfused with bicarbonate buffer gassed with 95% O2-5%

CO2 (pH 7.4) (perfusion buffer, PB). The bicarbonate buffer contained (in mM)

115.0 NaCl, 4.75 KCl, 2.15 KH2PO4, 25.0 NaHCO3, 0.65 MgSO4, 1.69 CaCl2,

and 11 glucose (Barbato et al., 1996). The PB was warmed at 37°C through a

water-jacketed glass cylinder/heat exchanger system with a warming/cooling

bath, and was circulated (at a constant flow rate of 5 ml/min) by a water

pump. The tissue temperature was maintained at 37°C by suspending the

heart in a water-jacketed chamber (Fig. 2).

37

FIGURE 1. Scheme of circulation of the perfusion buffer in an isolated heart subjected

to the Langendorff mode of perfusion.

FIGURE 2. Langendorff perfusion setup. The red arrow indicates the isolated organ.

Perfusion Protocols

At the beginning of the experiment, hearts were subjected to a 5 min-

stabilization period by perfusion with PB, a sufficient time period to wash

hearts from blood. During this step the heart beating and the coronary out-

flow were constantly monitored. When beating and/or coronary out-flow were

absent or interrupted, hearts were discarded. Hearts were then subjected to

the following perfusion protocols.

(i) Ischemia/Reperfusion (I/R), consisting of 40 min of global ischemia

(achieved by stopping the coronary flow), followed by 15 min of reperfusion

with the PB (gradually re-established at the flow-rate of 5 ml/min) (Fig. 3A).

The experimental significance of the I/R protocol is described in details in the

Introduction.

(ii) IPC protocol. A general description of the protocol, the significance, and

the effects of IPC is given in the Introduction. In our case, IPC consisted of

three episodes of 5 min of ischemia and 5 min of reperfusion, followed or not

by the previously described I/R protocol (Fig. 3B).

38

(iii) Perfusion with H2O2. This protocol was aimed at challenging hearts with a

non-ischemic oxidative injury, i.e. at probing the effects of oxidative stress in

hearts that were not previously weakened by a prolonged period of global

ischemia. To this purpose, isolated hearts were perfused for 15 or 30 min with

PB containing 1mM H2O2, immediately after the equilibration step (Fig. 3C).

Because of the high instability of H2O2, titration of the stock solution (10 M) was

always checked spectrophotometrically before each experiment (see below).

A

B

C

FIGURE 3. Schematic description of the perfusion protocols for I/R (A), IPC (B), and

perfusion with H2O2 (C).

When necessary, N-2-mercaptopropionyl-glycine (MPG, 1 mM) was added to

the PB. Control (normoxic) hearts were subjected to the 5 min-stabilization

protocol only.

For subsequent analyses, 5 ml-samples of the coronary effluent were

collected at 1 min-intervals during post-ischemic reperfusion, or perfusion

39

with H2O2, as previously described (Di Lisa et al., 2001; Carpi et al., 2009). At

the end of the experiments, hearts were quickly removed from the set-up,

dried with Whatman blotting paper to remove all PB residues, and cut into two

parts that were weighed separately. One part, homogenized (50 mg/ml) in PB

containing 0.5% (w/v) Triton X-100, was used to evaluate the release of LDH

(see below). The other part was flash-frozen, and stored in liquid nitrogen for

further biochemical/histochemical analyses.

H2O2 titration

For H2O2 titration we first diluted a 10 M stock solution to 100 mM. Then, 30, 60,

and 90 µl of the 100 mM solution were further diluted to a final volume of 3 ml.

The spectrophotometrical measurements were performed in quartz cuvettes at a

wavelength of 240 nm (Spectrophotometer Cary 50 Bio, Varian). Absorbance

readouts for each H2O2 diluition had to be comparable to those reported in the

following table. When this was not the case, the H2O2 stock solution was replaced.

Aliquots Absorbance Concentration

30 µl 0.4 1 mM

60 µl 0.8 2 mM

90 µl 1.2 3 mM

Estimation of LDH release

To assess the loss of heart viability after post-ischemic reperfusion, or during

perfusion with H2O2, we evaluated the release of the soluble enzyme LDH in

the coronary effluent, as an index of myocardial cell death (Di Lisa et al.,

2001). To this end, LDH activity was evaluated in both the coronary effluent,

collected during the (re)perfusion period, and the corresponding heart

homogenate. LDH activity was then expressed as the percentage activity in

the coronary effluent over the total (i.e., effluent + heart homogenate)

activity (Carpi et al., 2009; Schluter et al., 1991):

100(%)hom

×+

=ogenateeffluent

effluenteffluent LDHLDH

LDHLDH

40

FIGURE 4. Reaction mediated by LDH. LDH catalyzes the interconversion of pyruvate and

lactate with concomitant interconversion of NADH and NAD+.

Taking advantage of the functional properties of LDH, which catalyzes the

interconversion of pyruvate and lactate, with concomitant interconversion of

NADH and NAD+ (Fig. 4), LDH activity was determined spectrophotometrically

(Multiskan Ex. Lab System) by measuring the oxidation of NADH (absorption at

340 nm), according to a classic procedure (Bergmeyer et al., 1974; Di Lisa et

al., 2001). The essay was performed (at room temperature, RT) into 96 wells

plates. Briefly, the following reaction solution (180 µl), prepared just before

use, was dispensed into each well: 81.3 mM Tris/HCl, pH 7.2, 203.3 mM NaCl

(TRIS/NaCl solution), 240 µM NADH. Then, either homogenate (diluted 1:5 or

1:10 in Tris/NaCl solution) or coronary effluent samples (40 µl) were added to

each well. A first reading was performed as a blank. Reaction was then

started by adding 50 µl of 10 mM pyruvate (in Tris-NaCl solution) to the

reaction mixture, and the reaction kinetics were monitored for 4 min. LDH

activity is directly proportional to the oxidation of NADH, which can be easily

extrapolated by the Lambert-Beer equation (A = ε x C x L, where A is the

absorbance, ε is the molar extinction coefficient, C is the molar concentration

of NADH, L is the path length).

In I/R protocols, the total activity of LDH released in the coronary effluent at

the end of reperfusion was calculated according to the formula:

∑=

=15

1iieffluent LDHLDH

(in which LDHi is the LDH activity in the coronary effluent collected at the ith

min of reperfusion)

41

For perfusion with H2O2, the cumulative activity of LDH released in the

coronary effluent during the perfusion time-period was calculated according to

the formula:

∑=

=n

ii

neffluent LDHLDH

1

(in which LDHneffluent is the LDH activity in the coronary effluent after n min of

perfusion, being n=1-15/30)

Preparation of mitochondria from mouse hearts

Mice were killed by cervical dislocation and the hearts were quickly removed

and placed in ice-cold isolation solution (IS) containing 250 mM sucrose, 0.1

mM EGTA, 10 mM Tris/HCl, pH 7.4. Mitochondria were isolated within 30 min

after tissue dissection. Briefly, hearts were minced and washed several times

in IS to eliminate blood, homogenized by an Ultra-Turrax homogenizer in 3 ml

of IS, and then centrifuged at 900�g (4°C, 10 min) to remove the cell debris.

The supernatant was then centrifuged at 8000�g (4°C, 10 min). The cytosolic

fraction (supernatant) was kept for enzymatic activity assays and Western

blot analyses. The mitochondrial fraction (pellet) was resuspended either in a

buffer containing 50 mM potassium-phosphate (pH 8.0) and 0.5% Triton X-

100, for enzymatic assay, or in a sample buffer (SB) containing 2% (w/v)

sodium dodecyl-sulphate (SDS), 5% (w/v) glycerol, 125 mM Tris–HCl (pH

6.8), for Western blot analyses. The cytosolic fraction was diluted with 4 part

of methanol and the proteins precipitated by incubation over-night (ON) at -

80°C, then centrifuged at 8000�g (4°C, 30 min); the pellet was resuspended

in the SB for Western blot analyses. The protein content was measured using

a BCA protein assay kit (Thermo Scientific, Rockford, IL, USA).

42

SDS-PAGE and Western blotting

Sample preparation

Heart biopsies, stored in liquid nitrogen, were cut into 10 µm cryosections at -

25°C, then added with 1 ml of ice-cold phosphate-buffered saline (PBS), pH

7.2 containing 0.5 mM EDTA. Just before use, the solution was stirred under

vacuum to reduce the oxygen tension. To avoid protein degradation the entire

procedure was performed at 4°C. The suspension was then vortexed and

centrifuged (10 min) at 12000·g (Eppendorf Centrifuge 5417 R). The resulting

pellet was homogenized in SB by means of a Teflon potter, and the proteins

were denatured by 10 min boiling. When SDS-PAGE was performed under

reducing conditions, dithiothreitol (DTT, 100 mM) was added to the SB.

Conversely, when SDS-PAGE was performed under non-reducing conditions,

N-ethylmaleimide (NEM, 1mM) was added to the SB, to avoid artefacts due to

the oxidation of thiol groups in vitro. The total protein content of

homogenates was determined by the BCA protein assay kit.

Deglycosylation with peptide N-glycosidase F

For glycan removal with peptide N-glycosidase F (PNGase-F), proteins (40 µg)

denatured in SB were diluted five-fold in deglycosylation buffer (20 mM

sodium phosphate, 0.8% NP-40, 1.2% β-mercaptoethanol, 30 mM EDTA, and

1% (w/v) Triton X-100, pH 8.0). The protein suspension was then incubated

under agitation (24 h, 37°C) in the absence or in the presence of PNGase-F

(Roche) (2.5 U). Proteins were then diluted in SB and resolved by SDS-PAGE.

SDS-PAGE

SDS poly-acrylamide gel electrophoresis was performed in the mini gel format

(7 cm gel size), using the BioRad Protean III electrophoresis system (BioRad,

Hercules, CA, USA). The acrylamide concentration was always set to 4%

(w/v) in stacking gels, while – in separating gels – it was either 12% (for PrPC

and tropomyosin (Tm) immunoblotting), 8% (for p66Shc) or 15% (for SOD1

and Mn SOD (SOD2)). The following amounts of total proteins were loaded

into each lane: 15 µg for PrPC; 50 µg for SOD1 and 30 µg for SOD2, from

cytosolic or mitochondrial fractions, respectively; 50 and 20 µg for p66Shc,

from total heart homogenates or mitochondrial fractions, respectively; 5 µg

43

and 50 µg for the monomeric and the dimeric form of Tm, respectively (see

below).

Western blotting

After SDS-PAGE, proteins were electrophoretically transferred onto 0.45 µm

pore-size nitrocellulose membranes (Bio-Rad) (350 mA, 1 h, 4°C) in a buffer

containing 25 mM Tris, 192 mM glycine, 0.03% SDS and 20% methanol. The

electro-blotting efficiency was always checked by staining with Red Ponceau S

(Sigma). Unspecific binding sites on nitrocellulose membranes were blocked

by incubation (1h, RT) in PBS containing 0.1% (w/v) Tween-20 (PBS-T) and

either 3% (w/v) bovine serum albumine (BSA) (for PrPC immunodetection), or

5% (w/v) non-fat dry milk (for tropomyosin, p66Shc, SOD1 and SOD2).

Membranes were then incubated (ON, 4°C) with the desired antibody diluted

in PBS-T containing either 1%, 3% BSA or 5% non-fat dry milk. After

extensive washing with PBS-T, membranes were incubated (1h, RT) with

horseradish peroxidase-conjugated anti-mouse or anti-rabbit secondary

antibodies (Santa Cruz Biotechnology), diluted (1:4000) in PBS-T containing

either 1% BSA or 5% non-fat dry milk. Immunoreactive bands were visualized

using a chemiluminescence detection system (ECL, Pierce), and acquired by a

digital Kodak imaging workstation, checking that exposure times were within

the linear range of detection. Densitometry was performed on the digitalized

images of immunoblots using the Kodak 1D 3.5 computer program (Kodak,

New Haven, USA). The intensity values are the result of the sum of each pixel

intensity composing the analyzed area, considered in each lane, subtracted

for the background of the same area.

Antibodies

For Western blot analyses, the following mono- (M) and polyclonal (P)

antibodies (ab) were used (dilutions are indicated in parenthesis). Anti-PrP

mouse Mab 8H4 (1:7000) (a kind gift of Dr. M.S. Sy, Case Western

University, Cleveland, OH), raised against the human (173–185) sequence;

anti-tropomyosin mouse Mab CH1 (1:2000) (Sigma); anti-SOD1 rabbit Pab

(1:1000) (Abcam); anti-SOD2 rabbit Pab (1:2000) (Sigma); anti-Shc rabbit

Pab (1:2000) (BD Transduction Laboratories).

44

Estimation of tropomyosin oxidation

Western blot analyses were also used to quantify the oxidation of Tm, i.e.,

the relative presence in the heart homogenates of Tm dimers resulting from

the formation of disulphide cross-bridges (DCB). Of consequence, SDS-PAGE

for Tm immunodetection were always run under non-reducing conditions.

Given that, under the used conditions, the monomeric, reduced form of Tm

was always predominant over the oxidized (dimeric) form, two gels were run

in parallel for each set of samples. In one gel, to be blotted for evaluating the

immuno-intensity of the Tm dimers, 50 µg of total proteins were loaded onto

each well. Instead, in the other gel, which was used for evaluating the

immuno-signal of the Tm monomers, 5 µg of total proteins were loaded onto

each well. Quantitative analysis of the degree of tropomyosin oxidation was

performed on the densitometric values of the bands detected in immunoblots.

In particular, the intensity of the approx. 80 kDa band, indicative of Tm

dimers, therefore reflecting the formation of inter-molecular DCB, was

normalized to the band intensity of the corresponding Tm monomer (Fig. 5).

FIGURE 5. Schematic representation of the reversible oxidation of Cys residues leading to homodimers formation (cross-linking).

In situ superoxide detection

Tissue staining with hydroethidium (HE, Sigma) was used to evaluate the

accumulation of the �O2- in heart cryosections (Oudot et al., 2006). In the

presence of superoxide, HE is converted into the fluorogenic molecule 2-

hydroxyethidium, which becomes fluorescent upon binding to nuclear DNA

(Zielonka et al., 2008) (Fig. 5). Freshly prepared heart cryosections (10 µm)

45

were incubated (30 min) in a light-protected humidified chamber at 37°C in

the presence of HE (10 µM in DMSO), then rinsed twice with PBS, and covered

with a coverslip. The fluorescence images were acquired, with a constant

exposure time, by means of an Olympus IMT-2 inverted microscope, equipped

with a Xenon lamp and a 12-bit digital cooled CCD camera (Micromax,

Princeton Instruments, Monmouth Junction, NJ, USA). Fluorescence was

detected with 510–560 nm excitation and 590 nm emission filters. Automatic

computer-based analysis was performed with the same fluorescence threshold

for all sections. Four fields were randomly selected for each tissue section,

and the mean of their fluorescence was normalized to the intensity of a

reference sample, considered as 100%.

FIGURE 6. Proposed mechanism of 2-hydroxyethidium (2-OH-E+) formation from the reaction between hydroethidium (HE) and �O2

-.

46

Enzymatic activity assays

Superoxide dismutase activity assay

Cu/Zn SOD (SOD1) and Mn SOD (SOD2) activities were measured on the

cytosolic and mitochondrial fractions, respectively, prepared as described

previously. SOD activity was determined by quantifying

spectrophotometrically the inhibition of xanthine/xanthine oxidase–induced

cytochrome c reduction (McCord and Fridovich, 1969). The reaction mixtures

contained 50 mM potassium phosphate buffer (pH 8.0), 0.1 mM EDTA, 50 µM

cytochrome c, 100 µM xanthine and 50 µg of cytosolic proteins or 5 µg of

mitochondrial proteins (previously diluted in the reaction medium), for SOD1

and SOD2, respectively. Reactions were performed in plastic cuvettes at

25°C, and were started by adding xanthine oxidase (0.1 U/ml). Cytochrome c

reduction rate was determined by the slope of absorbance increase at a

wavelength of 550 nm (Spectrophotometer Cary 50 Bio, Varian). The inhibition

caused by heart homogenates was normalized to that afforded by 1 U of

purified SOD (Sigma-Aldrich). Data were then reported as percentage of the

mean value of WT samples.

Catalase activity assay

To determine CAT activity, tissues were homogenized in 50 mM phosphate

buffer (pH 7.0), 0.5% (w/v) Triton X-100 by means of a Teflon potter, and

the resulting suspension was centrifuged at 3000�g (10 min, 4°C) to discard

tissue debris. The supernatant was used for the enzymatic assay, after

determining the total protein content by the BCA protein assay kit. CAT

activity was assayed by measuring the consumption of H2O2 through the

absorbance decrease at a wavelength of 240 nm (Aebi, 1984)

(Spectrophotometer Cary 50 Bio, Varian). The reaction medium contained 50

mM phosphate buffer (pH 7.0) and 215 µg of total proteins. Reactions,

performed in a quartz cuvette at 25°C, were started by adding 10 mM H2O2.

Measurements were calibrated by means of a standard curve, generated by

using known amounts of purified CAT (Sigma-Aldrich), and then data were

expressed as percentage of the mean value obtained with WT hearts.

47

AIMS AND RATIONALE

The major purpose of this work was to contribute to the understanding of the

physiologic functions of PrPC. This issue is relevant not only for cell biology per

se, but especially for elucidating possible mechanisms of TSE-related

neurodegeneration, and the future development of suitable therapeutic

strategies. As clearly detailed in the Introduction, the richness of data on the

putative PrPC activities has not resulted in a comprehensive picture. This is

ascribable to the fact that use of cell (mainly neuronal) models have

contributed overwhelming – often confused – insights into the cellular and

molecular events which PrPC may come to play in, whereas animal models –

with few exceptions – have provided no concrete advancement in the

understanding of the true physiologic importance of the protein. In the

attempt to fill the gap between cell models and whole animals, in this work we

have used, for the first time in the prion field, an intact isolated organ – the

heart – to probe the function of PrPC against cell damage and, more

specifically, oxidative stress. The rationale of this choice relies on the notion

that PrPC is abundantly expressed in the heart, and on recent findings

correlating PrPC pathophysiology to the skeletal and cardiac muscles. In

addition, isolated hearts has the advantage of retaining the cell properties and

physiological environment, while being by far more amenable to experimental

manipulations than live animals. Thus, we have subjected hearts isolated from

mice with a different PrP genotype (WT, PrPC-KO and PrPC-OE) to different

perfusion protocols, that allowed us to specifically probe different myocardial

properties, ranging from the response to post-ischemic reperfusion damage,

which involves a broad set of physiologic parameters, to ischemic

preconditioning and non-ischemic oxidative insult. In the attempt to provide

explanation for our ex-vivo data, we have also analysed the efficiency of some

myocardial pro- and anti-oxidant systems in hearts with different PrPC levels,

and monitored the cell fate of PrPC in hearts subjected to the different stress

conditions.

48

49

RESULTS

As detailed previously, we used in our experiments WT FVB mice and

genetically modified congenic mice. These included PrPC-KO mice and a

transgenic mouse line (OE) reported to overexpress PrPC three-to-four fold

the natural level. The amounts of PrPC in hearts isolated from the different

used strains were routinely assessed by Western blot analysis (Fig. 1). Such

experiments showed that, though to a lesser extent than in the brain (lane 1),

PrPC could nonetheless be readily identified in WT hearts (lane 3), and that

higher amounts of the protein were present in the heart of PrPC-OE mice (lane

4) (OE to WT PrPC ratio being 3.0 ± 0.2, n = 3). Significantly, in hearts from

either WT and PrPC-OE mice, PrPC was found mainly in the mature, di-

glycosylated form. Conversely, no immunosignal was present in PrPC-KO

hearts (lane 2).

WTKO OE

35.5

29

MW (kDa)

D

M

U

WT Brain

1 2 3 4

WTKO OE

35.5

29

MW (kDa)

D

M

U

WT Brain WTKO OE

35.5

29

MW (kDa)

D

M

U

WT Brain

1 2 3 4

FIGURE 1. Immunodetection of PrPC in the heart of the different mouse strains. Heart

samples from 4 month-old WT, PrPC-KO and PrPC-OE male mice were homogenized and proteins

(15 µg per lane) were resolved on a 12% SDS-PAGE gel under non-reducing conditions,

electroblotted onto a nitrocellulose membrane, and then probed with anti-PrP Mab 8H4. For

comparison, a WT brain homogenate (5 µg of proteins) was also loaded in the gel. The PrPC

immunosignal is readily appreciable in the heart from WT mice, and is significantly increased in

the PrPC-OE heart. Conversely, no signal is evident in the KO sample. Arrows on the right indicate

the different PrPC glycoforms, i.e., un-glycosylated (U), mono-glycosylated (M) and di-

glycosylated (D). Molecular mass standards are reported on the left. Shown blot is representative

of 3 independent experiments that yielded comparable results.

50

Evaluation of the myocardial damage induced by I/R protocols in isolated hearts with different PrPC levels

Hearts isolated from PrPC-OE mice are protected against loss of viability

induced by post-ischemic reperfusion

A by now long tradition of experimental cardiology has established that

prolonged periods of no-flow ischemia in isolated hearts cause a complex set

of myocardial dysfunctions, including impairment of mitochondria, ATP

depletion, pH changes and ion dyshomeostasis. However, it is also accepted

that the major cause of injury in ischemic hearts is the oxidative damage that

follows the massive ROS production occurring after the re-establishment of

the coronary flow (Zweier et al., 1987; Vanden-Hoek et al., 2000).

This notion prompted us to assess whether PrPC was capable to protect the

myocardium from post-ischemic reperfusion. To this end, we first subjected

hearts –isolated from mice expressing different amounts of PrPC - to a 40

min-period of ischemia (I) followed by 15 min of reperfusion (R); then, we

monitored the occurrence of cell death by quantifying the release of LDH in

the coronary effluent over the entire reperfusion time-period. As shown in the

bar diagram of Fig. 2, and in Table 1 (first line), we found that the I/R

protocol produced ∼27% loss of cell viability in WT samples, and,

unexpectedly, that a similar value pertained to PrPC-KO hearts. In contrast,

PrPC-OE hearts had a significantly reduced cell death (∼ 22%).

51

WT KO OE

I/R 27.2 ± 2.1 (n = 38)

27.2 ± 1.8 (n = 38)

21.7 ± 1.4 (n = 35)

IPC + I/R 18.3 ± 2.4 (n = 14)

15.4 ± 2.4 (n = 12)

17.5 ± 2.0 (n = 9)

IPC (+MPG) +I/R 28.6 ± 5.4 (n = 4)

26.4 ± 4.9 (n = 4)

22.5 ± 4.2 (n = 4)

15 min H2O2 6.4 ± 1.6 (n = 8)

12.5 ± 2.0 (n = 11)

6.5 ± 2.2 (n = 11)

TABLE 1. LDH activity in the coronary effluent of hearts subjected to the different

perfusion protocols. LDH activity, expressed as the percentage of LDH activity in the coronary

effluent over the total (i.e., effluent + heart homogenate) activity, is a marker of cardiac cell

death. In the first column are indicated the protocols used; the second, third and fourth column

refer to the LDH activity in the effluent of WT, PrPC-KO and PrPC-OE hearts, respectively. Data are

mean ± standard error of the mean (s.e.m.). The number of experiments for each condition is

reported in parenthesis. For further details, see Materials and Methods.

0

5

10

15

20

25

30

35 ***

WT KO OE

LDH

act

ivity

in th

e co

rona

ry e

fflue

nt (

%)

0

5

10

15

20

25

30

35 ***

WT KO OEWT KO OE

LDH

act

ivity

in th

e co

rona

ry e

fflue

nt (

%)

FIGURE 2. Myocardial loss of viability induced by post-ischemic reperfusion is reduced

in PrPC-OE hearts. Hearts isolated from 4 month-old male WT, PrPC-KO and PrPC-OE mice were

subjected to 40 min of global ischemia (I) followed by 15 min of reperfusion (R). Loss of viability

was evaluated as the percentage of total LDH activity released in the coronary effluent during

post-ischemic reperfusion with respect to the total (coronary effluent + tissue homogenate) LDH

activity of the heart. Data indicate a significant reduction of I/R-induced myocardial damage in

PrPC-OE hearts with respect to both WT and PrPC-KO hearts. Values are mean ± s.e.m., n = 38

(WT and KO), 35 (OE). * p < 0.05, ** p < 0.01, Student’s t-test.

52

The over-expression of PrPC reduces the degree of oxidative stress caused by

post-ischemic reperfusion

Since excess production of ROS is the main feature of post-ischemic

reperfusion, we monitored the oxidative stress in the different hearts

subjected to I/R using two alternative methods. With the first, we evaluated

the oxidation of tropomyosin (Tm), a reliable index of oxidative damage of

cardiac contractile proteins (Canton et al., 2004; 2006), which we quantified

by assessing the amount of dimers that are generated by inter-molecular

disulphide cross bridges following the oxidation of the only Cys residue

present in each Tm molecule (Canton et al., 2006; for details see Fig. 5 in the

Materials and Methods section). As illustrated in Fig. 3, reporting data from

immunoblot experiments (under non reducing conditions) of heart samples

with an anti-Tm antibody, the degree of Tm dimers (of ∼80 kDa) produced by

the I/R injury was considerably reduced (by ∼40%) in PrPC-OE hearts

compared to the other mouse strains.

53

WT KO OE Cntr

80 Oxidized tropomyosin(dimer)

40Tropomyosin(monomer)

MW (kDa)WT KO OE Cntr

80 Oxidized tropomyosin(dimer)

40Tropomyosin(monomer)

MW (kDa)

0

20

40

60

80

100

120

140

WT KO OE

Nor

mal

ized

ban

d in

tens

ity (

%)

**

0

20

40

60

80

100

120

140

WT KO OEWT KO OE

Nor

mal

ized

ban

d in

tens

ity (

%)

***

FIGURE 3. Oxidation of tropomyosin after I/R injury is decreased in PrPC-OE hearts.

Hearts from mice with different PrPC levels, subjected to I/R (40 min/15 min), were homogenized,

proteins resolved onto 12% SDS-PAGE under non-reducing conditions, and then electro-blotted

onto nitrocellulose membranes. Tm oxidation, consequent to post-ischemic reperfusion, was

evaluated, by Western blotting with the anti-Tm Mab CH1, as the appearance of a high molecular

weight (∼80 kDa) immuno-reactive band due to the formation of inter-molecular disulphide cross

bridges. To better appreciate both the oxidized and the reduced (monomeric, ∼40 kDa) forms of

Tm, each set of samples was always run in double, by loading 5 µg and 50 µg of total proteins

into each lane for the detection of the Tm monomers and dimers, respectively (see Materials and

Methods for further details). In the upper panel, the results of a Western blot analysis,

representative of one out of five independent experiments, are reported. A sample of a normoxic

heart (Cntr) from a WT mice was also loaded into the gel as a negative control. In the lower

panel, the results of the densitometric analysis are reported. Values (mean ± s.e.m., n = 5 for

each PrPC genotype) are expressed as percentage of the WT samples. Tm oxidation after I/R

injury is largely reduced in PrPC-OE with respect to WT and PrPC-KO hearts. * p < 0.05, Student’s

t-test.

54

We then evaluated ROS accumulation in I/R-treated hearts by subjecting

tissue cryosections to the HE staining assay. HE reacts specifically with the

superoxide anion (�O2-), generating the fluorogenic 2-hydroxyethidium

product that can be easily detected and quantified by fluorescence microscopy

(Oudot et al., 2006). In line with the previous results, heart cyosections from

PrPC-OE mice were significantly less fluorescent than the WT and PrPC-KO

counterparts (Fig. 4). No fluorescent signal was detected in cyosections of

normoxic hearts, or when cryosections were treated with recombinant SOD

before the HE-staining procedure (data not shown).

KOWT OEKOKOWTWT OEOE

0

20

40

60

80

100

120

140

WT KO OE

****

Nor

mal

ized

fluo

resc

ence

inte

nsity

(%

)

0

20

40

60

80

100

120

140

WT KO OEWT KO OE

****

Nor

mal

ized

fluo

resc

ence

inte

nsity

(%

)

FIGURE 4. �O2- accumulation in cardiomyocytes after I/R injury is decreased in PrPC-OE

hearts. Tissue cryosections of hearts with different PrPC levels, subjected to I/R (40 min/15 min),

were processed for HE staining to quantify superoxide accumulation. In the upper panel,

fluorescence micrographs of HE-stained cryosections from WT, PrPC-KO and PrPC-OE mice,

representative of five independent experiments, are reported. In the lower panel, the quantitative

analysis of fluorescence intensities is shown. Values (mean ± s.e.m., n = 5 for each PrPC

genotype) are expressed as percentage of the WT samples. Accumulation of �O2- after I/R injury

55

is largely reduced in PrPC-OE with respect to WT and PrPC-KO hearts. * p < 0.05 and *** p <

0.001, Student’s t-test.

As mentioned, the considerable amounts of ROS produced in the reperfusion

step are fatal to the viability of myocardial cells, especially because cells get

undermined during the ischemic period. The aggressiveness of the entire I/R

protocol, involving a very complex set of myocardial dysfunctions, could thus

explain why, only when over-expressed, PrPC elicited an appreciable

protection over cell death. Accordingly, this may also explain why no

difference of cell death could be detected between PrPC-KO hearts and hearts

that express normal amounts of PrPC. For the same reasons, data acquired

with two different markers of oxidative stress, i.e., oxidation of myofibrillar

proteins and accumulation of �O2-, suggest that one, albeit perhaps not the

only, possible reason of the enhanced viability of I/R PrPC-OE hearts is the

capacity of PrPC to counteract the production, and/or accumulation, of ROS.

PrPC performs anti-oxidant functions in the heart

The absence of PrPC enhances the protective effects of ischemic

preconditioning

To explore in more details the direct involvement of PrPC against ROS, hearts

were subjected to ischemic preconditioning (IPC). IPC consists of a short

sequence of I/R brief episodes preceding a long ischemic period, which

significantly protects the heart from the damage induced by post-ischemic

reperfusion (Pain et al., 2000). The protective effects of IPC have been

ascribed to the production of sub-lethal amounts of ROS during the

preconditioning phase (Pain et al., 2000), which would trigger defence

mechanisms, and eventually counteract the large burst of ROS generated at

the onset of post-ischemic reperfusion (Vanden-Hoek et al., 1998).

Accordingly, known anti-oxidant compounds are able to diminish or suppress

the beneficial effect of IPC (Liu et al., 1998; Pain et al., 2000). We thus

reasoned that, were indeed PrPC part of the cell apparatus protecting against

the production and/or accumulation of ROS, the benefits of IPC should have

been found to inversely correlate with the levels of PrPC. This is precisely what

we observed after treating hearts with 3 cycles of short I/R episodes (5 min/5

56

min), followed by the standard I/R challenge (40 min/15 min). As reported in

Fig. 5 (white bars) and Table 1 (second line), the protection - estimated by

the relative reduction of LDH release in the coronary effluent - was maximal

(∼45%) in the absence of PrPC, intermediate (∼30%) in WT hearts, and totally

abrogated in PrPC-OE hearts. Interestingly, the degree of cardioprotection by

IPC in WT hearts is comparable to that afforded by PrPC overexpression in the

absence of IPC (Fig. 5, compare white (WT) and grey (OE) bars).

Next, we investigated if this observed difference in IPC efficiency could be

specifically ascribed to the anti-ROS potential of PrPC. We found that this was

the case for two reasons. One was that the presence, during IPC, of the

potent free radical-scavenger N-2-mercaptopropionyl-glycine (MPG)

completely abolished the preconditioning protective effects, irrespective of the

PrPC genotype (Fig. 5, red bars and Table 1, third line). The other was the

demonstration that, during IPC, the heart content of ROS was dependent on

PrPC expression. Indeed, Fig. 6 reports that the PrPC-KO heart, stained with

HE at the end of the preconditioning step, has more abundant ROS compared

to WT hearts. Further, a slight, though not significant, reduction of ROS was

observed in PrPC-OE hearts with respect to the WT counterpart.

57

0

5

10

15

20

25

30

35

WT KO OE

**

****

LDH

act

ivity

in th

e co

rona

ry e

fflue

nt (

%)

0

5

10

15

20

25

30

35

WT KO OE0

5

10

15

20

25

30

35

WT KO OE

****

********

LDH

act

ivity

in th

e co

rona

ry e

fflue

nt (

%)

FIGURE 5. PrPC reduces the protective effect of IPC on myocardial cell loss. Isolated

hearts were subjected to I/R without (grey bars) or with (white bars and red bars) a protocol of

IPC before the step of prolonged no-flow ischemia. Ischemic preconditioning (IPC) (3 cycles of 5

min of ischemia followed by 5 min of reperfusion) was run in the absence (white bars) or in the

presence (red bars) of 1 mM MPG in the perfusion buffer. The protection afforded by IPC was

evaluated as a decrease in the release of LDH during post-ischemic reperfusion. The beneficial

effect of IPC is maximal (∼45%) in PrPC-KO, intermediate (∼30%) in WT and non-significant in

PrPC-OE hearts. In the presence of MPG, the effect of IPC is completely abolished, irrespective of

PrPC expression levels. Values are mean ± s.e.m.; grey bars: n = 38 (WT and KO), 35 (OE);

white bars: n = 14 (WT), 12 (KO), 9 (OE); red bars: n = 4 (WT, KO, OE). * p < 0.05, ** p <

0.01, Student’s t-test. Other experimental details are as in the legend to Fig. 2.

58

KO OEWT KOKOKO OEOEOEWTWTWT

0

50

100

150

200

WT KO OE

Nor

mal

ized

fluo

resc

ence

inte

nsity

(%

)

** ***

0

50

100

150

200

WT KO OEWT KO OE

Nor

mal

ized

fluo

resc

ence

inte

nsity

(%

)

** ***

FIGURE 6. PrPC reduces ROS production during IPC. Isolated hearts were subjected to the

IPC protocol (3 cycles of 5 min of ischemia followed by 5 min of reperfusion) and then analysed

for the accumulation of �O2- by means of the HE-staining assay. In the upper panel, fluorescence

micrographs of HE-stained cryosections from WT, PrPC-KO and PrPC-OE mice, representative of

four independent experiments, are reported. In the lower panel, the quantitative analysis of

fluorescence intensities is shown. Values (mean ± s.e.m., n = 4 for each PrPC genotype) are

expressed as percentage of the WT samples. Accumulation of the �O2- after IPC is considerably

increased in PrPC-KO hearts with respect to the PrPC-expressing counterparts. ** p < 0.01 and

*** p < 0.001, Student’s t-test. Other experimental details are as in the legend to Fig. 5.

PrPC protects the heart from non-ischemic oxidative injury

To corroborate further the anti-oxidant potential of PrPC, we used a protocol

based on non-ischemic oxidative damage, i.e. perfusion with H2O2. In this

experimental paradigm, isolated hearts, after the equilibration step, were

perfused for 15 or 30 min with a buffer containing 1mM H2O2. Though severe,

this means of oxidative challenge is conceptually simpler that the I/R protocol

in that it allows to test the anti-ROS potential of myocardial cells that have

not been weakened by a previous ischemic challenge. In fact, at difference

from the cell death by post-ischemic reperfusion, most of which occurred

59

immediately after the re-establishment of the coronary flow, we observed

that, initially, the myocardial damage by H2O2 was small, but that it worsened

progressively with the duration of perfusion (Fig. 7). Importantly, with this

protocol the myocardial damage, during the first 15 min of perfusion, was

constantly more pronounced in PrPC-KO hearts than in the PrPC-expressing

hearts, the difference eventually becoming significant at the 14th and 15th min

of the perfusion period ((6.4 ± 1.6)% in WT, (12.5 ± 2.0)% in PrPC-KO, (6.5

± 2.2)% in PrPC-OE, at the end of the 15-min perfusion time-period) (see also

table 1, fourth line).

When perfusion with H2O2 was prolonged up to 30 min, the differences in

myocardial cell death between PrPC-KO and WT hearts disappeared, while

PrPC-OE hearts still resulted significantly more protected. This might be due to

PrPC downregulation during the first phase of the perfusion period (see Fig. 17

and the Conclusions and Perspectives section).

As expected, ROS are the main actors of the myocardial damage by H2O2,

given that loss of cell viability of PrPC-KO hearts after 15 min of perfusion,

was drastically reduced by the presence of MPG in the perfusion medium (Fig.

7, black line). In line with these results, we observed that the formation of Tm

dimers in hearts treated with H2O2 for 15 min was higher in the absence,

than in the presence, of PrPC (Fig. 8), indicating an increased degree of

myofibrillar protein oxidation.

Thus, in line with the results obtained with the IPC protocol, this data proves

once again the anti-ROS capacity of PrPC, and the protection that the protein

exerts on the myocardium against cell death and oxidative damage.

60

24 25 26 27 28 29 30

LDH

act

ivity

in th

e co

rona

ry e

fflue

nt (

%)

**

**

0

10

20

30

40

50

60

70

80

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

t (min)

**

= WT

= KO = OE

= KO + MPG

24 25 26 27 28 29 30

LDH

act

ivity

in th

e co

rona

ry e

fflue

nt (

%)

**

**

0

10

20

30

40

50

60

70

80

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

t (min)

**

= WT

= KO = OE

= KO + MPG

FIGURE 7. PrPC protects from H2O2-induced myocardial cell damage. Isolated hearts were

subjected to perfusion (15 or 30 min) with a buffer containing 1 mM H2O2. Myocardial cell loss

was evaluated as the cumulative LDH activity measured in the coronary effluent (for further

experimental details see Materials and Methods). After 14 and 15 min of perfusion, myocardial

cell viability is significantly increased in PrPC-KO hearts (red) with respect to WT (blue) and PrPC-

OE (green) hearts. When MPG (1 mM) is included in the perfusion medium, the incidence of cell

death in PrPC-KO hearts (black) is reduced to the level of PrPC-expressing hearts. Instead, at 26-

29 min of perfusion, myocardial cell death is significantly decreased in PrPC-OE hearts, while no

difference persists between WT and PrPC-KO hearts. Values are mean ± s.e.m., n = 12 (WT), 15

(KO and OE), 3 (KO + MPG) for the 15 min perfusion protocol, n = 4 for each genotype. * p <

0.05, Student’s t-test.

61

80 Oxidized tropomyosin(dimer)

40Tropomyosin(monomer)

WT KO OE CntrMW (kDa)

80 Oxidized tropomyosin(dimer)

40Tropomyosin(monomer)

WT KO OE CntrMW (kDa)

0

20

40

60

80

100

120

140

160

180

200

WT KO OE

* **

Nor

mal

ized

ban

d in

tens

ity (

%)

0

20

40

60

80

100

120

140

160

180

200

WT KO OE

** ****

Nor

mal

ized

ban

d in

tens

ity (

%)

FIGURE 8. Oxidation of tropomyosin after perfusion with H2O2 is increased in PrPC-KO

hearts. Hearts subjected to perfusion (15 min) with a buffer containing H2O2 (1 mM,) were

homogenized and assayed by Western blot for the presence of oxidized Tm as described in the

legend to Fig. 3. In the upper panel, the results of a Western blot analysis, representative of one

out of eight independent experiments, are reported. A sample of a heart from a WT mice perfused

(15 min) in the absence of H2O2 was also loaded into the gel as a negative control (Cntr). In the

lower panel, the results of the densitometric analysis are reported. Values (mean ± s.e.m., n = 8

for each PrPC genotype) are expressed as percentage of the WT samples. Tm oxidation after 15

min of perfusion with H2O2 is considerably increased in PrPC-KO with respect to WT and PrPC-OE

hearts. * p < 0.05 and ** p < 0.01 Student’s t-test.

62

Evaluation of the expression and/or activity of proteins involved the

oxidative response, in hearts with different PrPC levels

As detailed in the Introduction, since the discovery that PrPC had the

capability to bind copper at physiologic concentrations, a large number of

studies have been focussed on its possible (direct or indirect) role in the cell

defence against oxidative insults. Although the concept that PrPC possesses an

intrinsic SOD1-like activity is strongly disputed (and probably unrealistic), the

capability of the protein to modulate the expression and/or the activity of

different anti-oxidant systems is well documented (for detailed reviews see

Brown and Sassoon, 2002).

In light of these notions, we have tested the hypothesis that the different

capability to respond to the oxidative insult by hearts with different PrPC levels

might be due to variations in their endogenous anti-oxidant resources.

The enzymatic activity of CAT is diminished in PrPC–KO hearts

To unravel the molecular mechanisms at the basis of the PrPC-related

cardioprotection, we first compared the activity of CAT in hearts from WT,

PrPC-KO and PrPC-OE mice, not subjected to perfusion protocols. The

importance of CAT in the cellular detoxification from H2O2 is well recognized,

and reduced levels (by ~30%) of CAT activity have been already reported in

PrPC-KO hearts with respect to WT hearts (Klamt et al., 2001). We observed a

significant decrease (∼15%) of CAT activity in the hearts of PrPC-KO mice

compared to the PrPC-expressing counterparts. Conversely, no significant

difference was detected between WT and PrPC-OE hearts (Fig 9). This data

may contribute to justify the lower oxidative damage observed in PrPC-

expressing hearts subjected to perfusion with H2O2. Unfortunately, we were

not able to detect CAT in Western blot analyses, possibly due to the low

expression levels of the protein in the cardiac tissue (Ishikawa et al., 1986).

Of consequence, it remains an open question if PrPC modulates CAT activity,

or the myocardial content of the enzyme.

63

0

20

40

60

80

100

120

WT KO OE

* *

Nor

mal

ized

cata

lase

activ

ity(%

)

0

20

40

60

80

100

120

WT KO OE0

20

40

60

80

100

120

WT KO OE

** **

Nor

mal

ized

cata

lase

activ

ity(%

)

FIGURE 9. CAT activity is reduced in PrPC-KO hearts. Homogenates of non-perfused hearts

from mice with different PrPC levels were assayed for CAT activity by means of a standard

enzymatic assay. Briefly, CAT activity was measured by evaluating spectrophotometrically

(absorption at 240 nm) the rate of H2O2 consumption in the presence of a fixed amount (215 µg)

of total proteins from each homogenate. Measurements were calibrated by means of a standard

curve, created by use of known amounts of purified CAT, and then values were normalized to

those obtained with WT hearts (for further experimental details see Materials and Methods). Data

are expressed as mean ± s.e.m., n = 12 (WT), n = 8 (KO), n = 10 (OE). CAT activity is

significantly reduced in PrPC-KO hearts with respect to the PrPC- expressing counterparts.

* p < 0.05, Student’s t-test.

Hearts with different PrPC levels have no difference in superoxide dismutase

activities and expression

Another fundamental component of the ROS-scavenging systems of cells is

the enzyme SOD, catalyzing the dismutation of �O2- into molecular oxygen

and H2O2 (which is then removed by CAT and other enzymes). SOD exists in

three isoforms, the major two being the cytosolic Cu/Zn-dependent SOD

(SOD1), and the mitochondrial Mn-dependent SOD (SOD2). Given that a

reduced SOD activity has been repeatedly reported in PrPC-KO or prion-

infected brains (Klamt et al., 2001; Wong et al., 2001; but see also Hutter et

al., 2003), and that we observed different amounts of

production/accumulation of �O2- in perfused hearts, depending on the PrPC

expression levels, we decided to evaluate SOD activities and expression in

hearts with different PrPC amounts. SOD1 (Fig. 10) and SOD2 (Fig. 11)

64

activities were evaluated separately, by measuring SOD activity in the

cytosolic and mitochondrial fractions of non-perfused heart homogenates,

respectively. In both cases, no significant difference was observed in mice

with different PrPC genotype. Accordingly, no difference in both SOD1 and

SOD2 expression was detected by Western blot analyses on heart samples

from WT, PrPC-KO and PrPC-OE mice (data not shown).

0

20

40

60

80

100

120

WT KO OE

Nor

mal

ized

cyto

solic

SO

D a

ctiv

ity(%

)

0

20

40

60

80

100

120

WT KO OEWT KO OE

Nor

mal

ized

cyto

solic

SO

D a

ctiv

ity(%

)

FIGURE 10. SOD1 activity in the heart is unaffected by PrPC expression levels. The

cytosolic fraction of heart homogenates from mice with different PrPC levels was assayed for SOD

activity by quantifying spectrophotometrically the inhibition of xanthine/xanthine oxidase–induced

cytochrome c reduction (for further experimental details see Materials and Methods). Data (mean

± s.e.m., n = 4 for each mouse strain) were normalized to mean value obtained with WT hearts.

No significant difference exists between hearts with different PrPC amounts.

65

0

20

40

60

80

100

120

140

Nor

mal

ized

mito

chon

dria

lSO

D a

ctiv

ity(%

)

WT KO OE0

20

40

60

80

100

120

140

Nor

mal

ized

mito

chon

dria

lSO

D a

ctiv

ity(%

)

WT KO OEWT KO OE

FIGURE 11. SOD2 activity in the heart is unaffected by PrPC expression levels. The

mitochondrial fraction of heart homogenates from mice with different PrPC levels was assayed for

SOD activity. Data (mean ± s.e.m., n = 4 for each mouse strain) were normalized to mean value

obtained with WT hearts. No significant difference exists between hearts with different PrPC

amounts. Other experimental details are as in the legend to Fig. 10.

p66Shc expression is increased in PrPC–KO hearts

A very recent work has demonstrated that the protein p66Shc significantly

contributes to mitochondrial ROS formation and myocardial damage caused in

the heart by I/R challenge (Carpi et al., 2009). p66Shc is a splice variant of

p52Shc/p46Shc, two cytoplasmic adaptor proteins involved in the propagation of

intracellular signals from activated tyrosine kinases to Ras. Rather than

functioning as an adaptor protein, p66Shc is mainly involved in the intracellular

pathways that regulate ROS metabolism and apoptosis (Migliaccio et al.,

1997). A fraction of p66Shc localizes within the mitochondrial inter-membrane

space where it oxidizes reduced cytochrome c, contributing to ROS formation

and apoptosis (Cosentino et al., 2008; for an illustrated mechanism of p66Shc

functions see Conclusions and Perspectives, Fig. 1). Accordingly, mice ablated

for p66Shc expression display strongly reduced ROS formation and damage,

increased resistance to apoptotic stimuli, increased life-span (Migliaccio et al.,

1999), and, as anticipated, reduced myocardial damage upon I/R injury

(Carpi et al., 2009).

66

We therefore asked whether PrPC might influence the expression of p66Shc in

the heart. Surprisingly, by subjecting total heart homogenates to immunoblot

analyses with an antibody directed against all Shc isoforms, we observed a

significant 2.5-fold increase in the levels of p66Shc in PrPC-KO hearts with

respect to the PrPC-expressing counterparts (Fig. 12). Conversely, no

significant difference was observed, in the different mouse strains, in the

cardiac levels of p52Shc/p46Shc (data not shown).

The increased levels of p66Shc in PrPC-KO hearts were also reflected by a

larger mitochondrial content of the protein. Indeed, when probing the

mitochondrial fraction of heart homogenates for the presence of p66Shc, a

significant (∼60%) increase of the protein was observed in the absence of PrPC

with respect to PrPC-expressing hearts (Fig. 13).

Taken together, these results nicely fit with the data shown previously, and

may open new perspectives in the study of PrPC patho-physiology.

67

66

OEKOWTMW (kDa)

66

OEOEKOKOWTWTMW (kDa)

0

50

100

150

200

250

300

WT KO OE

Nor

mal

ized

ban

d in

tens

ity (

%)

** *

0

50

100

150

200

250

300

WT KO OEWT KO OE

Nor

mal

ized

ban

d in

tens

ity (

%)

**** **

FIGURE 12. p66Shc expression is increased in PrPC-KO hearts. Non-perfused hearts from

mice with different PrPC levels, were homogenized, proteins (50 µg) were resolved onto 8% SDS-

PAGE under reducing conditions, and then electro-blotted onto nitrocellulose membranes. p66Shc

expression was evaluated by Western blotting with an anti-Shc Pab, as the abundancy of the 66

kDa immuno-reactive band. In the upper panel, the results of a Western blot analysis,

representative of at least nine independent experiments, are reported. In the lower panel, the

results of the densitometric analysis are reported. Values (mean ± s.e.m., n = 9 (WT), 10 (KO)

and 11 (OE)) are expressed as percentage of the WT samples. p66Shc expression is largely

increased in PrPC-KO hearts compared to WT and PrPC-OE hearts. * p < 0.05 and ** p < 0.01,

Student’s t-test.

68

OEKOWT

66

MW (kDa)OEOEKOKOWTWT

66

MW (kDa)

0

20

40

60

80

100

120

140

160

180

200 **

Nor

mal

ized

ban

d in

tens

ity (

%)

WT KO OE

*

0

20

40

60

80

100

120

140

160

180

200 **

0

20

40

60

80

100

120

140

160

180

200 **

Nor

mal

ized

ban

d in

tens

ity (

%)

WT KO OEWT KO OE

**

FIGURE 13. Mitochondrial p66Shc levels are increased in PrPC-KO hearts. The

mitochondrial fractions from homogenates of non-perfused hearts were probed for the presence

of p66Shc. In the upper panel, the results of a Western blot analysis, representative of at least five

independent experiments, are reported. In the lower panel, the results of the densitometric

analysis are reported. Values (mean ± s.e.m., n = 6 (WT), 5 (KO and OE)) are expressed as

percentage of the WT samples. p66Shc presence is largely increased in PrPC-KO hearts compared

to WT and PrPC-OE hearts. * p < 0.05 and ** p < 0.01, Student’s t-test. Other experimental

details are as in the legend to Fig. 12, except that 20 µg of proteins were loaded into the gels.

69

The fate of PrPC during and after the ischemic and oxidative

challenges

It is well known that during I/R the levels of myocardial proteins may vary.

This up- or down-regulation may be due to inflammatory injury, with

endothelial dysfunction and neutrophil accumulation into the myocardium, as

well as generation of free radicals and cytokines, that, in turn, can activate

proteolysis pathways. Otherwise, the activation of the complement system,

which leads to inflammatory injury, can trigger cellular activation and protein

synthesis (Maroko et al., 1978; Lefer et al., 1994). In this scenario, the

expression of proteins involved in protective pathways may be crucial for cell

survival. Given that PrPC plays a role in the heart cell defences, we decided to

measure its expression during and after the used perfusion protocols.

PrPC levels are decreased after I/R, but not after ischemia alone, in WT and

PrPC-OE hearts

The finding that PrPC-OE hearts, but not WT hearts, were less damaged by I/R

injury than the PrPC-KO counterpart led us to hypothesize that the physiologic

expression levels of PrPC were not sufficient to counteract the insult. This also

prompted us to check the actual levels of PrPC at the end of the entire I/R

protocol in PrPC-expressing hearts. Interestingly, we found that after the I/R

challenge the expression of PrPC, assessed by Western blotting analyses, was

decreased by ∼20% in WT and ∼35% in PrPC-OE hearts with respect to the

corresponding non-perfused hearts (Fig. 14).

70

MW (kDa)

D

MU

35.5

29

WTCNTR

WTIPC+I/R

OECNTR

OEIPC+I/R

1 2 3 4

MW (kDa)

D

MU

35.5

29

WTCNTR

WTIPC+I/R

OECNTR

OEIPC+I/R

1 2 3 4

D

MU

35.5

29

WTCNTR

WTIPC+I/R

OECNTR

OEIPC+I/R

1 2 3 4

0

50

100

150

200

250

300

350

WT CNTR OE CNTRWT I/R OE I/R

Nor

mal

ized

band

inte

nsity

(%)

**

*

0

50

100

150

200

250

300

350

WT CNTR OE CNTRWT I/R OE I/R

Nor

mal

ized

band

inte

nsity

(%)

****

**

FIGURE 14. The expression of PrPC is decreased after the I/R challenge. Hearts from WT

and PrPC-OE mice, subjected, or not (CNTR), to I/R (40 min/15 min), were probed for the

expression of PrPC by Western blot analysis with anti-PrP Mab 8H4. In the upper panel, the results

of a Western blot analysis, representative of at least three independent experiments, are

reported. Arrows on the right indicate the different PrPC glycoforms, while on the left molecular

mass standards are reported. In the lower panel, the results of the densitometric analysis are

reported. Data (mean ± s.e.m., n = 3 for WT and PrPC-OE CNTR hearts, n = 6 for WT and n = 7

for PrPC-OE I/R-treated hearts) are expressed as percentage of the mean value of WT CNTR

hearts. PrPC expression after the I/R protocol is significantly reduced in both WT and PrPC-OE

hearts (lanes 2 and 4 of the Western Blot, dark bars of the diagram) with respect to the

corresponding untreated hearts (lanes 1 and 3 of the Western Blot, light bars of the diagram).

* p < 0.05 and ** p < 0.01, Student’s t-test. Other experimental details are as in the legend to

Fig. 1.

It was therefore reasonable to speculate that the reduction in myocardial PrPC

content, observed upon I/R in WT hearts, although not dramatic, was

71

nevertheless sufficient to cause a significant loss of the PrPC-dependent

protective functions. However, when probing the PrPC levels at the end of the

ischemic time-period (40 min), no difference was observed in both WT and

PrPC-OE hearts with respect to the untreated counterparts (Fig. 15). Thus,

when reperfusion starts and the destructive burst of ROS occurs, both WT and

PrPC-OE hearts still contain their original PrPC reservoir.

D

MU

35.5

29

MW (kDa)

WTCNTR

WTI 40’

OECNTR

OEI 40’

1 2 3 4

D

MU

35.5

29

MW (kDa)

WTCNTR

WTI 40’

OECNTR

OEI 40’

1 2 3 4

D

MU

35.5

29

MW (kDa)

WTCNTR

WTI 40’

OECNTR

OEI 40’

D

MU

35.5

29

MW (kDa)

WTCNTR

WTI 40’

OECNTR

OEI 40’

1 2 3 4

FIGURE 15. The expression PrPC is unchanged after 40 min of ischemia. Hearts from WT

and PrPC-OE mice, subjected, or not (CNTR), to a 40-min ischemia (I 40’), were assayed for the

expression of PrPC by Western blot analysis with anti-PrP Mab 8H4. The reported Western blot is

representative of three independent experiments, which yielded comparable results. PrPC

expression after the ischemic time-period is not appreciably changed in both WT and PrPC-OE

hearts (lanes 2 and 4) with respect to the corresponding untreated counterparts (lanes 1 and 3).

Other experimental details are as in the legends to Fig. 1 and 14.

PrPC levels are preserved when I/R is preceded by IPC

In view of the previous results we could conclude that the reduction in

myocardial PrPC content, observed after the I/R protocol, occurs during the

reperfusion step. Given that IPC protects hearts from the ROS challenge

caused by post-ischemic reperfusion, we next asked if IPC could also protect

the myocardium from the I/R-induced loss of PrPC. We found that indeed this

was the case, as the levels of the protein were preserved when both WT and

PrPC-OE hearts were subjected to I/R after the application of the IPC stimulus

(Fig. 16).

72

MW (kDa)

D

MU

35.5

29

WTCNTR

WTIPC+I/R

OECNTR

OEIPC+I/R

1 2 3 4

MW (kDa)

D

MU

35.5

29

WTCNTR

WTIPC+I/R

OECNTR

OEIPC+I/R

1 2 3 4

D

MU

35.5

29

WTCNTR

WTIPC+I/R

OECNTR

OEIPC+I/R

1 2 3 4

FIGURE 16. PrPC levels are preserved when I/R is preceded by IPC. Hearts from WT and

PrPC-OE mice, subjected, or not (CNTR), to I/R (40 min/15 min) preceded by IPC (3 cycles of 5

min I/5 min R) (IPC+I/R), were probed for the expression of PrPC by Western blot analysis with

anti-PrP Mab 8H4. The reported Western blot is representative of three independent experiments,

which yielded comparable results. PrPC expression after the IPC+I/R is not appreciably changed in

both WT and PrPC-OE hearts (lanes 2 and 4) with respect to the corresponding untreated

counterparts (lanes 1 and 3). Other experimental details are as in the legends to Fig. 1 and 14.

PrPC levels are largely reduced after perfusion with H2O2

To investigate further if oxidative stress is a major player in the loss of

myocardial PrPC during post-ischemic reperfusion, we analysed PrPC

expression in hearts subjected to perfusion (15 min) with H2O2 (1 mM). We

found that, in both WT and PrPC-OE hearts, PrPC levels were dramatically

reduced (by ∼50%) at the end of the perfusion period, compared to non-

perfused hearts (Fig. 17).

73

D

MU

35.5

29

MW (kDa)

WTCNTR

WTH2O2

OECNTR

OEH2O2

1 2 3 4

D

MU

35.5

29

MW (kDa)

WTCNTR

WTH2O2

OECNTR

OEH2O2

D

MU

35.5

29

MW (kDa)

WTCNTR

WTH2O2

OECNTR

OEH2O2

1 2 3 41 2 3 4

0

50

100

150

200

250

300

350

WT H2O2 OE H2O2WT CNTR OE CNTR

Nor

mal

ized

band

inte

nsity

(%)

***

**

0

50

100

150

200

250

300

350

WT H2O2 OE H2O2WT CNTR OE CNTR

Nor

mal

ized

band

inte

nsity

(%)

******

****

FIGURE 17. Perfusion with H2O2 produces a drastic loss of myocardial PrPC. Hearts from

WT and PrPC-OE mice, subjected, or not (CNTR), to perfusion (15 min) with H2O2 (1 mM), were

probed for the expression of PrPC by Western blot analysis with anti-PrP Mab 8H4. In the upper

panel, the results of a Western blot analysis, representative of at least three independent

experiments, are reported. In the lower panel, the results of the densitometric analysis are

reported. Data (mean ± s.e.m., n = 3 for WT and PrPC-OE CNTR hearts, n = 7 for WT and n = 9

for PrPC-OE H2O2-treated hearts) are expressed as percentage of the mean value of WT CNTR

hearts. PrPC expression after perfusion with H2O2 is largely reduced in both WT and PrPC-OE

hearts (lanes 2 and 4 of the Western Blot, dark bars of the diagram) with respect to the

corresponding untreated hearts (lanes 1 and 3 of the Western Blot, light bars of the diagram).

** p < 0.01 and *** p < 0.001 Student’s t-test. Other experimental details are as in the legend

to Fig. 1 and 14.

Which is the fate of myocardial PrPC during post-ischemic reperfusion, or

perfusion with H2O2?

The expression of the prion protein is significantly reduced after the injury by

post-ischemic reperfusion and perfusion with H2O2. Both protocols largely

include the production of ROS as a prime mediator of damage. Oxidative

74

stress in the heart is known to trigger the activation of matrix metallo-

proteases (MMPs), responsible for the proteolysis of several matrix proteins

(for a review, see Schulz 2007), and PrPC might be a target of MMPs (Parkin

et al., 2004). In addition, it has been reported that mature PrPC may

physiologically undergo two distinct endo-proteolytic cleavage events,

generating two membrane-bound C-terminal fragments, named C1 and C2, of

approx. 18 and 20 kDa, respectively, which bear both the glycosylation sites

(Chen et al., 1995; Vincent et al., 2001; Mangè et al., 2004). Interstingly, the

cleavage event generating the C2 fragment, occurring next to the octapeptide

repeat region, appears to be mediated by ROS (McMahon et al., 2001), and

seems to play a role in the cellular response to oxidative stress (Watt et al.,

2005). In light of these notions, we decided to verify if the reduction in

myocardial PrPC content was a consequence of increased proteolytic cleavage.

The C1 and C2 fragments are not easily distinguishable from one another, or

from the full-lenght protein, in Western blot analysis of crude homogenates,

as they carry the high molecular weight oligo-saccharidic chains. However,

they become easily appreciable after removal of glycans with the de-

glycosylating enzyme PNGase-F (Massimino et al., 2005). Therefore, we

analysed the PrPC expression pattern in heart homogenates treated with

PNGase-F. Our results indicated that, after either I/R (Fig 18, upper panel) or

perfusion with H2O2 (Fig 18, lower panel), the amount of C1 and C2 were not

significantly increased in both WT and PrPC-OE hearts compared to the

respective non-perfused hearts. From these data, we can conclude that I/R-

or H2O2-induced oxidative stress did not increase the endo-proteolytic

cleavage of PrPC.

Another possible explanation for the myocardial loss of PrPC entails that the

protein is released from the plasma membrane of cardiomyocytes into the

perfusion buffer. However, no detectable PrPC was never found in methanol-

precipitated protein samples from the coronary effluent of perfused hearts

(data not shown). In conclusion, PrPC is probably degraded, either in the

extracellular matrix, or after internalization into cells, but this issue demands

further elucidation.

75

35.5

+ +- -WT I/R

PNGaseF

WT CNTR

29

20

D

M

U

+ +- -OE I/ROE CNTR

MW (kDa)

1 3 52 4 6 7 8

18

C 2

C 1

35.5

+ +- -WT I/R

PNGaseF

WT CNTR

29

20

D

M

U

+ +- -OE I/ROE CNTR

MW (kDa)

1 3 52 4 6 7 8

18

35.5

+ +- -WT I/R

PNGaseF

WT CNTR

29

20

D

M

U

+ +- -OE I/ROE CNTR

MW (kDa)

1 3 52 4 6 7 8

18

C 2

C 1

35.5

+ +- -WT H2O2

PNGaseF

WT CNTR

29

20

D

M

U

+ +- -OE H2O2OE CNTR

MW (kDa)

1 3 52 4 6 7 8

18

C 2

C 1

35.5

+ +- -WT H2O2

PNGaseF

WT CNTR

29

20

D

M

U

+ +- -OE H2O2OE CNTR

MW (kDa)

1 3 52 4 6 7 8

18

35.5

+ +- -WT H2O2

PNGaseF

WT CNTR

29

20

D

M

U

+ +- -OE H2O2OE CNTR

MW (kDa)

1 3 52 4 6 7 8

18

C 2

C 1

FIGURE 18. The endo-proteolytic pattern of PrPC is unaffected by I/R and perfusion

with H2O2. Homogenates of WT and PrPC-OE hearts, subjected, or not (CNTR), to I/R (40 min/15

min) (upper panel), or perfusion (15 min) with H2O2 (1 mM) (lower panel), were treated (+), or

not (-), with the sugar-removing enzyme PNGase-F, and then probed for PrPC expression by

Western blot analysis with anti-PrP Mab 8H4. In both panels, the results of a Western blot,

representative of three independent experiments, are reported. After either I/R or H2O2-

treatment, a decrease in full-length PrPC levels is observed, whereas no significant difference is

evident in the amounts of the C1 and C2 endo-proteolytic PrPC products, at around 18 and 20

kDa, respectively, in both WT and PrPC-OE hearts. Other experimental details are as in the legend

to Fig. 1 and 14. To note that the differences between the de-glycosylated full-length PrPC band in

perfused and non-perfused PrPC-OE hearts might not be readily appreciable because blots have

been over-exposed in the attempt to visualize better the cleavage products.

76

77

CONCLUSIONS AND PERSPECTIVES

A wealth of evidence implicates PrPC in cell protective mechanisms. Most of

these data are, however, merely circumstantial, and – with rare exceptions –

do not support any physiologic significance for the putative PrPC function. For

example, cells deprived of PrPC were shown to be more susceptible to

oxidative and apoptotic injury, and decreased anti-oxidant defences and

increased oxidative damage have been repeatedly reported in the brain and

other tissues of PrPC-KO mice. These mice, however, do not succumb to

oxidative overload, nor display gross signs of physiologic disturbances, and

apparently live (under normal conditions) a regular and joyful life. The most

sensible justifications for these findings are that either PrPC is in fact

dispensable for life, or it becomes necessary only in yet unidentified

situations.

Quite surprisingly, the biologic functions of PrPC have never been verified in

isolated organs. These systems maintain the cell native environment, but they

are also more amenable than whole animals to experimental manipulations

aimed at elucidating the molecular and cellular mechanisms by which proteins

exert their functions. Recent findings have highlighted a role for PrPC in

protecting from ischemic brain injury. Given that oxidative stress is a prime

mediator of hypoxia-induced cell damage in both heart and brain tissues

(Dröge, 2002), this suggested an intriguing parallelism between heart and

brain with respect to PrPC protective activity. These are the basic reasons for

our choice of using perfused hearts as intact organ paradigms where to verify

the putative antagonism of PrPC over ischemic injury by ROS.

The use of different perfusion protocols gave us the possibility to differentiate

the insults imposed to cardiomyocytes. We firstly used an I/R protocol, which

is the most renowned, and perhaps physiologically relevant, experimental

paradigm of isolated hearts. Contrary to the most simplistic expectations, and

despite the large number of experiments performed, a perfusion protocol

consisting of 40 min ischemia followed by 15 min reperfusion failed to

underscore any significant difference in myocardial cell death between WT and

PrPC-KO mice. Instead, we observed a significant reduction of myocardial

78

damage in PrPC-OE hearts. This data nicely correlates with the reduced

superoxide accumulation and myofibrillar protein oxidation observed in PrPC-

OE hearts. It is also important to underline that overexpression of PrPC,

naturally occurring at different days after focal cerebral ischemia and hypoxia,

or mediated by adenoviral delivery of a PrP transgene, was demonstrated to

protect from ischemic brain injury (Weise et al., 2004; Shyu et al., 2005).

However, it is not easy to rationalize the fact that, in our I/R protocol, the

absence of PrPC does not result in any appreciable effect. As already

mentioned, post-ischemic reperfusion injury is the result of a complex set of

functional and metabolic modifications (Allen et al., 1990; Allen and Orchard,

1987). Thus we can speculate that the physiologic PrPC levels are not

sufficient to counteract the burst of ROS, which occurs at the onset of

reperfusion, in hearts that have been previously weakened by a prolonged

ischemic event.

Accordingly, when a strong oxidative challenge, i.e. perfusion (15 min) with 1

mM H2O2, is given without ischemia, a PrPC-KO-related phenotype becomes

clearly evident. It has to be pointed out that such a high H2O2 dose was

administrated in order to take the oxidative processes to their maximal

extent. This should override the endogenous antioxidant defences of the

cardiac myocytes (Canton et al., 2004), and elicit the sole contribute of PrPC

against the oxidative insult. In this case, a significant increase of myocardial

cell loss and protein oxidation is observed in the absence of PrPC compared to

PrPC-expressing hearts. In addition, the lack of differences between WT and

PrPC-OE hearts after 15 min of perfusion argues in favour of the hypothesis

that the natural levels of the protein are sufficient to protect the myocardium

from a simple, though strong, oxidative insult. Intriguingly, the PrPC

protective effects against the H2O2 insult is further corroborated by the results

obtained when prolonging the perfusion up to 30 min. Indeed, it is quite fair

to speculate that the strong downregulation of PrPC, observed after 15 min of

perfusion with H2O2 (see Fig. 17), abrogates the beneficial effects of the

protein in WT hearts. Of consequence, a more long-lasting protocol abolishes

the differences between WT and PrPC-KO hearts, while highlighting a

significantly reduced cell loss in PrPC-OE hearts, in which PrPC is still

appreciably detectable at the 15th min of perfusion.

PrPC’s antioxidant properties were further supported by the results obtained

with ischemic preconditioning. Though the protective mechanisms at the basis

79

of IPC have not been completely unravelled, the evidence of an involvement

of small amount of ROS produced during the intermittent short reperfusion

time periods has been widely documented (Pain et al., 2000; Vanden-Hoek et

al., 1998; see the Introduction for further details). In our experimental

paradigm, the extent of IPC-induced cardioprotection inversely correlates with

the levels of PrPC, being most remarkable in PrPC-KO hearts. It is thus

reasonable to conclude that increased ROS production during IPC in hearts

lacking PrPC triggers more powerfully the adaptive response that protects the

myocardium from the subsequent I/R insult. In line with this, decreased ROS

accumulation was observed in the PrPC-KO myocardium at the end of IPC,

while the addition of an anti-oxidant in the perfusion medium completely

abrogates IPC’s beneficial effect in PrPC-KO hearts, further supporting the

anti-oxidant potential of the protein.

The protection given by PrPC over injury by ROS has been already suggested

by several works. Evidence of a modulation of the cellular antioxidant

defences by PrPC, not only in the CNS but also in other tissues (Brown et al.,

1997c; White et al., 1999; Klamt et al., 2001), and of a putative direct

involvement in the scavenging of ROS have been reported (Brown et al.,

1997c). In particular, CAT activity was found significantly reduced in hearts

form PrPC-KO C57BL/6J mice with respect to the WT counterpart (Klamt et al.,

2001). CAT decomposes H2O2 to H2O and oxygen, and is important to prevent

the Fenton and Haber-Weiss reactions, which generate the highly reactive,

and tissue-damaging, hydroxyl radical. Our results support what previously

reported, in that a significant reduction of CAT activity was observed in PrPC-

KO compared to PrPC-expressing hearts. CAT is a very efficient enzyme,

bearing one of the highest turnover numbers of all enzymes. Of consequence

it can be physiologically relevant even when expressed at very low levels. We

were not able to detect myocardial CAT by Western blot analyses, indicating

that it is indeed scarcely expressed in the heart. Of consequence, we cannot

conclude if the observed reduction of CAT activity in the absence of PrPC is

ascribable to a decreased expression of the enzyme. The effects of PrPC on

SOD activity are strongly disputed. While some research groups have reported

a decreased SOD activity in PrPC-less paradigms (Brown et al., 1997c;

Milhavet et al., 2002; Wong et al., 2001; Klamt et al. 2001), or even an

intrinsic SOD activity of the recombinant PrP (Brown et al., 1999), others

have categorically rejected such a possibility (Waggoner et al., 2000; Hutter

80

et al., 2003). Our results support this second possibility, given that no

significant difference was seen in the activity of both cytosolic SOD1 and

mitochondrial SOD2 in hearts with different PrPC genotypes. Importantly,

following the work by Hutter and colleagues, we assessed the SOD activity by

means of xanthine/xanthine oxidase-based assay (Okado-Matsumoto and

Fridovich, 2001). This assay may be more sensitive and less error-prone than

the NBT-method mainly used in other studies. Indeed, the NBT-based assay

was shown to potentially interfere with xanthine oxidase (Ukeda et al., 1997)

and resulted in poor reproducibility (Hutter et al., 2003).

At present, however, we cannot conclude that the protection from oxidative

injury exerted by PrPC has to be ascribed to the unique modulation of CAT

activity, being this enzyme not particularly abundant in the heart (see also

Ishikawa et al., 1986). Nevertheless, it is possible that, under conditions of

extreme oxidative stress, which in most cases involve GSH depletion, CAT

becomes important in providing cytoprotection (Jones et al., 1978; Kang et

al., 1996). Previous studies performed on cerebellar granule neurons cultures

showed a higher susceptibility of PrPC-KO cells to H2O2 administration, not

ascribable to altered CAT activity, but rather linked to a reduced activity of

GSH reductase (White et al., 1999). To assess whether, or not, CAT activity

could be crucial in the detoxification from H2O2 in our paradigm, further

experiments, also addressing the activity of the GSH-linked systems in hearts

with different PrPC levels, are demanded.

Another novel and important insight came from our experiments on p66Shc

expression. As previously mentioned, p66Shc is a splice variant of

p52Shc/p46Shc, two cytoplasmic adaptor proteins involved in Ras signalling.

p66Shc, a fraction of which localizes within the mitochondrial inter-membrane

space where it oxidizes reduced cytochrome c, regulates ROS metabolism

(Migliaccio et al., 1997) (for an illustrated mechanism of p66Shc see Fig. 1).

Following the mitochondrial re-localization of p66Shc, part of the respiratory

chain electron flow is therefore diverged to the production of remarkable

amounts of H2O2, corresponding to approximately one third of the total

intracellular H2O2 pool (Giorgio et al., 2005). The biological significance of this

effect is underscored by the fact that mouse-derived p66Shc−/− cells

accumulate significantly less markers of oxidative stress (Giorgio et al.,

2005). Thereby, this molecule plays an important role in mediating oxidative

damage and apoptosis. For this reason, for the first time in the field of prion

81

biology, we measured the expression of p66Shc in hearts with different PrPC

genotype. Our results clearly demonstrated a significantly higher expression

of p66Shc in heart homogenates of PrPC-KO hearts with respect to the WT and

PrPC-OE counterparts, also reflected by an increased mitochondrial content of

the protein. This further supports the possible involvement of PrPC in

preventing oxidative stress and mitochondria-mediated apoptosis. Moreover a

role for p66Shc in apoptotic and necrotic myofibers death due to ischemia-

reperfusion injury has been repeatedly demonstrated (Migliaccio et al., 1999;

Napoli et al., 2003; Zaccagnini et al., 2004), and – most importantly – it has

been recently reported that p66Shc plays a major role in ROS mediated

myocardial damage consequent to post-ischemic reperfusion (Carpi et al.,

2009). It would be of importance to verify the presence of p66Shc after IPC in

hearts with different PrPC levels, being already demonstrated that a down-

regulation of this protein occurs in preconditioning-mediated protection of

human neuroblastoma cells (Andoh et al., 2001). Another important

correlation between p66Shc and PrPC could be ascribed to the finding that

dilated cardiomyopathy, which is associated with increased levels of oxidative

stress–mediated cell death, was observed in TSE-affected individuals, where

the physiological content of PrPC is diminished due to its conversion in the

pathological isoform (Ashwath et al., 2005). Indeed, the same pathology was

correlated to an increased p66Shc expression (Cesselli et al., 2001). It would

be very interesting to assess if, in the absence of PrPC, increased levels of

p66Shc also pertains to other cell types, in particular neurons, in which PrPC

should mainly perform its physiologic functions. Importantly, the expression

levels and mitochondrial translocation of p66Shc should be evaluated in TSE-

affected brains, in order to understand if a correlation exists between PrPC

depletion and p66Shc-mediated oxidative stress in prion diseases.

82

FIGURE 1. p66Shc is part of the oxidative stress-induced apoptosis mechanism. Free

radicals activate protein kinase C-β isoform to induce Ser36 phosphorylation of the p66Shc, allowing

transfer of the protein from the cytosol to the mitochondrion. In the mitochondrion, p66Shc binds

to a complex which includes members of the TIM-TOM import system. Pro-apoptotic stimuli

destabilize the p66Shc-mtHsp70 complex and lead to the release of p66Shc in its monomeric form.

Once activated, p66Shc oxidizes cytochrome C and catalyzes the reduction of O2 to H2O2. This

latter event triggers the opening of the mitochondrial permeability transition pore, with

subsequent increase of mitochondrial membrane permeability to ions, solutes and water, swelling

and disruption of the organelle, and consequent release of pro-apoptotic factors into the cytosol

(Cosentino et al., 2008).

Thus, we have provided strong evidence that PrPC plays a role in the

myocardial defence against both I/R-mediated, and non-ischemic, oxidative

damage, possibly by modulating the expression of pro- and anti-oxidant

systems. This function may be physiologically relevant, given that it is

reflected by increased cell survival in PrPC-expressing intact hearts. But how

can a protein that is located extracellularly impinge on processes, such as

ROS production and scavenging and protein expression, which are mainly

intracellular? In this context, it is important to remember that several lines of

evidence support the possibility that PrPC takes part to multi-component

83

signal transduction complexes at the cell surface. Accordingly, several

putative functional partners of PrPC, and different signalling pathways in which

the protein may come to play, have been proposed (for exhaustive reviews

see Aguzzi et al., 2008; Linden et al., 2008; Sorgato et al., 2009). The I/R

damage involves a large number of dysfunction correlated to the alteration of

different signalling pathways. Ionic imbalance, e.g. calcium overload, is one of

the major causes of death in the I/R model, implicated in a large variety of

processes, ranging from activation of calpains, to mitochondria mediated

apoptosis (Garcia-Dorado et al., 2009). The expression of proteins involved in

Ca2+ handling seems important in protecting from this damage. SERCA over-

expression, for example, has been shown to result in several beneficial effects

in animal models, such as a diminished occurrence of arrhythmias during

ischemia-induced Ca2+ overload (del Monte et al., 2004). It will be therefore

important to measure the expression of such proteins in hearts with different

PrPC genotypes, most importantly in light of the evidences that correlate Ca2+

homeostasis to PrPC (Sorgato and Bertoli, 2009), and the finding that PrPC

positively modulates SERCA expression (Brini et al., 2005). Other signalling

pathways that could link PrPC to cardioprotection are those involving Akt or

the MAP kinase Erk1/2. Indeed, it has been shown that, in a model of

ischemic brain injury, the absence of PrPC results in reduced and increased

activation of Akt and Erk1/2, respectively, both events resulting in enhanced

post-ischemic caspase-3 activity, and exacerbation of brain damage (Spudich

et al., 2005; Weise et al., 2006). On the contrary, overexpression of PrPC

decreases early post-ischemic Erk1/2 activation and protects against focal

cerebral ischemia (Weise et al., 2008). Importantly, it has been documented

that Akt protects hearts from I/R injury (Miao et al., 2000; Fujio et al., 2000).

In this same context, it is of value to underscore the recent finding that the

activation of the MEK-Erk1/2 pathway in cardiac myocytes promotes the

phosphorylation of p66Shc at Ser36, required for the mitochondrial re-

localization of the protein (Obreztchikova et al., 2006). This would be

consistent with our data, whereby the absence of PrPC might increase the

mitochondrial p66Shc pool, through increased activation of Erk1/2. Further

analyses need to be performed to validate these hypotheses.

Another intriguing finding of this work is the substantial reduction of

myocardial PrPC observed after I/R and perfusion with H2O2 in PrPC-expressing

hearts. It is well known that during ischemia and reperfusion the levels of

84

myocardial proteins may vary, as a consequence of altered protein synthesis

and/or degradation. It is also recognized that the generation of free radicals

can activate proteolytic pathways. At the end of the I/R protocol we observed

a significant decrease of PrPC levels in both WT and PrPC-OE mice. We asked if

this reduction might account for the lack of difference between WT and PrPC-

KO hearts in I/R protocols. These seems not to be the case, given that after

the 40 min ischemia time period, no reduction of myocardial PrPC content was

seen with respect to non-perfused hearts. Thus, at the beginning of

reperfusion, WT hearts still contained the natural amounts of PrPC. This result,

together with the loss of PrPC upon perfusion with H2O2, suggests that this

event might be due to ROS-dependent degradation of the protein. This

possibility is also supported by the finding that PrPC levels are preserved when

I/R is preceded by the anti-oxidant activity of IPC. PrPC may undergo two

distinct endo-proteolytic cleavage events, one of which is mediated by ROS

and may play a role in the cellular response to oxidative stress (Watt et al.,

2005). We did not observe, however, any significant increase in the

myocardial content of the two C-terminal PrPC fragments, C1 and C2,

following I/R or H2O2 administration. It is therefore likely, although not

conclusive, that PrPC is subjected to unspecific degradation, and/or reduced

synthesis, due to myocardial metabolic impairments under the imposed

insults. Further studies are requested to understand if these events may be

physiologically relevant.

In conclusion, the novel approach, based on the use of an extra-neural tissue,

i.e. intact hearts, gave us the possibility to verify, for the first time, the

putative antioxidant and cell-protective function of PrPC in isolated organs,

suggesting that this activity of PrPC might be physiologically important under

specific stress conditions. This work allows us to confirm previous findings,

achieved by different models, and opens new interesting insights in the field

of PrPC physiology.

85

BIBLIOGRAPHY

• Aebi H. Catalase in vitro. Methods Enzymol. 1984;105:121-6.

• Aguzzi A, Baumann F, Bremer J. The prion's elusive reason for being. Annu Rev Neurosci. 2008;31:439-77.

• Aguzzi A, Calella AM. Prions: protein aggregation and infectious diseases. Physiol Rev. 2009 Oct;89(4):1105-52.

• Allen DG, Orchard CH. Myocardial contractile function during ischemia and hypoxia. Circ Res. 1987 Feb;60(2):153-68.

• Allen KL, Busza AL, Proctor E, Williams SR, Van Bruggen N, Gadian DG, Crockard HA. Restoration of energy metabolism and resolution of oedema following profound ischaemia. Acta Neurochir Suppl (Wien). 1990;51:171-3.

• Alper T, Does the agent of scrapie replicate without nucleic acid? Nature. 1967 May 20;214(5090):764-6.

• Andoh T, Lee SY, Chiueh CC. Preconditioning regulation of bcl-2 and p66shc by human NOS1 enhances tolerance to oxidative stress. Faseb J. 2000;14:2144–2146.

• Andréoletti O, Simon S, Lacroux C, Morel N, Tabouret G, Chabert A, Lugan S, Corbière F, Ferré P, Foucras G, Laude H, Eychenne F, Grassi J, Schelcher F. PrPSc accumulation in myocytes from sheep incubating natural scrapie. Nat Med. 2004 Jun;10(6):591-3.

• Angers RC, Browning SR, Seward TS, Sigurdson CJ, Miller MW, Hoover EA, Telling GC. Prions in skeletal muscles of deer with chronic wasting disease. Science. 2006 Feb 24;311(5764):1117.

• Ashwath ML, Dearmond SJ, Culclasure T, Prion-associated dilated cardiomyopathy. Arch Intern Med. 2005 Feb 14;165(3):338-40.

• Barbato R, Menabo R, Dainese P, Carafoli E, Schiaffino S, Di Lisa F. Binding of cytosolic proteins to myofibrils in ischemic rat hearts, Circ. Res. 78 (1996) 821–828.

• Basler K, Oesch B, Scott M, Westaway D, Walchli M, Groth DF, McKinley MP, Prusiner SB, Weissmann C. Scrapie and cellular PrP isoforms are encoded by the same chromosomal gene. Cell. 1986 Aug 1;46(3):417-28.

• Beckman JS, Koppenol WH. Nitric oxide, superoxide, and peroxynitrite: the good, the bad, and ugly. Am J Physiol. 1996 Nov;271(5 Pt 1):C1424-37.

• Bellinger-Kawahara C, Diener TO, McKinley MP, Groth DF, Smith DR, Prusiner SB. Purified scrapie prions resist inactivation by procedures that hydrolyze, modify, or shear nucleic acids. Virology. 1987 Sep;160(1):271-4.

• Bergmeyer HU, Bernt E. in: H.U. Bergmeyer (Ed.), Lactate Dehydrogenase. Weinheim, 1974, pp. 607–612

86

• Bianchin MM, Walz R, Brentani RR, Martins VR. Dilated cardiomyopathy and Creutzfeldt-Jakob disease: evidence for a role of

cellular prion protein in the heart? Arch Intern Med. 2005 Jul 25;165(14):1663-4.

• Bosque PJ, Ryou C, Telling G, Peretz D, Legname G, DeArmond SJ, Prusiner SB. Prions in skeletal muscle. Proc Natl Acad Sci U S A. 2002 Mar 19;99(6):3812-7.

• Bounhar Y, Zhang Y, Goodyer CG, LeBlanc A. Prion protein protects human neurons against Bax-mediated apoptosis. J Biol Chem. 2001 Oct 19;276(42):39145-9.

• Braasch W, Gudbjarnason S, Puri PS, Ravens KG, Bing RJ. Early changes in energy metabolism in the myocardium following acute

coronary artery occlusion in anesthetized dogs. Circ Res. 1968 Sep;23(3):429-38.

• Brandner S, Raeber A, Sailer A, Blättler T, Fischer M, Weissmann C, Aguzzi A. Normal host prion protein (PrPC) is required for scrapie spread within the central nervous system. Proc Natl Acad Sci U S A. 1996 Nov 12;93(23):13148-51.

• Brini M, Miuzzo M, Pierobon N, Negro A, Sorgato MC. The prion protein and its paralogue Doppel affect calcium signaling in Chinese

hamster ovary cells. Mol Biol Cell. 2005 Jun;16(6):2799-808.

• Brown DR, Besinger A. Prion protein expression and superoxide dismutase activity. Biochem J. 1998 Sep 1;334 ( Pt 2):423-9.

• Brown DR, Qin K, Herms JW, Madlung A, Manson J, Strome R, Fraser PE, Kruck T, von Bohlen A, Schulz-Schaeffer W, Giese A, Westaway D, Kretzschmar H. The cellular prion protein binds copper in vivo. Nature. 1997a; Dec 18-25;390(6661):684-7.

• Brown DR, Sassoon J. Copper-dependent functions for the prion protein. Mol Biotechnol. 2002 Oct;22(2):165-78.

• Brown DR, Schmidt B, Groschup MH, Kretzschmar HA. Prion protein expression in muscle cells and toxicity of a prion protein

fragment. Eur J Cell Biol. 1998 Jan;75(1):29-37.

• Brown DR, Schmidt B, Kretzschmar HA, Effects of oxidative stress on prion protein expression in Pc12 cells. Int. J. Dev. Neurosci. 1997b, 146:104-12.

• Brown DR, Schulz-Schaeffer WJ, Schmidt B, Kretzschmar HA, Prion protein-deficient cells show altered response to oxidative stress

due to decreased SOD-1 activity. Exp. Neurol. 1997c, 146: 104-12

• Brown DR, Wong BS, Hafiz F, Clive C, Haswell SJ, Jones IM. Normal prion protein has an activity like that of superoxide dismutase.

Biochem J. 1999 Nov 15;344 Pt 1:1-5.

• Büeler H, Fischer M, Lang Y, Bluethmann H, Lipp HP, DeArmond SJ, Prusiner SB, Aguet M, Weissmann C. Normal development and behaviour of mice lacking the neuronal cell-surface PrP protein. Nature. 1992; 356, 577-82.

• Canton M, Neverova I, Menabò R, Van Eyk J, Di Lisa F. Evidence of myofibrillar protein oxidation induced by postischemic reperfusion in

87

isolated rat hearts. Am J Physiol Heart Circ Physiol. 2004 Mar;286(3):H870-7.

• Canton M, Skyschally A, Menabò R, Boengler K, Gres P, Schulz R, Haude M, Erbel R, Di Lisa F, Heusch G. Oxidative modification of tropomyosin and myocardial dysfunction following coronary

microembolization. Eur Heart J. 2006 Apr;27(7):875-81.

• Carpi A, Menabò R, Kaludercic N, Pelicci P, Di Lisa F, Giorgio M. The cardioprotective effects elicited by p66(Shc) ablation demonstrate

the crucial role of mitochondrial ROS formation in ischemia/reperfusion

injury. Biochim Biophys Acta. 2009 Jul;1787(7):774-80.

• Caughey B, Brown K, Raymond GJ, Katzenstein GE, Thresher W. Binding of the protease-sensitive form of PrP (prion protein) to sulfated

glycosaminoglycan and congo red J Virol. 1994 Apr;68(4):2135-41. Erratum in: J Virol 1994 Jun;68(6):4107.

• Caughey B, Race R, Chesebro B. Comparative sequence analysis, in vitro expression and biosynthesis of mouse PrP. Prog Clin Biol Res. 1989;317:619-36.

• Caughey B, Raymond GJ. The scrapie-associated form of PrP is made from a cell surface precursor that is both protease- and phospholipase-

sensitive. J Biol Chem. 1991 Sep 25;266(27):18217-23.

• Cereghetti GM, Schweiger A, Glockshuber R, Van Doorslaer S. Electron paramagnetic resonance evidence for binding of Cu(2+) to the

C-terminal domain of the murine prion protein. Biophys J. 2001 Jul;81(1):516-25.

• Cesselli D, Jakoniuk I, Barlucchi L, Beltrami AP, Hintze TH, Nadal-Ginard B, Kajstura J, Leri A, Anversa P. Oxidative stress-mediated cardiac cell death is a major determinant of ventricular

dysfunction and failure in dog dilated cardiomyopathy. Circ Res. 2001 Aug 3;89(3):279-86.

• Chen S, Mangé A, Dong L, Lehmann S, Schachner M. Prion protein as trans-interacting partner for neurons is involved in neurite

outgrowth and neuronal survival. Mol Cell Neurosci. 2003 Feb;22(2):227-33.

• Chiesa R, Pestronk A, Schmidt RE, Tourtellotte WG, Ghetti B, Piccardo P, Harris DA. Primary myopathy and accumulation of PrPSc-like like molecules in peripheral tissues of transgenic mice ex

pressing a prion protein insertional mutation. Neurobiol Dis. 2001 Apr;8(2):279-88.

• Chrétien F, Dorandeu A, Adle-Biassette H, Ereau T, Wingertsmann L, Brion F, Gray F. A process of programmed cell death as a mechanisms of neuronal death in prion diseases. Clin Exp Pathol. 1999;47(3-4):181-91.

• Cohen FE, Pan KM, Huang Z, Baldwin M, Fletterick RJ, Prusiner SB. Structural clues to prion replication. Science. 1994 Apr 22;264(5158):530-1.

• Cohen FE. Protein misfolding and prion diseases. J Mol Biol. 1999 Oct 22;293(2):313-20.

88

• Coitinho AS, Freitas AR, Lopes MH, Hajj GN, Roesler R, Walz R, Rossato JI, Cammarota M, Izquierdo I, Martins VR, Brentani RR. The interaction between prion protein and laminin modulates memory

consolidation. Eur J Neurosci. 2006 Dec;24(11):3255-64.

• Collinge J, Owen F, Poulter M, Leach M, Crow TJ, Rossor MN, Hardy J, Mullan MJ, Janota I, Lantos PL. Prion dementia without characteristic pathology. Lancet. 1990 Jul 7;336(8706):7-9.

• Collins SR, Douglass A, Vale RD, Weissman JS. Mechanism of prion propagation: amyloid growth occurs by monomer addition. PLoS Biol. 2004 Oct;2(10):e321.

• Cosentino F, Francia P, Camici GG, Pelicci PG, Lüscher TF, Volpe M. Final common molecular pathways of aging and cardiovascular disease: role of the p66Shc protein. Arterioscler Thromb Vasc Biol. 2008 Apr;28(4):622-8.

• del Monte F, Lebeche D, Guerrero JL, Tsuji T, Doye AA, Gwathmey JK, Hajjar RJ. Abrogation of ventricular arrhythmias in a model of ischemia and reperfusion by targeting myocardial calcium

cycling. Proc Natl Acad Sci U S A. 2004 Apr 13;101(15):5622-7.

• Demaison L, Moreau D, Martine L, Chaudron I, Grynberg A. Myocardial ischemia and in vitro mitochondrial metabolic efficiency. Mol Cell Biochem. 1999 Apr;194(1-2):291-300.

• Di Lisa F, Menabò R, Canton M, Barile M, Bernardi P. Opening of the mitochondrial permeability transition pore causes depletion of

mitochondrial and cytosolic NAD+ and is a causative event in the death

of myocytes in postischemic reperfusion of the heart. J Biol Chem. 2001 Jan 26;276(4):2571-5.

• Diarra-Mehrpour M, Arrabal S, Jalil A, Pinson X, Gaudin C, Piétu G, Pitaval A, Ripoche H, Eloit M, Dormont D, Chouaib S. Prion protein prevents human breast carcinoma cell line from tumor necrosis

factor alpha-induced cell death. Cancer Res. 2004 Jan 15;64(2):719-27.

• Dorandeu A, Wingertsmann L, Chrétien F, Delisle MB, Vital C, Parchi P, Montagna P, Lugaresi E, Ironside JW, Budka H, Gambetti P, Gray F. Neuronal apoptosis in fatal familial insomnia. Brain Pathol. 1998 Jul;8(3):531-7.

• Downey JM, Cohen MV, Ytrehus K, Liu Y. Cellular mechanisms in ischemic preconditioning: the role of adenosine and protein kinase C. Ann N Y Acad Sci. 1994 Jun 17;723:82-98.

• Dröge W. Free radicals in the physiological control of cell function. Physiol Rev. 2002 Jan;82(1):47-95.

• Ettaiche M, Pichot R, Vincent JP, Chabry J. In vivo cytotoxicity of the prion protein fragment 106-126. J Biol Chem. 2000 Nov 24;275(47):36487-90.

• Evans MD, Dizdaroglu M, Cooke MS. Oxidative DNA damage and disease: induction, repair and significance. Mutat Res. 2004 Sep;567(1):1-61.

89

• Forloni G, Angeretti N, Chiesa R, Monzani E, Salmona M, Bugiani O, Tagliavini F. Neurotoxicity of a prion protein fragment. Nature. 1993 Apr 8;362(6420):543-6.

• Fornai F, Ferrucci M, Gesi M, Bandettini di Poggio A, Giorgi FS, Biagioni F, Paparelli A. A hypothesis on prion disorders: are infectious, inherited, and sporadic causes so distinct? Brain Res Bull. 2006 Mar 31;69(2):95-100.

• Fujio Y, Nguyen T, Wencker D, Kitsis RN, Walsh K. Akt promotes survival of cardiomyocytes in vitro and protects against ischemia-

reperfusion injury in mouse heart. Circulation. 2000 Feb 15;101(6):660-7.

• Garcia-Dorado D, Agulló L, Sartorio CL, Ruiz-Meana M. Myocardial protection against reperfusion injury: the cGMP pathway.

Thromb Haemost. 2009 Apr;101(4):635-42.

• Gauczynski S, Peyrin JM, Haïk S, Leucht C, Hundt C, Rieger R, Krasemann S, Deslys JP, Dormont D, Lasmézas CI, Weiss S. The 37-kDa/67-kDa laminin receptor acts as the cell-surface receptor for

the cellular prion protein. EMBO J. 2001 Nov 1;20(21):5863-75.

• Giorgio M, Migliaccio E, Orsini F, Paolucci D, Moroni M, Contursi C, Pelliccia G, Luzi L, Minucci S, Marcaccio M, Pinton P, Rizzuto R, Bernardi P, Paolucci F, Pelicci PG. Electron transfer between cytochrome c and p66Shc generates reactive oxygen species that

trigger mitochondrial apoptosis. Cell. 2005 Jul 29;122(2):221-33.

• Glatzel M, Abela E, Maissen M, Aguzzi A. Extraneural pathologic prion protein in sporadic Creutzfeldt-Jakob disease. N Engl J Med. 2003 Nov 6;349(19):1812-20.

• Graner E, Mercadante AF, Zanata SM, Forlenza OV, Cabral AL, Veiga SS, Juliano MA, Roesler R, Walz R, Minetti A, Izquierdo I, Martins VR, Brentani RR Cellular prion protein binds laminin and mediates neuritogenesis. Brain Res Mol Brain Res. 2000b Mar 10;76(1):85-92.

• Graner E, Mercadante AF, Zanata SM, Forlenza OV, Cabral AL, Veiga SS, Juliano MA, Roesler R, Walz R, Minetti A, Izquierdo I, Martins VR, Brentani RR. Cellular prion protein binds laminin and mediates neuritogenesis. Brain Res Mol Brain Res. 2000a Mar 10;76(1):85-92.

• Graner E, Mercadante AF, Zanata SM, Martins VR, Jay DG, Brentani RR. Laminin-induced PC-12 cell differentiation is inhibited following laser inactivation of cellular prion protein. FEBS Lett. 2000b Oct 6;482(3):257-60.

• Griffith, J.S. Self-replication and scrapie. Nature. 1967 215(105):1043-4.

• Hajj GN, Lopes MH, Mercadante AF, Veiga SS, da Silveira RB, Santos TG, Ribeiro KC, Juliano MA, Jacchieri SG, Zanata SM, Martins VR. Cellular prion protein interaction with vitronectin supports axonal growth and is compensated by integrins. J Cell Sci. 2007 Jun 1;120(Pt 11):1915-26.

• Hampton MB, Orrenius S. Redox regulation of apoptotic cell death. Biofactors. 1998;8(1-2):1-5.

90

• Hansen SM, Berezin V, Bock E. Signaling mechanisms of neurite outgrowth induced by the cell adhesion molecules NCAM and N-

cadherin. Cell Mol Life Sci. 2008 Nov;65(23):3809-21.

• Hausenloy DJ, Yellon DM. Survival kinases in ischemic

preconditioning and postconditioning. Cardiovasc Res. 2006 May 1;70(2):240-53.

• Hearse DJ. Free radicals and the heart. Bratisl Lek Listy. 1991 Mar-Apr;92(3-4):115-8.

• Hearse DJ. Ischemia at the crossroads? Cardiovasc Drugs Ther. 1988 May;2(1):9-15.

• Hearse DJ. Myocardial protection during ischemia and reperfusion. Mol Cell Biochem. 1998 Sep;186(1-2):177-84.

• Hill AF, Joiner S, Linehan J, Desbruslais M, Lantos PL, Collinge J. Species-barrier-independent prion replication in apparently resistant

species. Proc Natl Acad Sci U S A. 2000 Aug 29;97(18):10248-53.

• Huang S, Liang J, Zheng M, Li X, Wang M, Wang P, Vanegas D, Wu D, Chakraborty B, Hays AP, Chen K, Chen SG, Booth S, Cohen M, Gambetti P, Kong Q. Inducible overexpression of wild-type prion protein in the muscles leads to a primary myopathy in

transgenic mice. Proc Natl Acad Sci U S A. 2007 Apr 17;104(16):6800-5.

• Hundt C, Peyrin JM, Haïk S, Gauczynski S, Leucht C, Rieger R, Riley ML, Deslys JP, Dormont D, Lasmézas CI, Weiss S. Identification of interaction domains of the prion protein with its 37-

kDa/67-kDa laminin receptor. EMBO J. 2001 Nov 1;20(21):5876-86.

• Hutter G, Heppner FL, Aguzzi A. No superoxide dismutase activity of cellular prion protein in vivo. Biol Chem. 2003 Sep;384(9):1279-85.

• Ishikawa T, Akerboom TPM, Sies H. Target Organ Toxicity. CRC Press Inc., Boca Raton FL, 1986; 129–143.

• Jennings RB. Treatment of acute myocardial ischemia. West J Med. 1987 Jul;147(1):62-4.

• Jones DP, Thor H, Andersson B, Orrenius S. Detoxification reactions in isolated hepatocytes. Role of glutathione peroxidase,

catalase, and formaldehyde dehydrogenase in reactions relating to N-

demethylation by the cytochrome P-450 system. J Biol Chem. 1978 Sep 10;253(17):6031-7.

• Jones S, Batchelor M, Bhelt D, Clarke AR, Collinge J, Jackson GS. Recombinant prion protein does not possess SOD-1 activity. Biochem J. 2005 Dec 1;392(Pt 2):309-12.

• Kabouridis PS. Lipid rafts in T cell receptor signaling. Mol Membr Biol. 2006 Jan-Feb;23(1):49-57.

• Kanaani J, Prusiner SB, Diacovo J, Baekkeskov S, Legname G. Recombinant prion protein induces rapid polarization and development

of synapses in embryonic rat hippocampal neurons in vitro. J Neurochem. 2005 Dec;95(5):1373-86.

91

• Kang YJ, Chen Y, Epstein PN. Suppression of doxorubicin cardiotoxicity by overexpression of catalase in the heart of transgenic

mice. J Biol Chem. 1996 May 24;271(21):12610-6.

• Kim BH, Lee HG, Choi JK, Kim JI, Choi EK, Carp RI, Kim YS. The cellular prion protein (PrPC) prevents apoptotic neuronal cell death and

mitochondrial dysfunction induced by serum deprivation. Brain Res Mol Brain Res. 2004 Apr 29;124(1):40-50.

• Klamt F, Dal-Pizzol F, Conte da Frota ML JR, Walz R, Andrades ME, da Silva EG, Brentani RR, Izquierdo I, Fonseca Moreira JC. Imbalance of antioxidant defense in mice lacking cellular prion protein.

Free Radic Biol Med. 2001 May 15;30(10):1137-44.

• Knight RS, Will RG. Prion diseases. J Neurol Neurosurg Psychiatry. 2004 Mar;75 Suppl 1:i36-42.

• Kovacs GC, Linleck-Pozza E, Chimelli L, Araujo AQ, Gabbai AA, Strobel T, Glatzel M, Aguzzi A, Budke H, Creutzfeld-Jacob and inclusion body myositis: abundant disease-associated prion protein in

muscle. Ann Neurol. 2004 Jan;55(1):121-5.

• Kramer ML, Kratzin HD, Schmidt B, Römer A, Windl O, Liemann S, Hornemann S, Kretzschmar H. Prion protein binds copper within the physiological concentration range. J Biol Chem. 2001 May 18;276(20):16711-9.

• Langendorff, O. Untersuchungen am uber lebenden Sangetier Herzen. Pflüegers Archiv. 1985; 61 : 291-322.

• Lasmézas CI, Deslys JP, Robain O, Jaegly A, Beringue V, Peyrin JM, Fournier JG, Hauw JJ, Rossier J, Dormont D. Transmission of the BSE agent to mice in the absence of detectable abnormal prion

protein.Science. 1997 Jan 17;275(5298):402-5.

• Lee JA, Allen DG. Mechanism of acute ischemic contractile failure of heart. Role of intracellular calcium. J Clin Invest. 1991 Aug;88(2):361-7.

• Lefer AM, Weyrich AS, Buerke M. Role of selectins, a new family of adhesion molecules, in ischaemia-reperfusion injury. Cardiovasc Res. 1994 Mar;28(3):289-94.

• Lemasters JJ, Thurman RG. The many facets of reperfusion injury. Gastroenterology. 1995 Apr;108(4):1317-20.

• Levraut J, Iwase H, Shao ZH, Vanden Hoek TL, Schumacker PT. Cell death during ischemia: relationship to mitochondrial depolarization

and ROS generation. Am J Physiol Heart Circ Physiol. 2003 Feb;284(2):H549-58.

• Li Y, Mei GY, Jiang HY, Wang GR, Tian C, Chen C, Wang X, Wang KX, Han J, Dong XP. The construction and identification of the PRNP vectors with ubiquitin or the lysosome-targeting signal. Zhonghua Shi Yan He Lin Chuang Bing Du Xue Za Zhi. 2008 Dec;22(6):419-21.

• Linden R, Martins VR, Prado MA, Cammarota M, Izquierdo I, Brentani RR, Physiology of the prion protein. Physiol Rev. 2008 Apr;88(2):673-728.

92

• Liu H, Cala PM, Anderson SE. Ischemic preconditioning: effects on pH, Na and Ca in newborn rabbit hearts during Ischemia/Reperfusion. J Mol Cell Cardiol. 1998 Mar;30(3):685-97.

• Lopes MH, Hajj GN, Muras AG, Mancini GL, Castro RM, Ribeiro KC, Brentani RR, Linden R, Martins VR. Interaction of cellular prion and stress-inducible protein 1 promotes neuritogenesis and

neuroprotection by distinct signaling pathways. J Neurosci. 2005 Dec 7;25(49):11330-9.

• Málaga-Trillo E, Sempou E. PrPs: Proteins with a purpose: Lessons from the zebrafish. Prion. 2009 Jul;3(3):129-33.

• Mallucci G, Collinge J. Update on Creutzfeldt-Jakob disease. Curr Opin Neurol. 2004 Dec;17(6):641-7.

• Mallucci GR, Ratte S, Asante EA, Linehan J, Gowland I, Jefferys JG, Collinge J. Post-natal knockout of prion protein alters hippocampal CA1 properties, but does not result in neurodegeneration.

EMBO J. 2002 Feb 1;21(3):202-10

• Mangé A, Béranger F, Peoc'h K, Onodera T, Frobert Y, Lehmann S. Alpha- and beta- cleavages of the amino-terminus of the cellular prion protein. Biol Cell. 2004 Mar;96(2):125-32.

• Manson JC, Clarke AR, Hooper ML, Aitchison L, McConnell I, Hope J. 129/Ola mice carrying a null mutation in PrP that abolishes mRNA production are developmentally normal. Mol Neurobiol. 1994 Apr-Jun;8(2-3):121-7.

• Marber M. Viewpoint: stunning and longterm perfusion-contraction matching are clinically indistinguishable components of clinical

hibernation and separation, even if practical, is unlikely to matter. Basic Res Cardiol. 1997;92 Suppl 2:26-9.

• Marciano PG, Brettschneider J, Manduchi E, Davis JE, Eastman S, Raghupathi R, Saatman KE, Speed TP, Stoeckert CJ Jr, Eberwine JH, McIntosh TK. Neuron-specific mRNA complexity responses during hippocampal apoptosis after traumatic brain injury. J Neurosci. 2004 Mar 24;24(12):2866-76.

• Maroko PR, Carpenter CB, Chiariello M, Fishbein MC, Radvany P, Knostman JD, Hale SL. Reduction by cobra venom factor of myocardial necrosis after coronary artery occlusion. J Clin Invest. 1978 Mar;61(3):661-70.

• Massimino ML, Ballarin C, Bertoli A, Casonato S, Genovesi S, Negro A, Sorgato MC. Human Doppel and prion protein share common membrane microdomains and internalization pathways. Int J Biochem Cell Biol. 2004 Oct;36(10):2016-31.

• Massimino ML, Ferrari J, Sorgato MC, Bertoli A. Heterogeneous PrPC metabolism in skeletal muscle cells. FEBS Lett. 2006 Feb 6;580(3):878-84.

• McCord JM, Fridovich I. The utility of superoxide dismutase in studying free radical reactions. I. Radicals generated by the interaction

of sulfite, dimethyl sulfoxide, and oxygen. J Biol Chem. 1969 Nov 25;244(22):6056-63.

93

• McKerracher L, Chamoux M, Arregui CO. Role of laminin and integrin interactions in growth cone guidance. Mol Neurobiol. 1996 Apr;12(2):95-116.

• McLennan NF, Brennan PM, McNeill A, Davies I, Fotheringham A, Rennison KA, Ritchie D, Brannan F, Head MW, Ironside JW, Williams A, Bell JE. Prion protein accumulation and neuroprotection in hypoxic brain damage. Am J Pathol. 2004 Jul;165(1):227-35.

• McMahon HE, Mangé A, Nishida N, Créminon C, Casanova D, Lehmann S. Cleavage of the amino terminus of the prion protein by reactive oxygen species. J Biol Chem. 2001 Jan 19;276(3):2286-91.

• Medori R, Montagna P, Tritschler HJ, LeBlanc A, Cortelli P, Tinuper P, Lugaresi E, Gambetti P. Fatal familial insomnia: a second kindred with mutation of prion protein gene at codon 178. Neurology. 1992 Mar;42(3 Pt 1):669-70.

• Miao W, Luo Z, Kitsis RN, Walsh K. Intracoronary, adenovirus-mediated Akt gene transfer in heart limits infarct size following

ischemia-reperfusion injury in vivo. J Mol Cell Cardiol. 2000 Dec;32(12):2397-402.

• Miele G, Alejo Blanco AR, Baybutt H, Horvat S, Manson J, Clinton M. Embryonic activation and developmental expression of the murine prion protein gene. Gene Expr. 2003;11(1):1-12.

• Migliaccio E, Giorgio M, Mele S, Pelicci G, Reboldi P, Pandolfi PP, Lanfrancone L, Pelicci PG. The p66shc adaptor protein controls oxidative stress response and life span in mammals. Nature. 1999;402:309 –313.

• Migliaccio E, Mele S, Salcini AE, Pelicci G, Lai KM, Superti-Furga G, Pawson T, Di Fiore PP, Lanfrancone L, Pelicci PG. Opposite effects of the p52shc/p46shc and p66Shc splicing isoforms on the EGF

receptor-MAP kinase-fos signalling pathway. EMBO J. 1997 Feb 17;16(4):706-16.

• Milhavet O, Lehmann S. Oxidative stress and the prion protein in transmissible spongiform encephalopathies. Brain Res Brain Res Rev. 2002 Feb;38(3):328-39.

• Mitteregger G, Vosko M, Krebs B, Xiang W, Kohlmannsperger V, Nölting S, Hamann GF, Kretzschmar HA. The role of the octarepeat region in neuroprotective function of the cellular prion protein.Brain Pathol. 2007 Apr;17(2):174-83.

• Monnet C, Gavard J, Mège RM, Sobel A. Clustering of cellular prion protein induces ERK1/2 and stathmin phosphorylation in GT1-7

neuronal cells. FEBS Lett. 2004 Oct 8;576(1-2):114-8.

• Moore RC, Lee IY, Silverman GL, Harrison PM, Strome R, Heinrich C, Karunaratne A, Pasternak SH, Chishti MA, Liang Y, Mastrangelo P, Wang K, Smit AF, Katamine S, Carlson GA, Cohen FE, Prusiner SB, Melton DW, Tremblay P, Hood LE, Westaway D. Ataxia in prion protein (PrP)-deficient mice is associated with upregulation of the novel PrP-like protein doppel. J Mol Biol. 1999 Oct 1;292(4):797-817.

• Moore RC, Mastrangelo P, Bouzamondo E, Heinrich C, Legname G, Prusiner SB, Hood L, Westaway D, DeArmond SJ, Tremblay P.

94

Doppel-induced cerebellar degeneration in transgenic mice. Proc Natl Acad Sci U S A. 2001 Dec 18;98(26):15288-93.

• Mouillet-Richard, S., Ermonval, M., Chebassier, C., Laplanche, J.L., Lehmann, S., Launay, J.M., Kellermann, O. Signal transduction through prion protein. Science, 2000 289, 1925-1928.

• Murry CE, Jennings RB, Reimer KA. Preconditioning with ischemia: a delay of lethal cell injury in ischemic myocardium. Circulation. 1986 Nov;74(5):1124-36.

• Napoli C, Martin-Padura I, de Nigris F, Giorgio M, Mansueto G, Somma P, Condorelli M, Sica G, De Rosa G, Pelicci P. Deletion of the p66Shc longevity gene reduces systemic and tissue oxidative stress,

vascular cell apoptosis, and early atherogenesis in mice fed a high-fat

diet. Proc Natl Acad Sci U S A. 2003;100:2112–2116.

• Nazor KE, Seward T, Telling GC. Motor behaviour and neuropathological deficits in mice deficient form normal prion protein

expression. Biochim Biophys Acta. 2007 Jun;1772(6):645-53.

• Nico PB, Lobão-Soares B, Landemberger MC, Marques W Jr, Tasca CI, de Mello CF, Walz R, Carlotti CG Jr, Brentani RR, Sakamoto AC, Bianchin MM. Impaired exercise capacity, but unaltered mitochondrial respiration in skeletal or cardiac muscle of

mice lacking cellular prion protein. Neurosci Lett. 2005 Nov 4;388(1):21-6.

• Obreztchikova M, Elouardighi H, Ho M, Wilson BA, Gertsberg Z, Steinberg SF. Distinct signaling functions for Shc isoforms in the heart. J Biol Chem. 2006 Jul 21;281(29):20197-204.

• Okado-Matsumoto A, Fridovich I. Assay of superoxide dismutase: cautions relevant to the use of cytochrome c, a sulfonated tetrazolium,

and cyanide. Anal Biochem. 2001 Nov 15;298(2):337-42.

• Oudot A, Martin C, Busseuil D, Vergely C, Demaison L, Rochette L. NADPH oxidases are in part responsible for increased cardiovascular superoxide production during aging. Free Radic Biol Med. 2006 Jun 15;40(12):2214-22.

• Pain T, Yang XM, Critz SD, Yue Y, Nakano A, Liu GS, Heusch G, Cohen MV, Downey JM. Opening of mitochondrial K(ATP) channels triggers the preconditioned state by generating free radicals. Circ Res. 2000 Sep 15;87(6):460-6.

• Pan KM, Baldwin M, Nguyen, Gasset M, Serban A, Groth D, Mehlhorn I, Huang Z, Fletterick RJ, Cohen FE, Conversion of alpha-helices into beta sheet features in the formation of scrapie prion

protein. Proc. Natl. Acad. Sci. USA 1993, 90: 10962-6.

• Pan T, Wong BS, Liu T, Li R, Petersen RB, Sy MS. Cell-surface prion protein interacts with glycosaminoglycans. Biochem J. 2002 Nov 15;368(Pt 1):81-90.

• Parkin ET, Watt NT, Turner AJ, Hooper NM. .Dual mechanisms for shedding of the cellular prion protein. Biol Chem. 2004 Mar 19;279(12):11170-8.

95

• Paterson AW, Curtis JC, Macleod NK. Complex I specific increase in superoxide formation and respiration rate by PrP-null mouse brain

mitochondria. J Neurochem. 2008 Apr;105(1):177-91.

• Pauly PC, Harris DA Copper stimulates endocytosis of the prion protein. J Biol Chem. 1998 Dec 11;273(50):33107-10

• Peden AH, Ritchie DL, Head MW, Ironside JW. Detection and localization of PrPSc in the skeletal muscle of patients with variant,

iatrogenic, and sporadic forms of Creutzfeldt-Jakob disease. Am J Pathol. 2006 Mar;168(3):927-35.

• Peters PJ, Mironov A Jr, Peretz D, van Donselaar E, Leclerc E, Erpel S, DeArmond SJ, Burton DR, Williamson RA, Vey M, Prusiner SB. Trafficking of prion proteins through a caveolae-mediated endosomal pathway. J Cell Biol. 2003 Aug 18;162(4):703-17.

• Peterson DB, Holian J, Garrison WM. Radiation chemistry of the alpha-amino acids. Gamma radiolysis of solid cysteine. J Phys Chem. 1969 May;73(5):1568-72.

• Prince RC, Gunson DE. Prions are copper-binding proteins. Trends Biochem Sci. 1998 Jun;23(6):197-8.

• Prusiner SB. Prions. Proc. Natl. Acad. Sci. 1998, 95, 13363-13383.

• Prusiner SB. Prions: novel infectious pathogens. Adv Virus Res. 1984;29:1-56.

• Race R, Chesebro B. Scrapie infectivity found in resistant species. Nature. 1998 Apr 23;392(6678):770.

• Radovanovic I, Braun N, Giger OT, Mertz K, Miele G, Prinz M, Navarro B, Aguzzi A. Truncated prion protein and Doppel are myelinotoxic in the absence of oligodendrocytic PrPC. J Neurosci. 2005 May 11;25(19):4879-88.

• Reimer KA, Jennings RB. The “wavefront phenomenon” of myocardial ischemic cell death. II. Trensmural progressin of necrosis

within the framework of ischemic deb size (myocardium at risk) and

collateral flow. Lab Invest. 1979 Jun;40(6):633-44.

• Rieger R, Edenhofer F, Lasmezas CI, Weiss S. The human 37-kDa laminin receptor precursor interacts with the prion protein in

eukaryotic cells. Nat Med. 1997 Dec;3(12):1383-8.

• Riek R, Hornemann S, Wider G, Billeter M, Glockshuber R, Wüthrich K. NMR structure of the mouse prion protein domain PrP(121-321). Nature. 1996 Jul 11;382(6587):180-2.

• Rodgers MA, Sokol HA, Garrison WM. The radiation-induced “hydrolysis” of the peptide boned. J Am Chem Soc. 1968 Jan 31;90(3):795-6.

• Rossi D, Cozzio A, Flechsig E, Klein MA, Rülicke T, Aguzzi A, Weissmann C. Onset of ataxia and Purkinje cell loss in PrP null mice inversely correlated with Dpl level in brain. EMBO J. 2001 Feb 15;20(4):694-702.

96

• Safar J, Roller PP, Gajduesek DC, Gibbs CJ Jr., Thermal stability and conformational transitions of scrapie amyloid (prion) protein

correlate with infectivity. Protein Sci. 1993, 2: 2206-16.

• Samaia HB, Brentani RR. Can loss-of-function prion-related diseases exist? Mol Psychiatry. 1998 May;3(3):196-7.

• Santuccione A, Sytnyk V, Leshchyns’ka I, Schachner M. Prion protein recruits its neuronal receptor NCAM to lipid rafts to activate

p59Fyn and to enhance neurite outgrowth. J Cell Biol. 2005 Apr 25;169(2):341-54.

• Sarkozi E, Askanas V, Engel WK. Abnormal accumulation of prion protein mRNA in muscle fibres of patients with sporadic inclusion-body

myositis and hereditary inclusion-body myopathy. Am J Pathol. 1994 Dec;145(6):1280-4.

• Schluter KD, Schwartz P, Siegmund B, Piper HM. Prevention of the oxygen paradox in hypoxic-reoxygenated hearts. Am J Physiol. 1991 Aug;261(2 Pt 2):H416-23.

• Schmitt-Ulms G, Legname G, Baldwin MA, Ball HL, Bradon N, Bosque PJ, Crossin KL, Edelman GM, DeArmond SJ, Cohen FE, Prusiner SB. Binding of neural cell adhesion molecules (N-CAMs) to the cellular prion protein. J Mol Biol. 2001 Dec 14;314(5):1209-25.

• Schneider B, Mutel V, Pietri M, Ermoval M, Mouillet-Richard S, Kellermann O. NADPH oxidase and extracellular regulated kinases1/2 are targets of prions protein signalling in neuronal and nonneuronal

cells. Proc Natl Acad Sci U S A. 2003 Nov 11;100(23):13326-31.

• Schulz R. Intracellular targets of matrix metalloproteinase-2 in cardiac disease: rationale and therapeutic approaches. Annu Rev Pharmacol Toxicol. 2007;47:211-42.

• Shmerling D, Hegyi I, Fischer M, Blättler T, Brandner S, Götz J, Rülicke T, Flechsig E, Cozzio A, von Mering C, Hangartner C, Aguzzi A, Weissmann C. Expression of amino-terminally truncated PrP in the mouse leading to ataxia and specific cerebellar lesions. Cell. 1998 Apr 17;93(2):203-14.

• Shyu WC, Lin SZ, Chiang MF, Ding DC, Li KW, Chen SF, Yang HI, Li H. Overexpression of PrPC by Adenovirus-mediated gene targeting reduces ischemic injury in a stroke rat model. J Neurosci. 2005 Sep 28;25(39):8967-77.

• Simons K, Ikonen E. Functional rafts in cell membranes. Nature. 1997 Jun 5;387(6633):569-72.

• Sommerschild HT, Kirkebøen KA. Preconditioning endogenous defence mechanisms of the heart. Acta Anaesthesiol Scand. 2002 Feb;46(2):123-37.

• Sorgato MC, Bertoli A. From cell protection to death: may Ca2+ signals explain the chameleonic attributes of the mammalian prion

protein? Biochem Biophys Res Commun. 2009 Feb 6;379(2):171-4.

• Sorgato MC, Peggion C, Bertoli A. Is, indeed, the prion protein a Harlequin servant of "many" masters? Prion. 2009 Oct;3(4):202-5.

• Sparkes RS, Simon M, Cohn VH, Fournier RE, Lem J, Klisak I, Heinzmann C, Blatt C, Lucero M, Mohandas T. Assignment of the

97

human and mouse prion protein genes to homologous chromosomes.

Proc Natl Acad Sci U S A. 1986 Oct;83(19):7358-62.

• Spudich A, Frigg R, Kilic E, Kilic U, Oesch B, Raeber A, Bassetti CL, Hermann DM. Aggravation of ischemic brain injury by prion protein deficiency: role of ERK-1/-2 and STAT-1. Neurobiol Dis. 2005 Nov;20(2):442-9.

• Steele AD, Lindquist S, Aguzzi A. The prion protein knockout mouse: a phenotype under challenge. Prion. 2007 Apr;1(2):83-93.

• Steele AD, Zhou Z, Jackson WS, Zhu C, Auluck P, Moskowitz MA, Chesselet MF, Lindquist S. Context dependent neuroprotective properties of prion protein (PrP).Prion. 2009 Oct 16;3(4).

• Stuermer CA, Langhorst MF, Wiechers MF, Legler DF, Von Hanwehr SH, Guse AH, Plattner H. PrPc capping in T cells promotes its association with the lipid raft proteins reggie-1 and reggie-2 and

leads to signal transduction. FASEB J. 2004 Nov;18(14):1731-3.

• Taylor DR, Hooper NM. The prion protein and lipid rafts. Mol Membr Biol. 2006 Jan-Feb;23(1):89-99.

• Thomzig A, Schulz-Schaeffer W, Kratzel C, Mai J, Beekes M. Preclinical deposition of pathological prion protein PrPSc in muscles of

hamsters orally exposed to scrapie. J Clin Invest. 2004 May;113(10):1465-72.

• Ukeda H, Maeda S, Ishii T, Sawamura M. Spectrophotometric assay for superoxide dismutase based on tetrazolium salt 3'--1--

(phenylamino)-carbonyl--3, 4-tetrazolium]-bis(4-methoxy-6-

nitro)benzenesulfonic acid hydrate reduction by xanthine-xanthine

oxidase. Anal Biochem. 1997 Sep 5;251(2):206-9.

• Uptain SM, Lindquist S. Prions as protein-based genetic elements. Annu Rev Microbiol. 2002;56:703-41.

• Valko M, Morris H, Cronin MT. Metals, toxicity and oxidative stress. Curr Med Chem. 2005;12(10):1161-208.

• Vanden Hoek TL, Becker LB, Shao Z, Li C, Schumacker PT. Reactive oxygen species released from mitochondria during brief

hypoxia induce preconditioning in cardiomyocytes. J Biol Chem. 1998 Jul 17;273(29):18092-8.

• Vanden Hoek TL, Shao Z, Li C, Schumacker PT, Backer LB. Mitochondrial electron transport can become a significant source of

oxidative injury in cardiomyocytes. J Mol Cell Cardiol. 1997 Sep;29(9):2441-50.

• Vey M, Pilkuhn S, Wille H, Nixon R, DeArmond SJ, Smart EJ, Anderson RG, Taraboulos A, Prusiner SB. Subcellular colocalization of the cellular and scrapie prion proteins in caveolae-like membranous

domains. Proc Natl Acad Sci U S A. 1996 Dec 10;93(25):14945-9.

• Vincent B, Paitel E, Saftig P, Frobert Y, Hartmann D, De Strooper B, Grassi J, Lopez-Perez E, Checler F. The disintegrins ADAM10 and TACE contribute to the constitutive and phorbol ester-

regulated normal cleavage of the cellular prion protein. J Biol Chem. 2001 Oct 12;276(41):37743-6.

98

• Waggoner DJ, Drisaldi B, Bartnikas TB, Casareno RL, Prohaska JR, Gitlin JD, Harris DA. Brain copper content and cuproenzyme activity do not vary with prion protein expression level. J Biol Chem. 2000 Mar 17;275(11):7455-8.

• Watt NT, Taylor DR, Gillott A, Thomas DA, Perera WS, Hooper NM. Reactive oxygen species-mediated beta-cleavage of the prion protein in the cellular response to oxidative stress. J Biol Chem. 2005 Oct 28;280(43):35914-21.

• Weise J, Crome O, Sandau R, Schulz-Schaeffer W, Bähr M, Zerr I. Upregulation of cellular prion protein (PrPC) after focal cerebral ischemia and influence of lesion severity. Neurosci Lett. 2004 Nov 30;372(1-2):146-50.

• Weise J, Doeppner TR, Müller T, Wrede A, Schulz-Schaeffer W, Zerr I, Witte OW, Bähr M. Overexpression of cellular prion protein alters postischemic Erk1/2 phosphorylation but not Akt

phosphorylation and protects against focal cerebral ischemia. Restor Neurol Neurosci. 2008;26(1):57-64.

• Weise J, Sandau R, Schwarting S, Crome O, Wrede A, Schulz-Schaeffer W, Zerr I, Bähr M. Deletion of cellular prion protein results in reduced Akt activation, enhanced postischemic caspase-3 activation,

and exacerbation of ischemic brain injury. Stroke. 2006 May;37(5):1296-300.

• Weissmann C, Flechsig E. PrP knock-out and PrP transgenic mice in prion research. Br Med Bull. 2003;66:43-60.

• Weissmann C. Molecular genetics of transmissible spongiform encephalopathies. J Biol Chem. 1999 Jan 1;274(1):3-6.

• White AR, Collins SJ, Maher F, Jobling MF, Stewart LR, Thyer JM, Beyreuther K, Masters CL, Cappai R. Prion protein-deficient neurons reveal lower glutathione reductase activity and increased

susceptibility to hydrogen peroxide toxicity. Am J Pathol. 1999 Nov;155(5):1723-30.

• Wong BS, Brown DR, Pan T, Whiteman M, Liu T, Bu X, Li R, Gambetti P, Olesik J, Rubenstein R, Sy MS. Oxidative impairment in scrapie-infected mice is associated with brain metals perturbations

and altered antioxidant activities. J Neurochem. 2001 Nov;79(3):689-98.

• Zaccagnini G, Martelli F, Fasanaro P, Magenta A, Gaetano C, Di Carlo A, Biglioli P, Giorgio M, Martin-Padura I, Pelicci PG, Capogrossi MC. p66ShcA modulates tissue response to hindlimb ischemia. Circulation. 2004;109:2917–2923.

• 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. Stress-inducible protein 1 is a cell surface ligand for cellular prion that triggers neuroprotection. EMBO J. 2002 Jul 1;21(13):3307-16.

• Zanusso G, Vattemi G, Ferrari S, Tabaton M, Pecini E, Cavallaro T, Tomelleri G, Filosto M, Tonin P, Nardelli E, Rizzuto N, Monaco S. Increased expression of the normal cellular isoform of prion protein

99

in inclusion-body myositis, inflammatory myopathies and denervation

atrophy. Brain Pathol. 2001 Apr;11(2):182-9.

• Zhang CC, Steele AD, Lindquist S, Lodish HF. Prion protein is expressed on long-term repopulating hematopoietic stem cells and is

important for their self-renewal. Proc Natl Acad Sci U S A. 2006 Feb 14;103(7):2184-9.

• Zielonka J, Vasquez-Vivar J, Kalyanaraman B. Detection of 2-hydroxyethidium in cellular systems: a unique marker product of

superoxide and hydroethidine. Nat Protoc. 2008;3(1):8-21.

• Zweier JL, Flaherty JT, Weisfeldt ML. Direct measurement of free radical generation following reperfusion of ischemic myocardium. Proc Natl Acad Sci U S A. 1987 Mar;84(5):1404-7.

100

101

AKNOWLEDGEMENTS

Prof. MC Sorgato and Prof. Di Lisa for giving me the possibility to do this

research. Dr. Alessandro Bertoli, whose competence has been fundamental for

my work. Dr. Roberta Menabò, Dr. Andrea Carpi, , Dr. Maria Lina Massimino,

Dr. Cristian Lazzari, Dr. Roberto Stella, Dr. Caterina Peggion, Dr. Luca

Rizzetto.

My sweetheart Ilaria, my Parents who supported me during these years, all

my friends.