structural studies of heterogeneous amyloid species of...

68
Structural studies of heterogeneous amyloid species of lysozymes and de novo protein albebetin and their cytotoxicity Vladimir Zamotin Department of Medical Biochemistry and Biophysics Umeå University Umeå, Sweden 2007

Upload: trinhmien

Post on 12-Aug-2019

213 views

Category:

Documents


0 download

TRANSCRIPT

Structural studies of heterogeneous

amyloid species of lysozymes and de novo protein albebetin and their

cytotoxicity

Vladimir Zamotin

Department of Medical Biochemistry and Biophysics Umeå University

Umeå, Sweden 2007

3

Моей маме, жене и доче

To my family

4

5

Abstract A number of diseases are linked to protein folding problems which lead to

the deposition of insoluble protein plaques in the brain or other organs. These diseases include prion diseases such as Creutzfeld-Jakob disease, Alzheimer's disease, Parkinson's disease and type II (non-insulin dependent) diabetes. The protein plaques are found to consist of amyloid fibrils - cross-beta-sheet polymers with the beta-strands arranged perpendicular to the long axis of the fibre. Studies of ex vivo fibrils and fibrils produced in vitro showed that amyloid structures possess similar tinctorial and morphological properties. These suggest that the ability to form amyloid fibrils is an inherent property of polypeptide chains.

The aims of this thesis were to investigate the structural properties of cytotoxic amyloid and examine the involved mechanisms. The model proteins used in the studies were the equine and hen lysozymes and de novo designed protein albebetin.

Lysozymes are naturally ubiquitous proteins. Equine lysozyme belongs to an extended family of structurally related lysozymes and α-lactalbumins and can be considered as an evolutional bridge between them. Hen lysozyme is one of the most characterized protein and its amyloidogenic properties were described earlier. De novo protein albebetin and its constructs are designed to perform the function of grafted polypeptide sequence. Fibrils of equine lysozyme are formed at acidic pH and elevated temperatures where a partially folded molten globule state is populated. We have shown that lysozyme assembles into annular and linear protofilaments in a calcium-dependent manner. We showed that albebetin and its constructs are inherently highly amyloidogenic under physiological conditions. Fibrillation proceeds via multiple pathways and includes a hierarchy of amyloid structures ranging from oligomers to protofilaments and fibrils, among which two distinct types of oligomeric intermediates were characterized. Pivotal oligomers comprise of 10-12 monomers and on-pathway amyloid-prone oligomers constitute of 26-30 molecules. We suggest that transformation of the pivotal oligomers into the amyloid-prone ones is a limiting stage in albebetin fibrillation. Cytotoxic studies of albebetin amyloid species have revealed that initial, pivotal oligomers do not effect on cell viability while amyloid-prone ones induce cell death. We suggest that oligomeric size is important for the stabilizing cross-beta-sheet core which is crucial for cell toxicity. Cytotoxic studies of both oligomers and fibrils of hen lysozyme have revealed that both species induce cell death. The amyloid sample containing cross-β-sheet oligomers induces an apoptosis-like cell death. The oligomers without cross-β-sheet appeared to be non-toxic, indicating that the stabilization of this structural pattern is critical for the induced toxicity. In contrast, the fibrils induce more rapid, necrosis-like death.

These studies gained insights into a structure–function relationship of different forms of amyloid and general pathways of cell death. This is an important step in understanding the mechanisms of amyloid-associated degeneration and defining specific therapeutic targets.

6

Contents

ABSTRACT 5

CONTENTS 6

LIST OF PUBLICATIONS 8

INTRODUCTION 10

Amyloid diseases and related proteins 10

MISFOLDING AND AGGREGATION 12

Generic properties of amyloids 12

Amyloid precursor state 16

Conditions of amyloid formation 18

Effect of mutation on amyloidogenic properties 21

Models of amyloid formation 23

CYTOTOXICITY OF AMYLOIDS 26

Amyloid toxic species 26

Cellular targets for amyloids 27

General pathways of cell death 30

STUDIED PROTEINS 33

Lysozymes 33

De novo designed protein albebetin and its constructs 37

7

APPLICATION OF ATOMIC FORCE MICROSCOPY IN AMYLOID STUDIES 39

AFM volume measurements 41

INTRODUCTION TO THE ARTICLES 44

CONCLUSIONS 50

ACKNOWLEDGEMENTS 52

REFERENCES 53

8

List of publications

The thesis is based upon the following papers:

I. Malisauskas, M., Zamotin, V., Jass, J., Noppe, W., Dobson, C.M.,

Morozova-Roche, L.A. (2003) Amyloid protofilaments from the

calcium-binding protein equine lysozyme: formation of ring and linear

structures depends on pH and metal ion concentration. J. Mol. Biol.

330, 879-890.

II. Morozova-Roche,* L.A., Zamotin,* V., Malisauskas, M., Ohman, A.,

Chertkova, R., Lavrikova, M.A., Kostanyan, I.A., Dolgikh, D.A.,

Kirpichnikov, M.P. (2004) Fibrillation of carrier protein albebetin and

its biologically active constructs. Multiple oligomeric intermediates and

pathways. Biochemistry, 43, 9610-9619. (*- with equal contributions)

III. Zamotin,* V., Gharibyan,* A., Gibanova, N.V., Lavrikova, M.A.,

Dolgikh, D.A., Kirpichnikov, M.P., Kostanyan, I.A., Morozova-Roche,

L.A. (2006) Cytotoxicity of albebetin oligomers depends on cross-β-

sheet formation. FEBS Letters, 580, 2451-2457. (*- with equal

contributions)

IV. Gharibyan, A.L., Zamotin, V., Yanamandra, K., Moskaleva, O.S.,

Margulis, B.A., Kostanyan, I.A., Morozova-Roche, L.A. (2007)

Lysozyme Amyloid Oligomers and Fibrils Induce Cellular Death via

Different Apoptotic/Necrotic Pathways. J. Mol. Biol. 365, 1337-1349.

9

The articles not included in the thesis:

1. Gruden, M.A., Davudova, T.B., Malisauskas, M., Zamotin, V.V.,

Sewell, R.D., Voskresenskaya, N.I., Kostanyan, I.A., Sherstnev, V.V.,

Morozova-Roche, L.A. (2004) Autoimmune responses to amyloid

structures of Aβ(25-35) peptide and human lysozyme in the serum of

patients with progressive Alzheimer's disease. Dement. Geriatr. Cogn.

Disord. 18, 165-171.

2. Malisauskas, M., Ostman, J., Darinskas, A., Zamotin, V., Liutkevicius,

E., Lundgren, E., Morozova-Roche, L.A. (2005) Does the cytotoxic

effect of transient amyloid oligomers from common equine lysozyme in

vitro imply innate amyloid toxicity? J. Biol. Chem. 280, 6269-6275.

3. Lavrikova, M.A., Zamotin, V.V., Malisauskas, M., Chertkova, R.V.,

Kostanyan, I.A., Dolgikh, D.A., Kirpichnikov, M.P., Morozova-Roche,

L.A. (2006) Amyloidogenic properties of the artificial protein albebetin

and its biologically active derivatives. The role of electrostatic

interactions in fibril formation. Biochemistry (Mosc). 71, 306-314.

4. Malisauskas, M., Darinskas, A., Zamotin, V.V., Gharibyan, A.,

Kostanyan, I.A., Morozova-Roche, L.A. (2006) Intermediate amyloid

oligomers of lysozyme: is their cytotoxicity a particular case or general

rule for amyloid? Biochemistry (Mosc), 71, 505-512.

5. Wilhelm, K.R., Yanamandra, K., Gruden, M.A., Zamotin, V.,

Malisauskas, M., Casaitec, V., Darinskasd, A., Forsgrene, L.,

Morozova-Roche L.A. (2007) Immune reactivity towards insulin, its

amyloid and protein S100B in blood sera of Parkinson’s disease

patients. European J. Neurology, 14, 327-334.

10

Introduction

Amyloid diseases and related proteins

Proteins, term originated from the Greek πρώτα ("prota")- "of primary

importance", were first described and named by Jöns Jakob Berzelius in 1838.

They are most multi-purposed macromolecules in living organisms,

participating in essentially all biological processes. They function as catalysts

and transporters; they store other molecules, provide mechanical support and

immune protection, transmit nerve impulses, control growth and differentiation

of cells and perform other functions. The structure of proteins and their ability

to carry out their correct functions are very tightly controlled, as even small

structural defects can lead to protein folding diseases. A number of human

diseases are associated with the deposition of protein aggregates. These include

prion diseases such as bovine spongiform encephalopathy and its human

equivalent Creutzfeld-Jakob disease, Alzheimer's and Parkinson’s diseases,

Skelefteå disease, type II (non-insulin dependent) diabetes and others. In most

of amyloid disorders, the protein deposits are accumulated extracellular. In

Parkinson's disease, however, amyloid deposits of α-synuclein were found

intracellularly.

The major constituent of the proteinaceous deposits are amyloid fibrils.

They can be constituted of different proteins or peptides, depending on their

location in the body. To date more than 30 proteins are known to be involved

in the human amyloid diseases (Table 1). In spite of their different primary

sequences and secondary and tertiary structures, when the individual

polypeptides are assembled into fibrils, the latter exhibit very similar structural

11

characteristics such as cross-β-sheet core with the β-strands arranged

perpendicular to the long axis of the fibre [Glenner, 1980; Blake & Serpell,

1996]. The common structural properties of the fibrils imply that the general

structural mechanism govern their formation and therefore the principles found

for one particular protein have a general application for understanding this

phenomenon as a whole.

Table 1. Human amyloid disease and associated proteins.

Disease Aggregating polypeptide Structure of polypeptide

Neurodegenerative diseases Alzheimer's disease Amyloid β peptide Natively unfolded Spongiform encephalopathies

Prion protein or fragments Natively unfolded (residues 1–120) and α-helical (residues 121–230)

Parkinson's disease α-Synuclein Natively unfolded Dementia with Lewy bodies

α-Synuclein Natively unfolded

Amyotrophic lateral sclerosis

Superoxide dismutase-1 All-β, Ig like

Huntington's disease Huntingtin with polyQ expansion

Largely natively unfolded

Spinocerebellar ataxia 17 TATA box-binding protein with polyQ expansion

α+β, TBP like (residues 159–339); unknown (residues 1–158)

Nonneuropathic systemic amyloidoses AL amyloidosis Immunoglobulin light chains

or fragments All-β, Ig like

AA amyloidosis Fragments of serum amyloid A protein

All-α, unknown fold

Senile systemic amyloidosis

Wild-type transthyretin All-β, prealbumin like

12

Disease Aggregating polypeptide Structure of polypeptide

Familial amyloidotic polyneuropathy

Mutants of transthyretin All-β, prealbumin like

Hemodialysis-related amyloidosis

β2-microglobulin All-β, Ig like

Apo amyloidosis N-terminal fragments of apolipoproteins

Natively unfolded

Lysozyme amyloidosis Mutants of lysozyme α+β, lysozyme fold Fibrinogen amyloidosis Variants of fibrinogen α-

chain Unknown

Nonneuropathic localized diseases Type II diabetes Amylin, also called islet

amyloid polypeptide (IAPP) Natively unfolded

Injection-localized amyloidosis

Insulin All-α

Corneal amylodosis associated with trichiasis

Lactoferrin α+β

Cataract γ-Crystallins All-β

Misfolding and aggregation

Generic properties of amyloids

Amyloid was discovered in 1854 by German physician Rudolph

Virchow, who named it in the belief that the iodine-staining component was

starch-like [Cohen & Jones, 1991; Sipe, 2000]. Positive response by iodine

staining led Virchow to the conclusion that the substance underlying the

evident macroscopic abnormality was cellulose and he gave it the name of

amyloid, derived from the Latin amylum and the Greek amylon [Cohen &

Jones, 1991]. In 1859 Friedreich and Kekule demonstrated that the amyloids

contain the depositions of proteinaceous mass. [Kyle, 2001].

The amyloid depositions were most commonly identified by various

types of microscopy and immunohistochemical dyes such as Congo red and

thioflavin T. (Figure 1 A, B)

A B C

Figure 1. Tinctorial properties of amyloid. Congo red staining of transthyretin (A) and thioflavin T staining of human lysozyme fibrils (B) and X-ray diffraction pattern of Ile56Thr (C) [Morozova-Roche et al., 2000; Saraiva, 2002].

Congo red stained deposits of amyloid were observed in a variety of

tissues using polarization light microscopy, as they demonstrate a characteristic

apple green birefringence under polarised light. Congophilia with apple green

birefringence was used as the first criterion of amyloid. [Missmahl & Hartwig,

1953] The secondary diazo dye Congo red is schematically presented in Figure

2. It has been suggested that it bindings specifically to amyloids via insertion

between the individual strands of cross-β-sheet [Carter & Chou, 1998]

Figure 2. Chemical structure of secondary diazo dye Congo red.

13

Benzothiazole dye thioflavin T is extensively applied in the histological

analysis of ex vivo amyloids as well as in in vitro amyloid studies. The

fluorescence of thioflavin T increases significantly upon binding to amyloid

fibrils. Krebs and co-workers suggested that thioflavin T molecules bind to

fibrils in the grooves on the face of the β-sheets that form the fibrillar backbone

[Krebs et al., 2005]. (Figure 3)

A B

Figure 3. Chemical structure of thioflavin T (A) and model of binding with amyloid structure (B) [Krebs et al., 2005].

Electron and atomic force microscopic studies of amyloids

[Chamberlain et al., 2000] showed that they exhibit a fibrillar morphology; the

fibrils range in width from 60 to 130 Å (average 75 to 100 Å) and in length up

to microns.

Figure 4. The structure of fibrils assembled from human lysozyme [Chamberlain et al., 2000]. Transmission electron microscopy (A); Atomic force microscopy (B). 14

X-ray diffraction analyses of isolated amyloid protein fibrils [Bonar et

al., 1969; Glenner et al., 1974; Sunde and Blake, 1998] showed specific

diffraction pattern, which was assigned to cross-β-sheet packing (Figure 1C).

The β-sheet formation accompanied the amyloid formation is also

monitored by using circular dichroism and Fourier-transformed infrared

spectroscopy (Figure 5). Up to date, all these methods are widely used in the

amyloid research.

A B

Figure 5. Fourier-transformed infrared spectrum of fibrillar amylin (A) and the far-UV circular dichorism spectrum (B) of fibrillar human lysozymes, demonstrating the β-sheet formation [Nilsson & Raleigh, 1999; Morozova-Roche et al., 2000].

During the last decade, a number of proteins not associated with

diseases have been found to convert into amyloid fibrils in vitro, showing the

physicochemical characteristics of those related to diseases (Table 2). The

similarity in dimensions and morphology, as well as in tinctorial properties

between ex-vivo and in vitro protein fibrils has been demonstrated. All fibrils

independent of their origin are stabilised by an array of inter- and

intramolecular hydrogen bonds between the main-chain amide and carbonyl

15

16

groups that are common to all peptides and proteins. Based on the accumulated

data, it has been suggested that the ability to form amyloid fibrils is an inherent

property of polypeptide chains and most if not all polypeptides have ability to

form amyloids, though their propensity to amyloid self-assembly can be

different (Dobson et al., 1999, 2001).

Table 2. Non-disease related amyloidogenic proteins

Protein References

Curlin Chapman et al., 2002

Hydrophobin EAS Wosten & Vocht, 2000

Myoglobin Fandrich et al., 2001

Proteins of the chorion of the eggshell Iconomidou et al., 2000

Spidroin Kenney et al., 2002

Ure2p (prion) Taylor et al., 1999

Sup35p (prion) King et al., 1997

HET-s (prion) Dos Reis et al., 2002

SH3 Guijarro et al., 1998

Cytochrome C Pertinhez et al., 2001

Equine lysozyme Malisauskas et al., 2003

α-lactalbumin Goers et al., 2002

Hen lysozyme Goda et al., 2000

Albebetin Morozova-Roche et al., 2004

Amyloid precursor state

17

In the amyloid field, particular attention was focused on amyloid

precursor state, which can be converted into fibrils. The multiple

conformational states of proteins are presented schematically in Figure 2 using

the energy folding landscape (Figure 6). Equilibrium partially folded states are

considered as stable counterparts of kinetic intermediates transiently populated

during protein folding kinetics [Heegaard et al., 2001]. Among equilibrium

partially folded states molten globule ones are the most extensively studied.

They are characterised by the native-like secondary structure, but labile and

large destabilised tertiary structure. Many globular proteins have been shown

to adopt several stable conformations, e.g. the native, molten globule, pre-

molten globule and unfolded states [Dunker et al., 2001; Kayed et al., 1999;

Lashuel et al., 2002; Merlini et al., 2003; Smith et al., 2003; Teplow et al.,

1998; Uversky et al., 2002]. A common observation is that fibrillization starts

from an intermediate state either partially unfolded or partially folded [Rochet

& Lansbury, 2000]. Many globular proteins such as human and hen lysozymes

[Morozova-Roche et al., 2000, Krebs et al., 2000], phosphoglycerate kinase

[Damaschun et al., 1999], cystatin C [Ekiel & Abrahamson, 1996],

acylphosphatase [Chiti et al., 1999] and transthyretin [Lai et al., 1996] and

other need to be partial unfolded in order to undergo fibril formation. In the

case of unfolded polypeptides such as α-synuclein [Uversky et al., 2000; 2001]

and islet amyloid polypeptide, the amyloid precursor intermediates have to be

partially folded. The free-energy barriers of the protein folding landscape

(Figure 3) can be rather high [Damaschun et al., 1999; Ferreira & De Felice,

2001; Kelly, 1999], leading to existence of parallel conformational states [Soto

& Saborio, 2001], including oligomeric, fibrillar and aggregated forms.

18

Conditions of amyloid formation

The amyloid-precursor state can be accumulated and the subsequent

aggregation can occur under variety of conditions, which have been studied

scrupulously [Ahmad et al., 2003; Canet et al., 2002; Hoyer et al., 2004; Lai, et

al., 1999; McParland et al., 2002; Smith et al., 2003]. The incubation at low

pH and elevated temperatures is widely used as a common condition to

produced fibrils from a number of proteins, including lysozymes (equine,

human and hen) [Malisauskas et al., 2003; Mishra et al., 2007; Morozova-

Roche et al., 2000], β2 microglobulin [Sasahara et al., 2007], insulin [Range et

al., 1997], transthyretin [Pasquato et al., 2007], apomyoglobin [Picotti et al.,

2007] and others.

These conditions cause the population of partially folded states in the

cases of lysozymes and β2 microglobulin. The low pH results in dissociation of

transthyretin tetramers and insulin hexamers into monomers or dimers,

respectively, which is critical for the amyloid assembly of these proteins

[Colon & Kelly, 1992; Ahmad et al., 2003].

Acylphosphatase and wild type hen lysozyme can be converted into

amyloid fibrils in the presence of moderate concentration of 2,2,2-

trifluoroethanol [Chiti et al., 1999; Krebs et al., 2000]. Under these conditions

hydrophobic residues and main-chain hydrogen bond donors and acceptors

become partially exposed, thus non-covalent intermolecular interactions and

particular hydrogen bonding within the polypeptide chain become favourable.

High hydrostatic pressure modulates protein – protein and protein –

solvent interactions through volume changes. Thereby it affects the equilibrium

between native and denatured conformational states as well as between the

monomeric, oligomeric and aggregated forms without the addition of

chemicals. High hydrostatic pressure has provided to be a useful tool to

populate and stabilize folding intermediates, which are characterised by more

perturbed and solvated structure as well as by smaller volume than the native

state [Randolph et al., 2002; Silva & Weber, 1993].

Figure 6. The energy folding landscape. The surface shows the multitude of conformations 'funneling' towards the native state via intramolecular contact formation, or towards the formation of amyloid fibrils via intermolecular contacts [Jahn & Radford, 2005].

Partially unfolded conformations with high propensity to form amyloid

have been populated by using this method for a range of proteins, including

insulin [Jansen et al., 2005], human lysozyme [De Feliche et al., 2004],

interferon-γ [Webb et al., 2001], interleukin-1 receptor antagonist [Seefeldt et

al., 2005], and β-lactoglobulin [Panick et al., 1999].

19

20

Amyloid fibrils can be induced by adding transient metal ions.

Miranker and co-workers demonstrated that Cu2+ significantly destabilizes β2

microglobulin and enhances fibrils formation of β2 microglobulin at neutral pH

[Morgan et al., 2001]. The addition of copper, iron, manganese or zinc ions

induced fibrillation of synthetic Aβ peptide [Huang et al., 1997; Atwood et al.,

1998]. Similar role of transient metal has been shown in the conversion of PrPC

to PrPSc [Jones et al., 2004]. In most of the cases the metal binding site of the

proteins involves histidine residues: in the case of Aβ peptide these are His6,

His13 and His14, while in PrP the binding site includes HGGGW segment of

the N-terminal domain. For β2 microglobulin different binding sites were

demonstrated for the native and amyloid precursor states; in the native state the

binding site involves His31, while in the non-native state the coordination of

Cu2+ is mediated by residues His13, His51 and His84, but not His31. [Eakin et

al., 2002].

The net charge of the proteins is critical for amyloid formation. The

kinetics and the quantity of the amyloid of albebetin, which carries a net charge

-12 under the neutral pH, was significantly enhanced by increasing ionic

strength, shielding the overall electrostatic repulsions between protein

molecules [Morozova-Roche et al., 2004]. Similarly, the fibrillation of

pharmaceutical peptides glucagon and insulin is also affected by both cations

and anions, changing the kinetics and morphology of amyloids by minimizing

the electrostatic interactions [Nielsen et al., 2001; Pedersen et al., 2006].

Chemical denaturants such as urea and guanidine-HCl at moderate

concentrations have been used extensively in protein-folding studies to trap

folding intermediates [Cymes et al., 1996; Ayed & Duckworth, 1999]. They

affect protein stability by disrupting hydrogen-bonding of the peptide

backbone, thus destabilizing secondary structure, and by allowing greater

solvation of hydrophobic side chains. For example, moderate urea

21

concentrations accelerates fibril formation of insulin [Ahmad et al., 2004] and

β-lactoglobulin [Hamada & Dobson, 2002; Rasmussen, 2007]. In prion protein

increasing urea concentration led at first to an increase in fibril formation

kinetics, which followed by a decrease in a rate of assembly, with the

maximum rate at 2.4 M urea, corresponding to the midpoint of protein

unfolding [Baskakov et al., 2004]. A bell-shaped dependence of the speed of

the fibrillation of β-lactoglobulin has been observed also by Hamada and co-

workers [Hamada & Dobson, 2002]. The authors showed that the rate of β-

lactoglobulin fibril formation decreased at the urea concentrations more than 5

M, although the population of unfolded protein molecules increased under this

condition. In order to aggregate both high population of unfolded proteins and

strong interactions between them are required. At the increasing concentration

of denaturant such as urea, the intermolecular interactions decline substantially

leading to decrease in amyloid assembly [Chiti et al., 1999; Bellotti et al.,

2000].

Finally for natively unfolded proteins such as α-synuclein, partial

folding has been shown to be essential first step in self-assembly [Der-

Sarkissian et al., 2003], underlining the importance of partially folded species

as amyloid precursors.

Effect of mutation on amyloidogenic properties

Most mutations associated with accelerated fibrillation in vitro and with

protein deposition in the body have been shown to destabilize the protein

native structure, increasing the concentration of partially folded conformers

[Rochet et al., 2000; Nielsen et al., 2001; Canet et al., 1999; Heegaard et al.,

22

2001]. In the case of human lysozyme, single point mutations in the gene cause

the fibril deposition in several tissues. Two amyloidogenic variants, Ile56Thr

and Asp67His, have been studied in detail and shown to be significantly less

stable than the wild-type protein [Morozova-Roche et al., 2000]. They showed

also a lack of cooperativity of the native structure, leading to an increased

concentration of partially folded states at equilibrium [Booth et al., 1997].

Transthyretin amyloidogenic variants are characterised by decreased stability

and increased dissociation rate constant of tetramer [Hammarstrom et al.,

2002]. Chiti and co-authors systematically analysed more than 50 mutants of

acylphosphatase. They demonstrated that destabilization of the native

conformation of acylphosphatase renders it more prone to amyloid formation.

The correlation drawn between the conformational stability of native

acylphosphatase and the tendency to form aggregates indicates that amyloid

fibrils originate from an ensemble of denatured conformations of the protein

under conditions in which the formation of non-covalent interactions is still

favourable [Chiti et al., 2000].

Nielsen and co-workers [Nielsen et al., 2001] have performed thorough

studies of amyloid properties of insulin mutations. They have demonstrated the

significant increase in the lag time of fibril formation of insulin upon amino

acid substitutions to more polar residues, as well as mutations to charged

residues. The authors suggested that mutations may affect the amount of

oligomeric intermediates formed under the conditions of experiments or may

perturb their structure, thus affecting the propensity to form fibrils. These

results demonstrate the importance of both hydrophobic and electrostatic

interactions in the initial stages of fibrillation [Nielsen et al., 2001].

Models of amyloid formation

The kinetics of fibril formation can be described as a nucleation

dependent process involving three major steps: nucleation, fibril growth and

stationary phases [Jarrett & Lansbury, 1993]. Nucleation involves the assembly

of several protein monomers into an organized structure – nucleus, serving the

role of precursor for the formation of fibrils. A lag phase of fibrillation kinetics

is related to the time required for the formation of nuclei. Subsequent addition

of monomers to the nuclei elongates them into fibrils [Teplow, 1998]. Seeding

or addition of preformed nuclei diminishes the lag-phase (Figure 7) [Han et al.,

1995; Wood et al., 1999; Morozova-Roche et al., 2000].

Figure 7. Aggregation of hen lysozyme. Data are shown for the aggregation of 1 mM hen lysozyme at pH 2.0 and 65°C monitored by thioflavin-T fluorescence in the absence of added fibrils (down triangles) and following addition of 5% (v/v) (black squares), 2.5% (v/v) (circles), 1% (v/v) (triangles), and 0.5% (v/v) fibrils preformed from hen lysozyme at pH 2.0 (diamonds). Seeding with a 5% (v/v) aliquot of a solution of amyloid fibrils containing purified full-length hen lysozyme gave essentially identical kinetics (gray squares) to aliquots formed simply by incubation at pH 2.0 and 65°C [Krebs et al., 2004].

23

24

There are several models of fibril formation based on the kinetic studies

such as monomer-derived conversion [Prusiner, 1982], template assisted [Soto,

2001] and nucleated polymerization model [Lomakin et al., 1996]. The

monomer derived conversion model suggests that a monomeric peptide can

adopt a conformation called the A state that is analogous to the conformation

adopted in the fibril. The rate-determining step is the A and S monomer

binding and converting S state monomer, which results in the A state dimer

(Figure 8 A). The dimer then dissociates and the A structured monomers

rapidly add to the end of the growing fibril [Prusiner, 1982]. In the template-

assisted model, a preassembled nucleus binds to a soluble state peptide present

in a random coil conformation. This step is followed by a rate-determining

structural change to add the peptide to the growing end of the fibril or filament,

presumably as part of the β-sheet-rich quaternary structure (Figure 8B) [Soto,

2001]. The nucleation polymerisation is characterized by the rate-limiting

formation of a nucleus arising from equilibrium between monomers. Once a

nucleus is established, assembly occurs by the addition of assembly competent

monomers to the growing end of the fibril (the nucleus). [Lomakin et al., 1996]

Serio and co-workers proposed the nucleation conformational conversion

model, which incorporated features of nucleated polymerization and template-

assisted ones. [Serio et al., 2000].

Figure 8. The models of protein conversion into amyloid fibrils. A, Templated assembly; B, monomer-directed conversion; C, nucleated polymerisation; D, nucleated conformational conversion. A-state denoted by smooth circles, rough circles represent S state [Kelly, 2000].

This model suggests that native oligomers accumulate and associate into a

“nucleus”, where conformational conversion takes place as a rate-determining

step. Oligomeric complexes are crucial intermediates forming the amyloid

nucleus. When these complexes undergo a conformational change on

association with the nuclei, rapid fibrillar assembly follows [Serio et al., 2000].

Each particular amyloid scenario need to be carefully investigated to conclude

which specific mechanisms is involved.

25

26

Cytotoxicity of amyloids

Amyloid toxic species

In spite of the large amount of experimental data on the role of protein

self-assembly in amyloid related diseases, the nature of the toxic amyloid

structures remains elusive. A number of studies have supported the view that

the prefibrillar aggregates are the most cytotoxic species. By means of electron

and atomic force microscopy, size exclusion chromatography and other

techniques it has been demonstrated that the broad range of structures

characterized by a different degree of assembly of misfolded proteins and

variable morphology are toxic in vitro. These include ADDLs – β-amyloid-

derived diffusible ligands [Walsh et al., 2002, Gong, et al., 2003], 8-mers to

20-mers of equine lysozyme [Malisauskas et al., 2005], 20-mer cross-β-sheet

containing oligomers of albebetin [Zamotin et al., 2006], misfolded monomers

and hexamers of transthyretin [Reixach et al., 2004], 8–9 nm globular

structures of transthyretin mutants [Andersson et al., 2002], 4– 200 nm

diameter granules of the PI3-SH3 domain, 4 – 8 nm width aggregates of HypF

[Bucciantini et al., 2002], 3 – 5 nm spherical aggregates of α-synuclein [Ding

et al., 2002; Volles et al., 2001] and others. Kayed and co-authors reported that

intermediate-size soluble oligomers of Aβ(1–40) and five other proteins form

structures sharing a similar conformational epitope, suggesting that they exert

cytotoxicity via similar mechanisms [ Kayed et al., 2003]. Even the process of

nucleation-dependent polymerization of Aβ(1–40) itself has been suggested as

a primary cause of cytotoxicity [Wogulis et al., 2005].

27

A greater controversy exists on the cellular toxicity of the amyloid

fibrils. Initially, they have been considered as an inert material substantially

accumulated in the tissues and organs [Andersson et al., 2002; Anderluh et al.,

2005; Bhatia et al., 2000; Chromy et al., 2003; Sirangelo et al., 2004]. There is

also a body of experimental evidence clearly demonstrating the neurotoxicicity

and cytotoxicity of amyloid fibrils.[ Forloni 1996; Forloni et al., 1996; Howlett

et al., 1995; Lorenzo& Yankner, 1994; Novitskaya et al., 2006; Weldon et al.,

1998; ] Amongst the recent findings, Petkova and co-authors showed that

mature fibrils from Aβ(1–40) produced under two different conditions, and

consequently characterized by different morphologies, exert significantly

different toxicity on neuronal cells [Petkova et al., 2005]. The origin of the

toxicity of amyloid species, especially in the body, require further

investigation.

Cellular targets for amyloids

The similarity and inherent toxicity of the amyloid structures suggests

that the toxic agents may share specific cellular targets [Kayed et al., 2003].

Cellular membranes, separating the ordered living cells from the entropic chaos

outside, are considered to be such common target [Hartley, 1999; Janson et al.,

1999]. Prefibrillar oligomers of α-synuclein and Aβ peptide were shown to lead

to neuron dysfunction and degeneration by forming the structures with pore-

like morphologies, which can be inserted into the cellular membranes, altering

their permeability [Lashuel & Lansbury, 2006]. The formation of amyloid

pores in mitochondria results in the loss of mitochondria membrane potential,

overloading with Ca2+ and subsequently to oxidative stress [De Giorgi et al.,

28

2002]. Some studies suggest that lipids can be absorbed by amyloids during

their formation on the membranes, leading to the membrane rupture due to

gradual lipid depletion [Sparr et al., 2004; Knight & Miranker, 2004].

Another main target involved in amyloid cytotoxicity is mitochondria

which can be affected by cytotoxic species in direct or indirect manner [Eckert,

2003; Brookes et al., 2004; Swerdlow et al., 2004; Takuma et al., 2005;

Nunomura et al., 2006]. Mitochondria are essential for the regulation of

intracellular Ca2+ homeostasis and as the principal generators of intracellular

reactive oxygen species. Therefore, the defects of mitochondrial functions can

result in excessive production of reactive oxygen species [Butterfield et al.,

2001; Wyttenbach et al., 2002]. The formation of the transmembrane pores are

linked to release of small proteins, such as cytochrome C and apoptosis-

inducing factor [Crompton, 1999; Rodriguez-Enriquez et al., 2004]. The

mitochondrial dysfunctions were proposed as potential mechanisms in the

development and pathogenesis of Alzheimer’s disease and most of the studies

was focused on the effect of Aβ peptide on mitochondrial activity [Ojaimi &

Byrne, 2001]. Growing evidence indicates that cell death can be initiated by the

Aβ peptide induced generation of free radicals [Butterfield et al., 1994],

leading to lipid peroxidation and oxidative stress. Oxidative stress,

subsequently, induces intracellular accumulation of Aβ [Misonou et al., 2000].

The experiments performed on isolated mitochondria demonstrated the ability

of Aβ to be directly toxic to the organelle [Casley et al., 2002]. It has been

shown that Aβ inhibits the mitochondrial electron transport chain complexes,

cytochrome C oxidase and promotes Ca2+-induced assembly of the

transmembrane pores. It has been also demonstrated that Aβ is present in

mitochondria, presumably in association with Aβ-binding alcohol

dehydrogenase, and promotes mitochondrial dysfunction and reactive oxygen

29

species production. This provides a direct link to Aβ-induced mitochondrial

toxicity [Lustbader, et al., 2004]. Recently it has been shown that Aβ induces

neuronal apoptosis via mechanisms involving induction of Fas ligand and the

release of Smac via AP-1/Bim activation in mitochondria [Morishima et al.,

2001; Yin et al., 2002]. In the case of Parkinson’s disease, it was shown, that

fibrillar deposits of α-synuclein also cause oxidative stress and mitochondria

dysfunction [Cooper et al., 2006].

Another abnormality found in neurodegenerative diseases is linked to

the damage of endoplasmatic reticulum. Endoplasmatic reticulum stress results

in the disruption of Ca2+ homeostasis, inhibition of protein N-linked

glycosylation, expression of mutant proteins and other types of cellular stresses

caused by the accumulation of misfolded proteins in the lumen of

endoplasmatic reticulum [Kaufman, 1999; Katayama et al., 2004]. It has been

suggested that the endoplasmatic reticulum stress-mediated apoptotic signal

pathways, specifically activation of caspase-4, may also contribute to the

pathogenesis of Alzheimer’s disease [Katayama et al., 2004]. As Aβ is

synthesized and accumulated in the endoplasmatic reticulum, Aβ has been

suggested as a direct mediator of the endoplasmatic reticulum stress responses

and apoptosis. Aβ(1–42) activates caspase-12 in primary neurons through

calpain activation, therefore, the caspase-12 knocked-out neurons are partially

resistant to Aβ-induced cell death [Nakagawa et al., 2000].

In familial amyloid polyneuropathies, the interaction of transthyretin

fibrils with RAGE receptors (receptor for advanced glycation end products)

were suggested as contributing to cellular stress and toxicity [Sousa et al.,

2001]. These indicate that fibrils can act via specific mechanisms, rather than

inducing only accidental cell damage.

30

General pathways of cell death

The general pathways of cell death induced by amyloids are also a

subject of intense debates. In the studies described above it has been shown

that amyloids activate caspases and receptor-mediated signaling pathways,

associated with apoptosis. There are, however, findings that HypF-N amyloid

exerts necrotic rather than apoptotic death of NIH-3T3 fibroblasts [Bucciantini

et al., 2002]. Here we give the definitions of apoptotic and necrotic cell deaths,

which can be useful for our further analysis. Apoptosis, or programmed cell

death, is a highly regulated process that allows a cell to self-degrade in order to

eliminate unwanted or dysfunctional cells in the body. During apoptosis, the

genome of the cell fractures, the cell shrinks and part of it disintegrates into

smaller apoptotic bodies (Figure 9). Unlike necrosis, where the cell dies by

swelling and bursting its content in the surrounding area, which causes an

inflammatory response, apoptosis is a very clean and controlled process.

During apoptosis the content of the cell is kept strictly within the cell

membrane during its degradation. The apoptotic cell is phagocytosed by

macrophages before the cell’s contents have a chance to leak into the

neighborhood [Van Cruchten & Van Den Broeck W, 2002]. Therefore,

apoptosis can prevent unnecessary inflammatory response.

Apoptosis can be triggered in a cell through either the extrinsic or

intrinsic pathways. The extrinsic pathway is initiated through the stimulation of

the transmembrane death receptors, such as the Fas or TNF receptors, located

on the cell membrane. In contrast, the intrinsic pathway is initiated through the

release of signal factors by mitochondria within the cell (Figure 10).

Figure 9. Hallmarks of the apoptotic and necrotic cell death process [Van Cruchten & Van Den Broeck W, 2002].

The caspase pathway leading to classical apoptosis [Mattson, 2000;

Strasser et al., 2000] is initiated by the release of cytochrome C from the

mitochondrial intermembrane space. Together with other essential factors, such

as ATP, it triggers assembly of the apotosome complex, which forms the

template for efficient caspase processing [Hengartner, 2000; Strasser et al.,

2000;]. The second mitochondrial death pathway leads to necrosis, without

necessarily activating caspases. Intracellular control of this pathway can be

attenuated by antioxidants [Schulze-Osthoff et al., 1992; Vercammen et al.,

1998]. The third distinct mitochondria pathway involves the release of the

apoptosis-inducing factor from the intermembrane space [Susin et al., 1999]. It

induces caspase-independent formation of large (50 kb) chromatin fragments

[Strasser et al., 2000; Susin et al., 1999]. The biochemical difference between

various cell death pathways is reflected in morphology of the condensed

chromatin (Figure 10). Often, a few pathways are activated simultaneously 31

[Jäättelä et al., 1998; Mattson, 2000; Susin et al., 1999]. The cell fate is then

determined by the relative speed of each process in a given model. The

amyloids can also activate the extrinsic apoptotic pathways, which are

followed by intrinsic pathways switching between apoptosis and necrosis,

depending on the timing and severity of mitochondria derangement.

Figure 10. Mitochondrial regulated pathways of cell death. A cell may switch back and forth between different death pathways [Leist & Jaattela, 2001].

32

33

Studied proteins

Lysozymes

Lysozyme is a protective enzyme which lyses bacteria by catalyzing the

hydrolysis of mucopolysaccharides of bacterial cell walls. Human lysozyme is

found in a variety of organs, including spleen, lung and kidney as well as in

extracellular fluids such as plasma, milk, saliva and tears. Five pathogenic

mutants of human lysozyme (I56T, F57I, W64R, D67H, W112R) are found to

be associated with non-neuropathic systemic amyloidosis, where mutant

lysozymes are deposited in kidney, lung, liver and intestines [Pepys et al.,

1993; Röcken et al., 2006; Valleix et al., 2002; Yazaki et al., 2003]. The

effects of several amyloidogenic mutations on the folding properties of

lysozyme have been investigated [Booth et al., 1997; Canet et al., 1999;

Merlini and Bellotti, 2005]. All mutants show the lower stability of the native

fold, which enables them to enter partially folded molten globule states under

physiological conditions. The molten globule state is also populated for the

wild-type lysozyme, but under more severe denaturing conditions, which may

explain why wild-type lysozyme is not found in amyloid deposits.

Hen egg white and equine lysozymes belong to the same family of c-

type lysozymes and were extensively studied during the last decades [Krishna

& Englander, 2007; Matagne & Dobson, 1998; Permyakov et al., 2006;

Mizuguchi et al., 1998; Morozova-Roche et al., 1997; Morozova et al., 1995].

In contrast to most c-type lysozymes, which do not bind calcium, equine

lysozyme is a calcium-binding protein. The calcium-binding properties

affected its structural organisation, stability and subsequently folding

34

properties. Equine lysozyme displays about 50 % sequence homology to

human and hen lysozymes, similar degree of the amino acid sequence

homology exists between the members of lysozyme family and other group of

proteins – α-lactalbumins. Lysozymes and α-lactalbumins show remarkable

homology in their tertiary and secondary structures. Equine lysozyme occupies

a special position within this family, combining the structural and folding

properties of non-calcium-binding lysozymes and calcium-binding α-

lactalbumins. Equine lysozyme was suggested to serve a role of an

evolutionary link between these groups of proteins [Nitta et al., 1987; Acharya

et al., 1994]. Specifically, it contains the active-site residues Glu35 and Asp52,

involved in lysozyme enzymatic activity [Tsuge et al., 1991; 1992], and the

conserved, high-affinity calcium-binding site of α-lactalbumins (Figure 11)

[Permyakov and Berliner, 2000].

As shown schematically in Figure 11, lysozymes and α-lactalbumins consist of

two sub-domains: the α-domain is comprised by four major α-helices, while the

β-domain contains a triple-stranded β-sheet and several long loops. All proteins

of this super-family possess also highly conserved four disulfide bridges. The

equine lysozyme calcium-binding property was initially predicted by Stuart

and co-workers [Stuart et al., 1986], since in the domain interface it has the

same calcium-binding ligands as α-lactalbumins (Figure 11). Then the binding

constant of 2×106 M-1 was determined experimentally [Nitta et al., 1987] and

the structure of the binding site confirmed by crystallographic and NMR

structural analysis [Acharya et al., 1994; Tsuge et al., 1991; 1992]. The

presence of the calcium-binding site most likely contributes to equine

lysozyme’s lower stability and cooperativity compared to other lysozymes

[Morozova et al., 1991, 1995; Morozova-Roche et al., 1999; Van Dael et al.,

1993]. If hen and human lysozymes undergo highly cooperative two-state

35

denaturation above 70 °C, and even the amyloidogenic variants of human

lysozyme are destabilised by only ca. 12 °C [ Privalov, 1979; Booth et al.,

1997], then by contrast, apo-equine lysozyme displays a three-state transition

in a wide temperature range from room temperature to 80 °C (Figure 12)

[Griko et al., 1995; Permyakov et al., 2006]. Calcium-binding significantly

stabilises the protein structure, shifting the first transition, which depend on the

calcium concentration, towards higher temperatures. However, even in the

presence of high calcium concentration the thermal denaturation of equine

lysozyme cannot be described by two-state, but three-state model.

Figure 11. Equine lysozyme – an evolutional link between lysozymes and calcium-binding α-lactalbumins [Blanch et al., 2000].

Equine lysozyme enters into various molten globule states during the

thermal, chemical or pH denaturation [Morozova et al., 1991; 1995;

Morozova-Roche et al., 1999; Van Dael et al., 1993]. Amide-proton hydrogen

exchange analysis provided residue-specific information on the nature of

equine lysozyme molten globule stability [Morozova et al., 1995]. The acidic

A-state of equine lysozyme shows high protection of the A, B and D α-helices

and particularly the amides of hydrophobic residues facing the helical interface

and making contacts with each other. The model of molten globule was

proposed, which possesses an extended hydrophobic core based on the native-

like tertiary interactions of the three out of four major α-helices. This was the

most stable molten globule within lysozyme-α-lactalbumin family [Morozova

et al., 1995, Morozova-Roche et al., 1997]. Thus, the first thermodynamic

transition of equine lysozyme corresponds to the conversion of the native state

to molten globule, while the unfolding of the core, rather than the whole α-

domain [Griko et al., 1995], gives an enthalpic contribution to the second

transition from molten globule to thermally-unfolded state [Kobashigawa et al.,

2000; Koshiba et al., 2000; Morozova et al., 1995; Van Dael et al., 1993].

Figure 12. Specific heat capacity of equine lysozyme at various calcium concentrations [Permyakov et al., 2006].

Hen lysozyme also populates an equilibrium partially folded state under

strongly acidic conditions as well as a kinetic intermediate [Kuwajima et al.,

1985; Sasahara et al., 2000]. In our research we have explored the partially

folded states of equine lysozyme and destabilised conformation of hen

lysozyme as amyloid precursors to produce various forms of amyloids

described below. 36

De novo designed protein albebetin and its constructs

De novo designed albebetin (ABB) is a 7.4 kDa protein, containing 73

amino acid residues in its sequence. ABB is characterised by a double repeat of

αββ-motif [Dolgikh et al., 1991; Fedorov et al., 1992], arranged in such a way

that four β-strands form an anti-parallel β-sheet covered by α-helices (Figure

13).

The compactness of ABB was optimised by the short loops connecting

the elements of the secondary structure and limiting the number of

conformations which can be adopted. While ABB possesses a well defined

secondary structure, it is characterised by a labile tertiary structure, which

correlates with its low immunogenicity [Bocharova et al., 2002].

Figure 13. Schematic representation of the ABB molecule [Koradi et al., 1996]. In order to increase the solubility of ABB, 22 charged amino acid residues

were introduced throughout the protein primary structure, including 17 Glu and

37

Asp residues and 5 Arg and Lys residues, creating a high net charge of -12 at

the neutral pH (Figure 14).

The large electrostatic repulsion contributes to an overall instability of

ABB. As a result, albebetin is characterized by the conformational mobility of

molten globule type at neutral pH and room temperature. While the molten

globules are most commonly induced by the additional perturbations or

destabilising conditions, neither of them are required in the case of albebetin. It

exists in the molten globule state under physiological conditions by definition

of its design.

N-tgEnhR-

N-lKEKKysp- spED lggN- mDpgDpDc tgg lEqllRR

α helix

tvhvEv β strand

vEvEv β strand Fused

Albebetin

lggs tgg spEDR-C EclEqllRR α helix

pgDp vEvEv β strand

tvhvEv β strand

Figure 14. Primary structure of ABB. Peptides of interferon-α2 or HLDF-6 fused to ABB are shown as arrows. Positively charged amino acids are denoted by italic capital letters. Negatively charged amino acids are denoted with bold capital letters [Morozova-Roche et al., 2004]

Due to the low immunogenicity of albebetin, it was planned to be used

as a carrier protein in drug delivery. The octapeptide LKEKKYSP of human

interferon-α2 or the hexapeptide TGENHR of human leukemia differentiation

factor were fused to the N-terminus of albebetin [Aphasizheva et al., 1998],

which resulted in two biologically active constructs possessing the

functionality of the fused peptides. The first construct, ABB-I, activates the

thymocyte blast transformation similarly to interferon-α2 [Chertkova et al.,

2003], while the second, ABB-DF, induces the differentiation and inhibits

proliferation of human leukemia cells similarly to molecules of differentiation

38

39

factor [Dolgikh et al., 1996; Kostanyan et al., 1994]. These peptides do not

perturb the structure of albebetin; both constructs retain the molten globule

state. Due to importance of the molten globule in amyloid scenario as an

amyloid precursor state, the amyloidogenic propensities of albebetin and its

two constructs were analysed in our research.

Application of atomic force microscopy in amyloid studies

Atomic force microscopy (AFM) allows two and three-dimensional

imaging and measurements of unstained and uncoated biological samples in air

or fluid. AFM belongs to an extended family of scanning probe microscopy

methods. AFM was first described by Binnig, Quate and Gerber in 1986 and

since then the method is rapidly developing.

AFM measurements are based on the interaction of the very sharp tip-

probe with the atomically flat surfaces on which the object of interest is

located. Images are formed by recording the interaction forces between the

probe and the surface as the cantilever carrying the probe is scanned over the

sample row by row. An optical system is used to measure the changes of the

reflection of the laser beam from the gold-coated back of the cantilever onto a

position-sensitive photodiode. This allows registering the changes in the

position of the incident laser at sub-nanometer scale. AFM operates in several

modes that can be chosen depending on the sample, environment and

measurements required. The contact mode is the original AFM imaging mode,

which can be implemented in both air and fluid. In this mode, the AFM probe

is brought into contact with the sample surface and raster-scanned across the

surface. Changes in the cantilever deflection during scanning are monitored

40

using an electronic feedback circuit. The soft biological samples, though, can

be drugged and damaged applying this mode.

The tapping mode is most commonly used in biological sample

imaging, In this mode the imaging probe is vertically oscillated at or near the

resonant frequency of the integrated cantilever. The probe-sample interactions

reflected in the changes of the amplitude and phase of oscillation. These

changes are transformed into the height or topographical image.

Simultaneously, amplitude and phase images can be produced. The advantage

of the tapping imaging that all these images can be obtained in a single scan.

Because the interactions between the tip and the surface depend not only on the

topography of the sample, but also on other characteristics, such as hardness,

elasticity, adhesion or friction, the movement of the cantilever depends also on

these properties, which are reflected in the phase shift during the tapping mode

scan. Therefore, the phase imaging can detect, for example, different

components in polymers related to their stiffness or areas of different

hydrophobicity in hydrogels immersed in saline solutions.

There are two major approaches to produce the cantilever oscillations in

the tapping mode. This can be achieved by acoustic vibrations induced by a

piezoelectric transducer attached to the cantilever holder. This method is called

an acoustic mode. Recently another approach becomes increasingly popular, in

which the cantilever is caused to oscillate directly without involving the

cantilever housing. The cantilever coated with a magnetic material is driven

into oscillation by magnetic field generated by a solenoid positioned close to

the cantilever housing. This mode is called magnetic mode. The application of

the magnetic mode has particular advantages during imaging in liquid, due to

minimising the noise produced by the system. The resolution of AFM in height

measurements both in air and in liquid environment can be of ± 0.1 nm.

Schematic diagram of AFM is presented in Figure 15.

41

AFM volume measurements

AFM can be used to determine the molecular volumes of individual particles

including single protein molecules as it was suggested by Schneider and co-

workers [Schneider et al., 1998]. We used this method to identify the volumes

of amyloid oligomers and consequently their stoichiometries. For this purpose

we measured the parameters of individual particles.

Figure 15. Schematic diagram of AFM [www.molec.com].

AFM usually overestimates the diameter of particles due to the

geometry of the tip, which introduces the broadening effect in the image. In

order to compensate this we measured the height and the diameter at half-

maximal height of the individual particles. Their volumes were calculated by

using the model of sphere segment which is presented in the Figure 16 and

equation 1.

Figure 16. Spherical cup model.

The equitation 1:

VAFM = (πh/6)×(3r2 + h2) (1)

Here, h is the particle height and r is the radius at half-height.

The molecular volume of monomeric protein was estimated by equation 2:

Vm = (M0/N0)×(V1 + dV2), (2) where M0 is the protein molecular weight, N0 is Avogadro’s number, d is the

extent of protein hydration (0.4 mol of H2O/mol of protein), V1 and V2 are the 42

43

partial specific volumes of individual protein (0.74 cm3 g -1) and water (1 cm3

g-1), respectively.

The number of molecules in oligomeric species was determined by the

equation 3.

n= VAFM / Vm, (3) where VAFM is the volume calculated from AFM measurements, Vm is volume of the single molecule.

44

Introduction to the articles

The results of this thesis are important for general understating the

mechanisms of amyloid formation and amyloid induced cytotoxicity. In this

section the main results and conclusions of the four articles, constituting the

thesis, are reviewed. The figures are referred to their numbers as given in the

original articles.

Paper I. Amyloid protofilaments from the calcium-binding protein equine

lysozyme: formation of ring and linear structures depends on pH and

metal ion concentration.

The destabilising mutations in human lysozyme are associated with the

hereditary systemic amyloidoses. Equine lysozyme is a naturally destabilised

protein, which is characterised by the formation of a range of molten globule

states. Equine lysozyme is considered as an evolutionary link between c-type

lysozymes and α-lactalbumins, as it performs the bacterial lyses similar to

lysozymes but binds calcium like α-lactalbumins. Here we have explored the

conditions, under which equine lysozyme populates molten globule state, in

order to produce amyloids. We have found that it self-assembled into

protofilaments of amyloid type under both pH 2.0 and pH 4.5 and elevated

temperatures. The morphology of its amyloid structures strongly depends on

environmental conditions such as calcium concentration, pH and temperature.

Fibrillar structures were characterised by using atomic force microscopy,

biochemical and spectroscopic methods. We demonstrated that the calcium-

bound equine lysozyme assembles into linear protofilaments, while the apo-

protein forms ring-shaped amyloids (Figure1, 2). The amyloid-rings are

45

characterised by diameters 45 - 50 nm at pH 4.5 and 70 - 80 nm at pH 2.0,

respectively. They bind amyloid dyes Congo red and thioflavin T similar to the

disease-related ring structures found for Aβ peptide and α-synuclein, which

were considered as a major pathogenic cause in Alzheimer’s and Parkinson’s

diseases. We suggested that the formation of annular amyloids is not limited to

disease-related polypeptides, but the amyloid-rings represent a second generic

type of amyloids in addition to the more common linear filaments.

We found that during the exposure to the low pH and elevated

temperatures equine lysozyme undergoes partial acidic hydrolysis, We have

shown that under appropriate conditions and limited exposure to the acidic pH

the full length equine lysozyme is preserved and gives rise to amyloid. This

indicates that a full length protein is able to assemble into amyloid structures

and partially hydrolysis is not required for this process. Upon further exposure

to acid pH two major peptides of 1 - 80 and 54 - 125 amino acid residues from

the primary structure of equine lysozyme were accumulated as shown by using

mass-spectrometry and gel electrophoresis. They are both prone to amyloid

formation, which indicates that the acidic hydrolysis can contribute to

amyloidogenic process on a longer time scale.

We have shown that equine lysozyme amyloid protofilaments do not

undergo maturation and significant growth. This was attributed to the presence

of the very stable core in its structure. Although the amino acid sequences most

likely forming the fibrillar cross-β-sheet correspond to the equine lysozyme β-

domain, the stable α-helical core forms a non-adhesive fibrillar interface,

preventing further fibrillar growth. These results suggest that it is not only the

propensity to form β-sheet structure governs the amyloid formation, but also

the conformational plasticity of the sequences involved in the fibrillar

interface. Therefore, the design of the regions with a high local stability and

46

cooperatively, in addition to the overall stabilisation of protein molecules, can

be explored in amyloid-preventing strategies.

Paper II. Fibrillation of carrier protein albebetin and its biologically active

constructs. Multiple oligomeric intermediates and pathways.

Albebetin was designed to adopt a molten globule conformation

under neutral pH. It is comprised of two αββ repeats, the compactness of

albebetin is achieved by the introducing short linking loops between the

elements of its secondary structure. Albebetin is characterised by a

conformational mobility of molten globule and rather low immunogenicity.

Albebetin was intended to be used as a drug carrier protein. For these purposes

the constructs of ABB were created by grafting to the N-terminus octapeptide

peptide from human intereferon-α2 and hexapeptide from human leukemia

differentiation factor. Both constructs possess the activity of the grafted

peptides, the first one, ABB-I, efficiently activates the thymocyte blast

transformation, while the second one, ABB-DF induces differentiation and

inhibits proliferation of human leukaemia cells.

We have demonstrated that albebetin and its constructs are highly

amyloidogenic at physiological conditions and form a wide range of amyloids

after hours and days of incubation under physiological pH. The combination of

methods including atomic force microscopy, Congo red and thioflavin T

binding and circular dichroism were used to follow development the amyloid

fibrils and their morphology. We have shown that the fibrillation proceeds via

multiple pathways and includes a hierarchy of amyloid structures ranging from

oligomers to protofilaments and fibrils. Comparative height and volume

microscopic measurements allowed us to identify two distinct types of

oligomeric intermediates: pivotal oligomers 1.2 nm in height comprised of 10 -

47

12 monomers and on-pathway amyloid-competent oligomers ca. 2 nm in height

constituted of 26 - 30 molecules. The pivotal oligomers can assemble into

chains and rings with “bead-on-string morphology”, in which a “bead”

corresponds to an individual oligomer. The rings and chain assemblies of

oligomers are very stable and, once formed, they remain in solution

simultaneously with fibrils. The amyloid-competent oligomers give rise to

protofilaments and fibrils, and their formation is concomitant with an

increasing level of thioflavin T binding. The amyloid nature of filamentous

structures was also confirmed by a pronounced thioflavine T and Congo red

binding as well as characteristic β-sheet-rich far-UV circular dichroism

spectrum. We suggest that transformation of the pivotal oligomers into the

amyloid-prone ones is a limiting stage in amyloid assembly. Peptides, either

fused to albebetin or added into solution, and an increased ionic strength

promote fibrillation of albebetin, carrying a net charge of -12 at the neutral pH,

by counterbalancing critical electrostatic repulsions. This finding demonstrates

that the fibrillation of newly designed polypeptide-based products can produce

multimeric amyloid species with a potentially “new” functionality, raising

questions about their potential safety in biomedical applications.

Paper III. Cytotoxicity of albebetin oligomers depends on cross-β-sheet

formation.

The cytotoxicity of amyloid prefibrillar and fibrillar species is a focus

of much current research. It is viewed as a common amyloid characteristic,

which the polypeptides involved in amyloid formation do not possess prior

their assembly into amyloids. The advantage of using albebetin in the

cytotoxicity studies is related to the fact that it forms amyloid under

physiological conditions and therefore its amyloid structures will not be

48

perturbed, when subjected to the conditions of cytotoxicity experiments.

Albebetin populated also well-defined and distinct amyloid oligomers of two

types, as described above, and protofilaments. Therefore we were able to

explore the cytotoxicity of two types of amyloid oligomeric assemblies, with

and without cross-β-sheet. We have shown that the initial or pivotal oligomers

consisted of 10–15 molecules as determined by atomic force microscopy, do

not bind thioflavin-T and do not affect viability of granular neurons and SH-

SY5Y cells. When these oligomers were converted into the larger ones

possessing ca. 30 - 40 albebetin molecules and developing cross-β-sheet

structure, they induce the reduction of the viability of both types of neuronal

cells. Neither monomers nor protofilaments of albebetin displayed cellular

toxicity on both neuronal cells. We have suggested that oligomeric size is

important for stabilizing cross-β-sheet core. The stable core may prevent

amyloid oligomers from the dissociation upon the penetration through cellular

membranes as well as inside of the cellular environment. The cytotoxicity of de

novo protein albebetin accentuates the universality of this phenomenon. As

albebetin was designed and intended to use as a carrier-protein for drug

delivery, its cytotoxicity once more necessitates the examination of

amyloidogenic properties prior subjecting the proteinaceous compounds to the

large scale applications.

Paper IV. Lysozyme amyloid oligomers and fibrils induce cellular death

via different apoptotic/necrotic pathways.

Amyloid cytotoxicity was further examined using hen lysozyme as a

model protein. Among the newly gained amyloid properties, its cytotoxicity

plays a key role. Lysozyme is a ubiquitous protein which has been shown to be

49

involved in systemic amyloidoses in vivo and able to form amyloid under

destabilizing conditions in vitro. We characterized both oligomers and fibrils of

hen lysozyme by atomic force microscopy and spectroscopic technique. The

well-defined amyloid species were subjected to the cellular toxicity assays after

different periods of aging. We demonstrated that both oligomers and fibrils of

hen lysozyme induce a dose (5 - 50 μM) and time-dependent (6 - 48 h) effect

on the viability of neuroblastoma SH-SY5Y cells. Using a wide range of cell

toxicity assays to probe the apoptotic and necrotic cell death pathways, we

have demonstrated that the oligomers and fibrils act on a different time scale.

We revealed that fibrils induce a decrease of cell viability after 6 h due to

membrane damage shown by inhibition of tetrazole reduction (WST-1 assay),

lactate dehydrogenase release and propidium iodide intake. By contrast, the

amyloid oligomers activate caspases after 6 h, but cause the cell viability

decline only after 48 h, as shown by fluorescent-labelled annexin V binding to

externalized phosphatidylserine, propidium iodide DNA staining, lactate

dehydrogenase release, and by typical apoptotic shrinking of cells. We

concluded that amyloid oligomers induce apoptosis-like cell death, while the

fibrils lead to necrosis-like death. As polymorphism is a common property of

an amyloid, we demonstrated that it is not a single uniform species, but rather a

continuum of cross-β-sheet-containing amyloids that are cytotoxic. An

abundance of lysozyme once more illuminates a universality of the amyloid-

associated toxicity, and accentuates that amyloid toxicity should be assessed in

all clinical applications involving proteinaceous materials.

50

Conclusions

• We have demonstrated that equine lysozyme forms amyloids under

conditions in which its molten globule states are populated. The

morphology of equine lysozyme amyloids is strongly dependent on

environmental conditions such as calcium concentration, pH and

temperature.

• We suggested that the ring-shaped amyloids represent an alternative

generic form of amyloid assembly in addition to most commonly

populated linear protofilaments. The ring formation is not limited to

Alzheimer’s and Parkinson’s disease-related amyloids, but they can be

formed even by such an abundant protein as lysozyme.

• The presence of very stable hydrophobic core in equine lysozyme

structure prevents amyloid protofilaments from significant elongation

and lateral assembly. The strategy of designing of the specific highly

stable and cooperative areas within protein molecule can be explored as

amyloid preventing approach in additional to the overall stabilisation of

the protein fold.

• We have shown that full-length equine lysozyme forms amyloid,

though on the longer time-scale acidic hydrolysis can also contribute to

the fibril formation.

• We have shown that de novo protein albebetin and its constructs with

biologically active peptides are inherently highly amyloidogenic,

forming a range of amyloid species under physiological conditions.

• We demonstrated the multiple pathways of amyloid assembly of

albebetin, including formation of a hierarchy of on- and off-pathway

oligomeric intermediates. The pivotal oligomers comprised of 12

51

protein monomers are the key-point, in which the amyloid pathways

split apart. The conversion of the pivotal oligomers into amyloid-prone

ones is considered as a rate limiting step in amyloid assembly.

Amyloid-prone oligomers are characterized by the cross-β-sheet

formation, giving rise to amyloid protofilaments and fibrils.

• Electrostatic interactions have been shown to be critical in amyloid

assembly of albebetin. The fusion of the short polar peptides to the N-

terminus of albebetin as well as increasing ionic strength significantly

enhances its amyloidogenicity.

• We demonstrated that the amyloid toxicity of albebetin is clearly

correlates with the development the cross-β-sheet in their oligomeric

assemblies. The pivotal oligomers, which do not contain cross-β-sheet

core, are apparently non-toxic, while the larger amyloid-prone

oligomers, comprised of ca. 30 and more molecules and possessing

well-developed cross-β-sheet pattern, induce the cell death. The initial

monomeric albebetin and its protofilaments do not display toxicity on

the neuronal cells.

• We demonstrated that amyloid oligomers and fibrils of hen lysozyme

induce death of SH-SY5Y cells via different mechanisms.

• We conclude that oligomers induce apoptosis-like cell death, while the

fibrils lead to necrosis-like death.

52

Acknowledgements I would like to express my sincere gratitude to:

Entire department of Medical Biochemistry and Biophysics for creating the

nice atmosphere.

My supervisor Ludmilla Morozova-Roche, for everything you have done for

me both in science and in life in general. I will never get better boss than you!

Gerhard Gröbner, for the support, infinite optimism and for the introducing the

phrase “gut, gut” in my lexicon.

Mantas, for the friendship and for the stories beginning “once upon a time we

visited Systembolaget”.

Anna, Kristina, Kiran – we are the very group.

Our administrator, Ingrid Råberg, for the guidance through the peculiarities of

life in Sweden.

Gena, for the hockey and the brain storm in chess.

Maribel, for your shine. No pasaran!

Katya, Mattias, Linus, Peter - tack att ni finns.

And all russian-speaking community in Umeå.

Моим друзьям и тем, кто был в Умео визитом на короткое время, я вас

всех помню.

Кате, за твой созидательно-подстригательный креатив, способный

потрясти окружающих, включая мою жену.

Моей маме за ее безграничную любовь и веру!

Моим любимым, жене и дочери, за приобретение смысла в жизни! Я

люблю вас, мои родные!

Thank you!!!

53

References

Acharya K.R., Stuart D.I., Phillips D.C., McKenzie H.A. and Teahan C.G. (1994) Models of the three-dimensional structures of echidna, horse, and pigeon lysozymes: calcium-binding lysozymes and their relationship with α-lactalbumins. J. Prot. Chem. 6, 569-584.

Ahmad A., Millett I. S., Doniach S., Uversky V.N. Fink A L. (2003) Partially folded intermediates in insulin fibrillation. Biochemistry 42, 11404–11416.

Ahmad A., Millett I.S., Doniach S., Uversky V.N., Fink A.L. (2004) Stimulation of insulin fibrillation by urea-induced intermediates. J. Biol. Chem. 279, 14999-5013.

Anderluh G., Gutierrez-Aguirre I., Rabzelj S., Ceru S., Kopitar-Jerala N., Macek P., Turk V, Zerovnik E. (2005) Interaction of human stefin B in the prefibrillar oligomeric form with membranes. Correlation with cellular toxicity. FEBS J. 272, 3042–3051.

Andersson K., Olofsson A., Nielsen E. H., Svehag S. E., Lundgren E. (2002) Only amyloidogenic intermediates of transthyretin induce apoptosis. Biochem. Biophys. Res. Commun. 294, 309–314.

Aphasizheva, I. Y., Dolgikh, D. A., Abdullaev, Z. K., Uversky, V. N., Kirpichnikov, M. P., and Ptitsyn, O. B. (1998) Can grafting of an octapeptide improve the structure of a de novo protein? FEBS Lett. 425, 101-104.

Atwood C.S., Moir R.D., Huang X., Scarpa R.C., Bacarra N.M., Romano D.M., Hartshorn M.A., Tanzi R.E., Bush A.I. (1998) Dramatic aggregation of Alzheimer abeta by Cu(II) is induced by conditions representing physiological acidosis. J. Biol. Chem. 273, 12817-12826.

Ayed A., Duckworth H.W. (1999) A stable intermediate in the equilibrium unfolding of Escherichia coli citrate synthase. Protein Sci. 8, 1116-1126.

Baskakov I.V., Legname G., Gryczynski Z., Prusiner S.B. (2004) The peculiar nature of unfolding of the human prion protein. Protein Sci. 13, 586-595.

Bellotti V., Mangione P. and Merlini G. (2000) Review: Immunoglobulin light chain amyloidosis - the archetype of structural and pathogenic variability. J. Struct. Biol. 130, 280–289

Blake C.C.F. and Serpell L.C. (1996) Synchrotron X-ray studies suggest that the core of the transthyretin amyloid fibril is a continuous beta-sheet helix. Structure 4, 989-998.

54

Bhatia R., Lin H., Lal R. (2000) Fresh and globular amyloid β protein (1–42) induces rapid cellular degeneration: evidence for AβP channel-mediated cellular toxicity. FASEB J. 14, 1233–1243.

Bocharova O.V., Moshkovskii S.A., Chertikova R.V., Abdullaev Z.Kh., Kolesanova E.F., Dolgikh D.A. and Kirpichnikov M.P. (2002) Introduction of biologically active fragments of interferon-R2 and insulin into the artificial protein albebetin affects immunogenicity of the final construct. J. Mol. Biol. 6, 84-90.

Bonar L., Cohen A.S., Skinner M.M. (1969) Characterization of the amyloid fibril as a cross-beta protein. Proc. Soc. Exp. Biol. Med. 131, 1373–1375.

Booth D.R., Sunde M., Bellotti V., Robinson C.V., Hutchinson W.L., Fraser P.E., Hawkins P.N., Dobson C.M., Radford S.E., Blake C.C., Pepys M.B. (1997) Instability, unfolding and aggregation of human lysozyme variants underlying amyloid fibrillogenesis. Nature 385, 787-793.

Blanch E.W., Morozova-Roche L.A., Hecht L., Noppe W., Barron L.D. (2000) Raman optical activity characterization of native and molten globule states of equine lysozyme: comparison with hen lysozyme and bovine alpha-lactalbumin. Biopolymers 57, 235-248.

Brange J., Dodson G.G., Edwards D.J., Holden P.H., Whittingham J.L. (1997) A model of insulin fibrils derived from the x-ray crystal structure of a monomeric insulin (despentapeptide insulin). Proteins 27, 507-516.

Brookes P.S., Yoon Y., Robotham.JL., Anders M.W., Sheu.SS. (2004) Calcium, ATP, and ROS: a mitochondrial love-hate triangle. Am. J. Physiol. Cell. Physiol. 287, C817-C833

Bucciantini M., Giannoni E., Chiti F., Baroni F., Formigli L., Zurdo J., Taddei N., Ramponi G., Dobson C.M., Stefani M (2002) Inherent toxicity of aggregates implies a common mechanism for protein misfolding diseases. Nature 416, 507–511.

Butterfield D.A., Hensley K., Harris M., Mattson M., Carney J. (1994) Betaamyloid peptide free radical fragments initiate synaptosomal lipoperoxidation in a sequence-specific fashion: implications to Alzheimer’s disease. Biochem. Biophys. Res. Commun. 200, 710–715.

Butterfield D.A., Drake J., Pocernich C., Castegna A. (2001) Evidence of oxidative damage in Alzheimer's disease brain: central role for amyloid beta-peptide. Trends Mol. Med. 7, 548-554.

Canet D., Sunde M., Last A.M., Miranker A., Spencer A., Robinson C.V., Dobson C.M. (1999) Mechanistic studies of the folding of human

55

lysozyme and the origin of amyloidogenic behavior in its disease-related variants. Biochemistry 38, 6419–6427.

Canet D., Last A.M., Tito P., Sunde M., Spencer A., Archer D.B., Redfield C., Robinson C.V., Dobson C.M. (2002) Local cooperativity in the unfolding of an amyloidogenic variant of human lysozyme. Nat. Struct. Biol. 9, 308–315.

Casley C.S., Canevari L., Land J.M., Clark J.B., Sharpe M.A. (2002) Beta-amyloid inhibits integrated mitochondrial respiration and key enzyme activities. J. Neurochem. 80, 91–100.

Carter D.B., Chou K.C. (1998) A model for structure-dependent binding of Congo red to Alzheimer beta-amyloid fibrils. Neurobiol. Aging. 19, 37-40.

Chamberlain A.K., MacPhee C.E., Zurdo J., Morozova-Roche L.A., Hill H.A., Dobson C.M., Davis JJ. (2000) Ultrastructural organization of amyloid fibrils by atomic force microscopy. Biophys. J. 79, 3282-3293.

Chapman M.R., Robinson L.S., Pinkner J.S., Roth R., Heuser J., Hammar M., Normark S., Hultgren S.J. (2002) Role of Escherichia coli curli operons in directing amyloid fiber formation. Science 295, 851-855

Chertkova R.V., Kostanian I.A., Astapova M.V., Surina E.A., Dolgikh D.A. and Kirpichnikov M.P. (2003) An artificial protein, possessing biological activity of HL-60 human promyelocytic leukemia differentiation factor. Bioorg. Khim. 29, 30-37.

Chiti F., Webster P., Taddei N., Clark A., Stefani M., Ramponi G. & Dobson C.M. (1999) Designing conditions for in vitro formation of amyloid protofilaments and fibrils. Proc. Natl Acad. Sci. USA 96, 3590–3594.

Chiti F, Taddei N, Bucciantini M, White P, Ramponi G, Dobson CM. 2000 Mutational analysis of the propensity for amyloid formation by a globular protein. EMBO J. 19, 1441-1449.

Chiti F., Taddei N., Baroni F., Capanni C., Stefani M., Ramponi G. and Dobson C.M. (2002) Kinetic partitioning of protein folding and aggregation. Nature Struct. Biol. 9, 137–143.

Chromy B.A., Nowak R.J., Lambert M.P., Viola K.L., Chang L., Velasco P.T., Jones B.W., Fernandez S.J., Lacor P.N., Horowitz P., Finch C.E., Krafft G.A., Klein W.L. (2003) Selfassembly of Aβ(1-42) into globular neurotoxins. Biochemistry 42, 12749–12760.

Cohen A.S., Jones L.A. (1991) Amyloidosis. Curr. Opin. Rheumatol. 3, 125-38.

Cooper A.A., Gitler A.D., Cashikar A., Haynes C.M., Hill K.J., Bhullar B., Liu K., Xu K., Strathearn K.E., Liu F., Cao S., Caldwell K.A.,

56

Caldwell G.A., Marsischky G., Kolodner R.D., Labaer J., Rochet J.C., Bonini N.M., Lindquist S. (2006) Alpha-synuclein blocks ER-Golgi traffic and Rab1 rescues neuron loss in Parkinson's models. Science 313, 24–328.

Crompton M. (2004) Mitochondria and aging: a role for the permeability transition? Aging Cell. 3, 3-6.

Cymes G.D., Grosman C., Delfino J.M., Wolfenstein-Todel C. (1996) Detection and characterization of an ovine placental lactogen stable intermediate in the urea-induced unfolding process. Protein Sci. 5, 2074-2079.

Damaschun G., Damaschun H., Fabian H., Gast K., Krober R., Wieske M. & Zirwer D. (1999) Conversion of yeast phosphoglycerate kinase into amyloid-like structure. Proteins 39, 204–211.

Damaschun G., Damaschun H., Gast K. and Zirwer D. (1999) Proteins can adopt totally different folded conformations. J. Mol. Biol. 291, 715–725.

De Felice F.G., Vieira M.N., Meirelles M.N., Morozova-Roche L.A., Dobson C.M., Ferreira S.T. (2004) Formation of amyloid aggregates from human lysozyme and its disease-associated variants using hydrostatic pressure. FASEB J. 10, 1099-1101.

De Giorgi F., Lartigue L., Bauer M.K., Schubert A., Grimm S., Hanson G.T., Remington S.J., Youle R.J., Ichas F. (2002) The permeability transition pore signals apoptosis by directing Bax translocation and multimerization. FASEB J. 16, 607-609.

Der-Sarkissian A, Jao CC, Chen J, Langen R. (2003) Structural organization of alpha-synuclein fibrils studied by site-directed spin labeling. J. Biol. Chem. 278, 37530-37535.

Ding T.T., Lee S.J., Rochet J C. & Lansbury, P.T., Jr (2002). Annular alpha-synuclein protofibrils are produced when spherical protofibrils are incubated in solution or bound to brain-derived membranes. Biochemistry 41, 10209–10217.

Dobson C.M. (1999) Protein misfolding, evolution and disease. Trends Biochem. Sci.24, 329-332.

Dobson C.M. (2001) The structural basis of protein folding and its links with human disease. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 356, 133-145.

Dolgikh D.A., Fedorov A.N., Chermeris V.V., Chernov B.K., Finkel’shtein A.V., Shul’ga A.A., Alakhov I.B., Kirpichnikov M.P., and Ptitsyn O.B. (1991) Preparation and study of albebetin, an artificial protein with a given spatial structure. Dokl. Akad. Nauk SSSR 320, 1266-1269.

57

Dolgikh D. A., Uversky V. N., Gabrielian A. E., Chemeris V. V., Fedorov A. N., Navolotskaya E. V., Zav’yalov V. P. and Kirpichnikov M. P. (1996) The de novo protein with grafted biological function: transferring of interferon blast- transforming activity to albebetin. Protein Eng. 9, 195-201.

Dos Reis S, Coulary-Salin B, Forge V, Lascu I, Begueret J, Saupe S.J. (2002) The HET-s prion protein of the filamentous fungus Podospora anserina aggregates in vitro into amyloid-like fibrils. J. Biol. Chem. 277, 5703-5706.

Dunker A.K., Lawson J.D., Brown C.J., Williams R.M., Romero P., Oh J.S., Oldfield C.J., Campen A.M., Ratliff C.M., Hipps K.W., Ausio J., Nissen M.S., Reeves R., Kang C., Kissinger C.R., Bailey R.W., Griswold M.D., Chiu W., Garner E.C. Obradovic Z. (2001) Intrinsically disordered protein. J. Mol. Graph. Model. 19, 26–59.

Eakin C.M., Knight J.D., Morgan C.J., Gelfand M.A., Miranker A.D. (2002) Formation of a copper specific binding site in non-native states of beta-2-microglobulin. Biochemistry 41, 10646-10656.

Eckert A., Keil U., Marques C.A., Bonert A., Frey C., Schussel K., Muller W.E. (2003) Mitochondrial dysfunction, apoptotic cell death, and Alzheimer's disease. Biochem. Pharmacol. 66, 1627-1634.

Ekiel I. & Abrahamson M. (1996) Folding-related dimerization of human cystatin C. J. Biol. Chem. 271, 1314–1321.

Fandrich M., Fletcher M.A., Dobson C.M. (2001) Amyloid fibrils from muscle myoglobin. Nature 410, 165-166.

Ferreira S.T. & De Felice F.G. (2001) Protein dynamics, folding and misfolding: from basic physical chemistry to human conformational diseases. FEBS Lett. 498, 129–134.

Fedorov A.N., Dolgikh D.A., Chemeris V.V., Chernov B.K., Finkelstein A.V., Schulga A.A., Alakhov Y.B., Kirpichnikov M.P. and Ptitsyn O.B. (1992) De novo design, synthesis and study of albebetin, a polypeptide with a predetermined threedimensional structure. Probing the structure at the nanogram level. J. Mol. Biol. 225, 927-931.

Forloni G., Bugiani O., Tagliavini F. & Salmona M. (1996) Apoptosis-mediated neurotoxicity induced by β-amyloid and PrP fragments. Mol. Chem. Neuropathol. 28, 163–171.

Forloni G. (1996) Neurotoxicity of β-amyloid and prion peptides. Curr. Opin. Neurol. 9, 492–500.

Guijarro JI, Sunde M, Jones JA, Campbell ID, Dobson CM. 1998 Amyloid fibril formation by an SH3 domain. Proc Natl Acad Sci U S A. 95, 4224-4228.

58

Glenner G.G., Eanes E.D., Bladen, H.A. (1974) Beta-pleated sheet fibrils. A comparison of native amyloid with synthetic protein fibrils, J. Histochem. Cytochem. 22, 1141–1158.

Glenner G. (1980) Amyloid deposits and amyloidosis. N. Engl. J. Med. 302, 1283-1292.

Goda S., Takano K., Yamagata Y., Nagata R., Akutsu H., Maki S., Namba K., Yutani K. (2000) Amyloid protofilament formation of hen egg lysozyme in highly concentrated ethanol solution. Protein Sci. 9, 369-375.

Goers J., Permyakov S.E., Permyakov E.A., Uversky V.N., Fink A.L. (2002) Conformational prerequisites for alpha-lactalbumin fibrillation. Biochemistry 41, 12546-12551.

Gong Y., Chang L., Viola K.L., Lacor P.N., Lambert M.P., Finch C.E. Grant A. Krafft G.A. and Klein W.L. (2003) Alzheimer's diseaseaffected brain: presence of oligomeric Aβ ligands (ADDLs) suggests a molecular basis for reversible memoryloss. Proc. Natl. Acad. Sci.USA 100, 10417–10422.

Griko Y.V., Freire E., Privalov G., van Dael H. and Privalov P.L. (1995) The unfolding thermodynamics of c-type lysozymes: a calorimetric study of the heat denaturation of equine lysozyme. J Mol Biol. 252, 447-459.

Han H., Weinreb P.H. and Lansbury P.T.Jr (1995) The core Alzheimer’s peptide NAC forms amyloid fibrils which seed and are seeded by beta-amyloid: is NAC a common trigger or target in neurodegenerative disease? Chem. Biol. 2, 163–169.

Hamada D., Dobson C.M. (2002) A kinetic study of beta-lactoglobulin amyloid fibril formation promoted by urea. Protein Sci. 10, 2417-2426.

Hammarstrom P, Jiang X, Hurshman A.R., Powers E.T., Kelly J.W. (2002) Sequence-dependent denaturation energetics: A major determinant in amyloid disease diversity. Proc Natl Acad Sci USA 99, 16427-16432.

Hartley D.M., Walsh D.M., Ye C.P., Diehl T., Vasquez S., Vassilev P.M., Teplow D.B., Selkoe D.J. (1999) Protofibrillar intermediates of amyloid beta-protein induce acute electrophysiological changes and progressive neurotoxicity in cortical neurons. J. Neurosci. 19, 8876-8884.

Heegaard N.H., Sen J.W., Kaarsholm N.C., Nissen, M.H. (2001) Conformational intermediate of the amyloidogenic protein beta 2-microglobulin at neutral pH. J. Biol. Chem. 276, 32657–32662.

Hengartner M.O. (2000)The biochemistry of apoptosis. Nature 407, 770–776.

59

Howlett D.R., Jennings K.H., Lee D.C., Clark M.S., Brown F., Wetzel R. Wood S.J., Camilleri P., Roberts G.W. (1995) Aggregation state and neurotoxic properties of Alzheimer β-amyloid peptide. Neurodegeneration 4, 23–32.

Hoyer W., Cherny D., Subramaniam V. and Jovin T. M. (2004) Impact of the acidic C-terminal region comprising amino acids 109–140 on alpha-synuclein aggregation in vitro. Biochemistry 43, 16233–16242.

Huang X., Atwood C.S., Moir R.D., Hartshorn M.A., Vonsattel J.P., Tanzi R.E., Bush A.I. (1997) Zinc-induced Alzheimer's Abeta(1-40) aggregation is mediated by conformational factors. J. Biol. Chem. 272, 26464-26470.

Iconomidou V.A, Vriend G., Hamodrakas S.J. (2000) Amyloids protect the silkmoth oocyte and embryo. FEBS Lett. 479, 141-145.

Jäättelä M., Wissing D., Kokholm, K, Kallunki T. & Egeblad M. (1998) Hsp70 exerts its anti-apoptotic function downstream of caspase-3-like proteases. EMBO J. 17, 6124–6134.

Jahn T.R., Radford S.E. (2005) The Yin and Yang of protein folding. FEBS J. 272, 5962-5970.

Jansen R., Dzwolak W., Winter R. (2005) Amyloidogenic self-assembly of insulin aggregates probed by high resolution atomic force microscopy. Biophys. J. 88, 1344-1353.

Janson J., Ashley R.H., Harrison D., McIntyre S., Butler P.C. (1999) The mechanism of islet amyloid polypeptide toxicity is membrane disruption by intermediate-sized toxic amyloid particles. Diabetes 48, 491-498.

Jarrett J.T. and Lansbury P.T. (1993) Seeding "one-dimensional crystallization" of amyloid: A pathogenic mechanism in Alzheimer's disease and scrapie? Cell 73, 1055-1058.

Jones C.E., Abdelraheim S.R., Brown D.R., Viles J.H. (2004) Preferential Cu2+ coordination by His96 and His111 induces beta-sheet formation in the unstructured amyloidogenic region of the prion protein. J. Biol. Chem. 279, 32018-32027.

Katayama T., Imaizumi K., Manabe T., Hitomi J., Kudo T., Tohyama M. (2004) Induction of neuronal death by ER stress in Alzheimer’s disease. J. Chem. Neuroanat. 28, 67–78.

Kaufman R.J. (1999) Stress signaling from the lumen of the endoplasmic reticulum: coordination of gene transcriptional and translational controls. Genes. Dev. 13, 1211–1233.

Kayed R., Bernhagen J., Greenfield N., Sweimeh K., Brunner H., Voelter W. and Kapurniotu A. (1999) Conformational transitions of Islet

60

Amyloid Polypeptide (IAPP) in amyloid formation in vitro. J. Mol. Biol. 287 781–796.

Kayed R., Head E., Thompson J.L., McIntire T.M., Milton S.C., Cotman C.W. & Glabe C.G. (2003) Common structure of soluble amyloid oligomers implies common mechanism of pathogenesis. Science 300, 486–489.

Kelly J.W. (1998) The alternative conformations of amyloidogenic proteins and their multi-step assembly pathways. Curr. Opin. Struct. Biol. 8, 101–106.

Kenney J.M., Knight D., Wise M.J., Vollrath F. (2002) Amyloidogenic nature of spider silk. Eur J Biochem. 269, 4159-4163.

King C.Y., Tittmann P., Gross H., Gebert R., Aebi M., Wuthrich K. (1997) Prion-inducing domain 2-114 of yeast Sup35 protein transforms in vitro into amyloid-like filaments. Proc. Natl. Acad. Sci. USA 94, 6618-6622.

Knight J.D., Miranker A.D. (2004) Phospholipid catalysis of diabetic amyloid assembly. J. Mol. Biol. 341, 1175-1187.

Kobashigawa Y., Demura M., Koshiba T., Kumaki Y., Kuwajima K. and Nitta K. (2000) Hydrogen exchange study of canine milk lysozyme: stabilization mechanism of the molten globule. Proteins 40, 579-589.

Koradi R., Billeter M., and Wuthrich K. (1996) MOLMOL: a program for display and analysis of macromolecular structures. J. Mol. Graphics 14, 51-55.

Koshiba T., Yao M., Kobashigawa Y., Demura M., Nakagawa A., Tanaka I., Kuwajima K. and Nitta K. (2000) Structure and thermodynamics of the extraordinarily stable molten globule state of canine milk lysozyme. Biochemistry 39, 3248-3257.

Kostanyan I. A., Astapova M. V., Starovoytova E. V., Dranitsina S. M. and Lipkin V. M. (1994) A new human leukemia cell 8.2 kDa differentiation factor: isolation and primary structure. FEBS Lett. 356, 327-329.

Krebs M.R., Wilkins D.K., Chung E.W., Pitkeathly M.C., Chamberlain A.K., Zurdo J., Robinson C.V., Dobson C.M. (2000) Formation and seeding of amyloid fibrils from wild-type hen lysozyme and a peptide fragment from the beta-domain. J. Mol. Biol. 300, 541-549.

Krebs M.R., Morozova-Roche L.A., Daniel K., Robinson C.V., Dobson C.M. (2004) Observation of sequence specificity in the seeding of protein amyloid fibrils. Protein Sci. 13, 1933-1938.

Krebs M.R., Bromley E.H., Donald A.M. (2005) The binding of thioflavin T to amyloid fibrils: localisation and implications. J. Struct. Biol. 149, 30-37.

Krishna M.M., Englander S.W. (2007) A unified mechanism for protein folding: predetermined pathways with optional errors. Protein Sci. 16, 449-464.

Kuwajima K., Hiraoka Y., Ikeguchi M., Sugai S. (1985) Comparison of the transient folding intermediates in lysozyme and -lactalbumin. Biochemistry 24, 874-881

Kyle R.A. (2001) Amyloidosis: a convoluted story. Brit. J. Haem. 114, 529-538.

Lai Z., Colon W. & Kelly J.W. (1996) The acid-mediated denaturation pathway of transthyretin yields a conformational intermediate that can self-assemble into amyloid. Biochemistry 35, 6470–6482.

Lashuel H.A., Hartley D., Petre B.M., Walz T., Lansbury P.T. Jr. (2002) Neurodegenerative disease: amyloid pores from pathogenic mutations. Nature 418, 291.

Lashuel H.A., Lansbury P.T. Jr. (2006) Are amyloid diseases caused by protein aggregates that mimic bacterial pore-forming toxins? Q Rev Biophys. 39, 167-201.

Leist M., Jaattela M. (2001) Four deaths and a funeral: from caspases to alternative mechanisms. Nat. Rev. Mol. Cell Biol. 2, 589-598.

Lomakin A., Chung D.S., Benedek G.B., Kirschner D.A. and Teplow D.B. (1996) On the nucleation and growth of amyloid beta-protein fibrils: detection of nuclei and quantitation of rate constants. Proc. Natl. Acad. Sci. USA 93, 1125–1129.

Lorenzo A. & Yankner B. A. (1994) β-amyloid neurotoxicity requires fibril formation and is inhibited by congo red. Proc. Natl Acad. Sci. USA 91, 12243–12247.

Lustbader J.W., Cirilli M., Lin C., Xu H.W., Takuma K., Wang N., Caspersen C., Chen X., Pollak S., Chaney M., Trinchese F., Liu S., Gunn-Moore F., Lue L.F., Walker D.G., Kuppusamy P., Zewier Z.L., Arancio O., Stern D., Yan S.S., Wu H. (2004) ABAD directly links Abeta to mitochondrial toxicity in Alzheimer's disease. Science 304, 448-452.

Malisauskas M., Zamotin V., Jass J., Noppe W., Dobson C.M., Morozova-Roche L.A. (2003) Amyloid protofilaments from the calcium-binding protein equine lysozyme: formation of ring and linear structures depends on pH and metal ion concentration. J. Mol. Biol. 330, 879-890.

Malisauskas, M., Ostman, J., Darinskas, A., Zamotin, V., Liutkevicius, E., Lundgren, E. & Morozova-Roche, L. A. (2005). Does the cytotoxic effect of transient amyloid oligomers from common equine lysozyme in vitro imply innate amyloid toxicity? J. Biol. Chem. 280, 6269–6275.

61

62

Matagne A., Dobson C.M. (1998) The folding process of hen lysozyme: a perspective from the 'new view'. Cell Mol. Life. Sci. 54, 363-371.

Mattson M.P. (2000) Apoptosis in neurodegenerative disorders. Nature Rev. Mol. Cell Biol. 1, 120–129

McParland V.J., Kalverda A.P., Homans S.W., and Radford S.E. (2002) Structural properties of an amyloid precursor of beta(2)-microglobulin. Nat. Struct. Biol. 9, 326–331.

Merlini G. and Bellotti V. (2003) Molecular mechanisms of amyloidosis. N. Engl. J. Med. 349, 583–596.

Merlini G. and Bellotti V. (2005) Lysozyme: a paradigmatic molecule for the investigation of protein structure, function and misfolding. Clin. Chim. Acta 357, 168–172.

Mishra R., Sorgjerd K., Nystrom S., Nordigarden A., Yu Y.C., Hammarstrom P. (2007) Lysozyme amyloidogenesis is accelerated by specific nicking and fragmentation but decelerated by intact protein binding and conversion. J. Mol. Biol. 366, 1029-1044.

Misonou H., Morishima-Kawashima M., Ihara Y. (2000) Oxidative stress induces intracellular accumulation of amyloid beta-protein (Abeta) in human neuroblastoma cells. Biochemistry 39, 6951–6959.

Missmahl H.P. and Hartwig M. (1953) Polarisationsoptischeuntersuchungen an der amyloidsubstance, Virchows. Arch. Pathol. Anat. 324, 489–508.

Mizuguchi M., Arai M., Ke Y., Nitta K., Kuwajima K. (1998) Equilibrium and kinetics of the folding of equine lysozyme studied by circular dichroism spectroscopy. J. Mol. Biol. 283, 265-277.

Morgan C.J., Gelfand M., Atreya C., Miranker A.D. (2001) Kidney dialysis-associated amyloidosis: a molecular role for copper in fiber formation. J. Mol. Biol. 309, 339-345.

Morishima Y., Gotoh Y., Zieg J., Barrett T., Takano H., Flavell R., Davis R.J., Shirasaki Y, Greenberg M.E. (2001) Beta-amyloid induces neuronal apoptosis via a mechanism that involves the c-Jun N-terminal kinase pathway and the induction of Fas ligand. J. Neurosci. 21, 7551–7560.

Morozova L., Haezebrouck P. and Van Cauwelert F. (1991) Stability of equine lysozyme. Thermal unfolding behaviour. Biophys. Chem. 41, 185–191.

Morozova L.A., Haynie D.T., Arico-Muendel C., Van Dael H., Dobson C.M. (1995) Structural basis of the stability of a lysozyme molten globule. Nat. Struct. Biol. 10, 871-875.

Morozova-Roche L.A., Arico-Muendel C.C., Haynie D.T., Emelyanenko V.I., Van Dael H., Dobson C.M. (1997) Structural characterisation and

63

comparison of the native and A-states of equine lysozyme. J. Mol. Biol. 268, 903-921.

Morozova-Roche L.A., Jones J.A., Noppe W. & Dobson C.M. (1999) Independent nucleation and heterogeneous assembly of structure during folding of equine lysozyme. J. Mol. Biol. 289, 1055-1073.

Morozova-Roche L.A., Zurdo J., Spencer A., Noppe W., Receveur V., Archer D.B., Joniau M., Dobson C.M. (2000) Amyloid fibril formation and seeding by wild-type human lysozyme and its disease-related mutational variants. J. Struct. Biol. 130, 339-351.

Morozova-Roche L.A., Zamotin V., Malisauskas M., Ohman A., Chertkova R., Lavrikova M.A., Kostanyan I.A., Dolgikh D.A, Kirpichnikov M.P. (2004) Fibrillation of carrier protein albebetin and its biologically active constructs. Multiple oligomeric intermediates and pathways. Biochemistry 43, 9610-9619.

Nakagawa T., Yuan J. (2000) Cross-talk between two cysteine protease families. Activation of caspase-12 by calpain in apoptosis. J Cell Biol. 150, 887–894.

Nielsen L., Khurana R., Coats A., Frokjaer S., Brange J., Vyas S., Uversky V.N., Fink A.L. (2001) Effect of environmental factors on the kinetics of insulin fibril formation: elucidation of the molecular mechanism. Biochemistry 40, 6036-6046.

Nielsen L., Frokjaer S., Brange J., Uversky V.N. and Fink A.L. (2001) Probing the mechanism of insulin fibril formation with insulin mutants. Biochemistry 40, 8397–8409.

Nilsson M.R., Raleigh D.P. (1999) Analysis of amylin cleavage products provides new insights into the amyloidogenic region of human amylin. J. Mol. Biol. 294, 1375-13785.

Nitta K., Tsuge H., Sugai S. and Shimazaki K. (1987) The calcium-binding property of equine lysozyme. FEBS Lett. 223, 405-408.

Novitskaya V., Bocharova O.V., Bronstein I. & Baskakov I.V. (2006) Amyloid fibrils of mammalian prion protein are highly toxic to cultured cells and primary neurons. J. Biol. Chem. 281, 13828–13836.

Nunomura A., Castellani R.J., Zhu X., Moreira P.I., Perry G., Smith M.A. (2006) Involvement of oxidative stress in Alzheimer disease J. Neuropathol. Exp. Neurol. 65, 631-641.

Ojaimi J., Byrne E. (2001) Mitochondrial function and alzheimer's disease. Biol. Signals Recept. 10, 254-262.

Panick G., Malessa R. and Winter R. (1999) Differences between the pressure- and temperature-induced denaturation and aggregation of beta-lactoglobulin A, B, and AB monitored by FT-IR spectroscopy and small-angle X-ray scattering, Biochemistry 38, 6512–6519.

64

Pasquato N., Berni R., Folli C., Alfieri B., Cendron L., Zanotti G. (2007) Acidic pH-induced conformational changes in amyloidogenic mutant transthyretin. J. Mol. Biol. 366, 711-719.

Pedersen J.S., Dikov D., Flink J.L., Hjuler H.A. Christiansen G., Otzen D.E. (2006) The changing face of glucagon fibrillation: structural polymorphism and conformational imprinting. J. Mol. Biol. 355, 501-523.

Pepys M.B., Hawkins P.N., Booth D.R., Vigushin D.M., Tennent G.A. and Soutar A.K. (1993) Human lysozyme gene mutations cause hereditary systemic amyloidosis, Nature 362, 553–557.

Permyakov E.A. and Berliner L.J. (2000) α-Lactalbumin: structure and function. FEBS Lett. 473, 269-274.

Permyakov E.A., Permyakov S.E., Deikus G.Y., Morozova-Roche L.A., Grishchenko V.M., Kalinichenko L.P., Uversky V.N. (2003) Ultraviolet illumination-induced reduction of alpha-lactalbumin disulfide bridges. Proteins 51, 498-503.

Permyakov S.E., Khokhlova T.I., Nazipova A.A., Zhadan A.P., Morozova-Roche L.A. and Permyakov E.A. (2006) Calcium-binding and temperature induced transitions in equine lysozyme: new insights from the pCa-temperature "phase diagrams". Proteins 65, 984-998.

Pertinhez T.A., Bouchard M., Tomlinson E.J., Wain R., Ferguson S.J., Dobson C.M., Smith L.J. (2001) Amyloid fibril formation by a helical cytochrome. FEBS Lett. 495, 184-186.

Petkova A.T., Leapman R.D., Guo Z., Ya, W.M., Mattson M.P. & Tycko R. (2005) Self-propagating, molecular-level polymorphism in Alzheimer's β-amyloid fibrils. Science 307, 262–265.

Picotti P., De Franceschi G., Frare E., Spolaore B., Zambonin M., Chiti F., de Laureto P.P., Fontana A. (2007) Amyloid fibril formation and disaggregation of fragment 1-29 of apomyoglobin: insights into the effect of ph on protein fibrillogenesis. J. Mol. Biol. 367, 1237-1245.

Privalov P.L. (1979) Stability of proteins: small globular proteins. Adv. Protein Chem. 33, 167-241.

Prusiner S.B. (1982) Novel proteinaceous infectious particles cause scrapie. Science 216, 136–144

Randolph T.W., Seefeldt M., Carpenter J.F. (2002) High hydrostatic pressure as a tool to study protein aggregation and amyloidosis. Biochim. Biophys. Acta 1595, 224–234.

Rasmussen P., Barbiroli A., Bonomi F., Faoro F., Ferranti P., Iriti M., Picariello G., Iametti S. (2007) Formation of structured polymers upon controlled denaturation of beta-lactoglobulin with different chaotropes. Biopolymers 86, 57-72.

65

Reixach N., Deechongkit S., Jiang X., Kelly J.W. & Buxbaum J.N. (2004) Tissue damage in the amyloidoses: transthyretin monomers and nonnative oligomers are the major cytotoxic species in tissue culture. Proc. Natl. Acad. Sci. USA, 101, 2817–2822.

Rochet J.C. and Lansbury P.T. Jr (2000) Amyloid fibrillogenesis: themes and variations. Curr. Opin. Struct. Biol. 10, 60–68.

Rodriguez-Enriquez S., He L., Lemasters J.J (2004) Role of mitochondrial permeability transition pores in mitochondrial autophagy. Int. J. Biochem. Cell Biol. 36, 2463-24672.

Röcken C., Becker K., Fändrich M., Schroeckh V., Stix B., Rath T. Kahne T., Dierkes J., Roessner A., Albert F.W. (2006) ALys amyloidosis caused by compound heterozygosity in exon 2 (Thr70Asn) and exon 4 (Trp112Arg) of the lysozyme gene. Hum. Mutat. 27, 119–120.

Saraiva M.J.M. (2002) Hereditary transthyretin amyloidosis: molecular basis and therapeutical strategies Exp. Rev. Mol. Med. 14 May

Sasahara K., Demura M., Nitta K. (2000) Partially unfolded equilibrium state of hen lysozyme studied by circular dichroism spectroscopy. Biochemistry 39, 6475-6482.

Sasahara K., Yagi H., Naiki H., Goto Y. (2007) Heat-triggered conversion of protofibrils into mature amyloid fibrils of beta(2)-microglobulin. Biochemistry 46, 3286-3293.

Schneider S.W., Larmer J., Henderson R.M., and Oberleithner H. (1998) Molecular weights of individual proteins correlate with molecular volumes measured by atomic force microscopy. Pfluegers Arch. 435, 362-367.

Seefeldt M.B., Ouyang J., Froland W.A., Carpenter J.F. and Randolph T.W. (2004) High-pressure refolding of bikunin: Efficacy and thermodynamics, Protein Sci. 13, 2639–2650.

Serio T. R., Cashikar A. G., Kowal A. S., Sawicki G. J., Moslehi J. J., Serpell L. et al. (2000) Nucleated conformational conversion and the replication of conformational information by a prion determinant. Science 289, 1317–1321

Schulze-Osthoff K., Bakker A.C., Vanhaesebroeck B., Beyaert R., Jacob W.A., Fiers W. (1992) Cytotoxic activity of tumor necrosis factor is mediated by early damage of mitochondrial functions. Evidence for the involvement of mitochondrial radical generation. J. Biol.Chem. 267, 5317-5323.

Silva J.L. and Weber G. (1993) Pressure stability of proteins, Annu. Rev. Phys. Chem. 44, 89–113.

Sipe, J.D. & Cohen, A.S. (2000) History of the amyloid fibril. J. Struct. Biol. 130, 88–98.

66

Sirangelo I., Malmo C., Iannuzzi C., Mezzogiorno A., Bianco M. R., Papa M. & Irace G. (2004) Fibrillogenesis and cytotoxic activity of the amyloid forming apomyoglobin mutant W7F and W14F. J. Biol. Chem. 279, 13183–13189.

Smith D.P., Jones S., Serpell L.C., Sunde M., Radford S.E. (2003) A systematic investigation into the effect of protein destabilisation on beta-2-microglobulin amyloid formation. J. Mol. Biol. 330, 943–954.

Soto C. & Saborio G.P. (2001) Prions: disease propagation and disease therapy by conformational transmission. Trends Mol. Med. 7, 109–114.

Soto C. (2001) Protein misfolding and disease; protein refoldingand therapy. FEBS Lett. 498, 204–207.

Sparr E., Engel M.F., Sakharov D.V., Sprong M., Jacobs J., de Kruijff B., Hoppener J.W., Killian J.A. (2004) Islet amyloid polypeptide-induced membrane leakage involves uptake of lipids by forming amyloid fibers. FEBS Lett. 577, 117-120.

Strasser A., O'Connor L., Dixit V.M. (2000) Apoptosis signaling. Annu. Rev. Biochem. 69, 217-45.

Stuart D.I., Acharya K.R., Walker N..P, Smith S.G., Lewis M. and Phillips D.C. (1986) α-Lactalbumin possesses a novel calcium binding loop. Nature 324, 84-87.

Sousa M.M., Du Yan S., Fernandes R., Guimaraes A., Stern D. & Saraiva, M.J. (2001) Familial amyloid polyneuropathy: receptor for advanced glycation end products-dependent triggering of neuronal inflammatory and apoptotic pathways. J. Neurosci. 21, 7576–7586.

Sunde M., and Blake C.C.F. (1998) From the globular to the fibrous state: protein structure and structural conversion in amyloid formation, Q. Rev. Biophys. 31, 1–39.

Susin S.A., Lorenzo H.K., Zamzami N., Marzo I., Snow B.E., Brothers G.M., Mangion J., Jacotot E., Costantini P., Loeffler M., Larochette N., Goodlett D.R., Aebersold R., Siderovski D.P, Penninger J.M., Kroemer G. (1999) Molecular characterization of mitochondrial apoptosis-inducing factor. Nature 397, 441-446.

Swerdlow R.H., Khan S.M. (2004) A "mitochondrial cascade hypothesis" for sporadic Alzheimer's disease. Med. Hypotheses. 63, 8-20.

Takuma K., Yan S.S., Stern D.M, Yamada K. (2005) Mitochondrial dysfunction, endoplasmic reticulum stress, and apoptosis in Alzheimer's disease. J. Pharmacol. Sci. 97, 312-316

Taylor K.L., Cheng N., Williams R.W., Steven A.C., Wickner R.B. (1999) Prion domain initiation of amyloid formation in vitro from native Ure2p. Science 283, 1339-1343.

67

Teplow B., (1998) Structural and kinetic features of amyloid beta-protein fibrillogenesis. Amyloid-Int. J. Exp. Clin. Invest. 5, 121–142

Tsuge H., Koseki K., Miyano M., Shimazaki K., Chuman T., Matsumoto T., Noma M., Nitta K. and Sugai S. (1991) A structural study of calcium-binding equine lysozyme by two-dimensional 1H-NMR. Biochim. Biophys. Acta 1078, 77-84.

Tsuge H., Ago H., Noma M., Nitta K., Sugai S. and Miyano M. (1992) Crystallographic studies of a calcium binding lysozyme from equine milk at 2.5 Å resolution. J. Biochem. 111, 141-143.

Uversky V.N. (2002) What does it mean to be natively unfolded? Eur. J. Biochem. 269, 2–12

Uversky V.N., Gillespie J.R. & Fink A.L. (2000) Why are 'natively unfolded' proteins unstructured under physiologic conditions? Proteins 41, 415–427.

Uversky, V.N., Li, J. & Fink, A.L. (2001) Evidence for a partially folded intermediate in α-synuclein fibril formation. J. Biol. Chem. 276, 10737–10744.

Valleix S., Drunat S., Philit J.B., Adoue D., Piette J.C., Droz D., MacGregor B, Canet D., Delpech M., Grateau G. (2002) Hereditary renal amyloidosis caused by a new variant lysozyme W64R in a French family, Kidney Int. 61, 907–912.

Van Cruchten S., Van Den Broeck W. (2002) Morphological and biochemical aspects of apoptosis, oncosis and necrosis. Anat Histol Embryol. 31, 214-223.

Van Dael H., Haezebrouck P., Morozova L., Arico-Muendel C. and Dobson C.M. (1993) Partially folded states of equine lysozyme. Structural characterization and significance for protein folding. Biochemistry 32, 11886-11894.

Vercammen D., Brouckaert G., Denecker G., Van de Craen M., Declercq W., Fiers W, Vandenabeele P. (1998) Dual signaling of the Fas receptor: initiation of both apoptotic and necrotic cell death pathways. J. Exp. Med. 188, 919–930.

Volles M.J., Lee S.J., Rochet J.C., Shtilerman M.D., Ding T.T., Kessler J.C. & Lansbury P. T., Jr (2001) Vesicle permeabilization by protofibrillar alpha-synuclein: implications for the pathogenesis and treatment of Parkinson's disease. Biochemistry 40, 7812–7819.

Walsh D.M., Klyubin I., Fadeeva J.V., Cullen W.K., Anwyl R., Wolfe M.S., Rowan M.J., Selkoe D.J. (2002) Naturally secreted oligomers of amyloid beta protein potently inhibit hippocampal long-term potentiation in vivo. Nature 416, 535-539

68

Webb J.N., Webb S.D., Cleland J.L., Carpenter J.F. and Randolph T.W. (2001) Partial molar volume, surface area, and hydration changes for equilibrium unfolding and formation of aggregation transition state: High-pressure and cosolute studies on recombinant human IFN-gamma. Proc. Natl. Acad. Sci. USA 98, 7259–7264.

Weldon D.T., Rogers S.D., Ghilardi J.R., Finke M.P., Cleary J.P., O'Hare E., Esler W.P., Maggio J.E., Mantyh P.W. (1998) Fibrillar β-amyloid induces microglial phagocytosis, expression of inducible nitric oxide synthase, and loss of a select population of neurons in the rat CNS in vivo. J. Neurosci. 18, 2161–2173.

Wogulis M., Wright S., Cunningham D., Chilcote T.,Powell K. & Rydel, R.E. (2005) Nucleation-dependent dependent polymerization is an essential component of amyloid-mediated neuronal cell death. J. Neurosci. 25, 1071–1080.

Wood S.J., Wypych J., Steavenson S., Louis J. C., Citron, M. and Biere A.L. (1999) alpha-synuclein fibrillogenesis is nucleation-dependent. Implications for the pathogenesis of Parkinson’s disease. J. Biol. Chem. 274, 19509–19512.

Wosten H.A., de Vocht M.L. (2000) Hydrophobins, the fungal coat unravelled. Biochim. Biophys. Acta. 1469, 79-86.

Wyttenbach A., Sauvageot O., Carmichael J., Diaz-Latoud C., Arrigo A.P., Rubinsztein D.C. (2002) Heat shock protein 27 prevents cellular polyglutamine toxicity and suppresses the increase of reactive oxygen species caused by huntingtin. Hum. Mol. Genet. 11, 1137-1151

Yazaki M., Farrell S.A. and Benson M.D. (2003) A novel lysozyme mutation Phe57Ile associated with hereditary renal amyloidosis, Kidney Int. 63, 1652–1657.

Yin K.J., Lee J.M., Chen S.D., Xu J., Hsu C.Y. (2002) Amyloid-beta induces Smac release via AP-1/Bim activation in cerebral endothelial cells. J Neurosci 22, 9764–9770.

Zhao H., Tuominen E.K., Kinnunen P.K. (2004) Formation of amyloid fibers triggered by phosphatidylserine-containing membranes. Biochemistry 43, 10302-10307.

Zamotin V., Gharibyan A., Gibanova N.V., Lavrikova M.A., Dolgikh D.A., Kirpichnikov M.P. and Morozova-Roche L.A. (2006) Cytotoxicity of albebetin oligomers depends on cross-β-sheet formation. FEBS Letters, 580, 2451–2457.