29.03.2013 Views

Review: History of the Amyloid Fibril - Biochemistry and Molecular ...

Review: History of the Amyloid Fibril - Biochemistry and Molecular ...

Review: History of the Amyloid Fibril - Biochemistry and Molecular ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Journal <strong>of</strong> Structural Biology 130, 88–98 (2000)<br />

doi:10.1006/jsbi.2000.4221, available online at http://www.idealibrary.com on<br />

<strong>Review</strong>: <strong>History</strong> <strong>of</strong> <strong>the</strong> <strong>Amyloid</strong> <strong>Fibril</strong><br />

Jean D. Sipe* ,1 <strong>and</strong> Alan S. Cohen<br />

*Center for Scientific <strong>Review</strong>, National Institutes <strong>of</strong> Health, Be<strong>the</strong>sda, Maryl<strong>and</strong> 20892; <strong>and</strong> †<strong>Amyloid</strong> Program,<br />

Boston University School <strong>of</strong> Medicine, Boston, Massachusetts 02118<br />

Rudolph Virchow, in 1854, introduced <strong>and</strong> popularized<br />

<strong>the</strong> term amyloid to denote a macroscopic<br />

tissue abnormality that exhibited a positive iodine<br />

staining reaction. Subsequent light microscopic<br />

studies with polarizing optics demonstrated <strong>the</strong> inherent<br />

birefringence <strong>of</strong> amyloid deposits, a property<br />

that increased intensely after staining with<br />

Congo red dye. In 1959, electron microscopic examination<br />

<strong>of</strong> ultrathin sections <strong>of</strong> amyloidotic tissues<br />

revealed <strong>the</strong> presence <strong>of</strong> fibrils, indeterminate in<br />

length <strong>and</strong>, invariably, 80 to 100 Å in width. Using<br />

<strong>the</strong> criteria <strong>of</strong> Congophilia <strong>and</strong> fibrillar morphology,<br />

20 or more biochemically distinct forms <strong>of</strong> amyloid<br />

have been identified throughout <strong>the</strong> animal<br />

kingdom; each is specifically associated with a<br />

unique clinical syndrome. <strong>Fibril</strong>s, also 80 to 100 Å in<br />

width, have been isolated from tissue homogenates<br />

using differential sedimentation or solubility. X-ray<br />

diffraction analysis revealed <strong>the</strong> fibrils to be ordered<br />

in <strong>the</strong> beta pleated sheet conformation, with<br />

<strong>the</strong> direction <strong>of</strong> <strong>the</strong> polypeptide backbone perpendicular<br />

to <strong>the</strong> fibril axis (cross beta structure). Because<br />

<strong>of</strong> <strong>the</strong> similar dimensions <strong>and</strong> tinctorial properties<br />

<strong>of</strong> <strong>the</strong> fibrils extracted from amyloid-laden<br />

tissues <strong>and</strong> amyloid fibrils in tissue sections, <strong>the</strong>y<br />

have been assumed to be identical. However, <strong>the</strong><br />

spatial relationship <strong>of</strong> proteoglycans <strong>and</strong> amyloid P<br />

component (AP), common to all forms <strong>of</strong> amyloid, to<br />

<strong>the</strong> putative protein only fibrils in tissues, has been<br />

unclear. Recently, it has been suggested that, in<br />

situ, amyloid fibrils are composed <strong>of</strong> proteoglycans<br />

<strong>and</strong> AP as well as amyloid proteins <strong>and</strong> thus resemble<br />

connective tissue micr<strong>of</strong>ibrils. Chemical <strong>and</strong><br />

physical definition <strong>of</strong> <strong>the</strong> fibrils in tissues will be<br />

needed to relate <strong>the</strong> in vitro properties <strong>of</strong> amyloid<br />

The opinions contained herein do not necessarily represent<br />

those <strong>of</strong> <strong>the</strong> Department <strong>of</strong> Health <strong>and</strong> Human Services or <strong>the</strong><br />

National Institutes <strong>of</strong> Health.<br />

1 To whom correspondence should be addressed at Center for<br />

Scientific <strong>Review</strong>, National Institutes <strong>of</strong> Health, 6701 Rockledge<br />

Drive, Room 4106, MSC 7814, Be<strong>the</strong>sda, MD 20892. Fax: 301-<br />

480-2644. E-mail: sipej@csr.nih.gov.<br />

1047-8477/00 $35.00<br />

Copyright © 2000 by Academic Press<br />

All rights <strong>of</strong> reproduction in any form reserved.<br />

Received December 6, 1999, <strong>and</strong> in revised form January 19, 2000<br />

88<br />

protein fibrils to <strong>the</strong> pathogenesis <strong>of</strong> amyloid fibril<br />

formation in vivo. © 2000 Academic Press<br />

Key Words: amyloid fibril; amyloid protein; amyloidosis;<br />

electron microscopy; history<br />

ORIGIN OF THE TERM AMYLOID<br />

The term amyloid was introduced in 1854 by <strong>the</strong><br />

German physician scientist Rudolph Virchow (reviewed<br />

by Cohen, 1986). Utilizing <strong>the</strong> best scientific<br />

methodology <strong>and</strong> medical knowledge <strong>of</strong> <strong>the</strong> time,<br />

Virchow used iodine to stain cerebral corpora amylacea<br />

that had an abnormal macroscopic appearance.<br />

The macroscopic appearance <strong>of</strong> <strong>the</strong> brain tissue<br />

was similar to previous descriptions <strong>of</strong> o<strong>the</strong>r<br />

tissues, perhaps as early as 1639, as lardaceous<br />

liver, waxy liver, <strong>and</strong> spongy <strong>and</strong> “white stone” containing<br />

spleens (Fig. 1). When Virchow discovered<br />

that <strong>the</strong> corpora amylacea stained pale blue on<br />

treatment with iodine, <strong>and</strong> violet upon <strong>the</strong> subsequent<br />

addition <strong>of</strong> sulfuric acid, he concluded that <strong>the</strong><br />

substance underlying <strong>the</strong> evident macroscopic abnormality<br />

was cellulose <strong>and</strong> gave it <strong>the</strong> name amyloid,<br />

derived from <strong>the</strong> Latin amylum <strong>and</strong> <strong>the</strong> Greek<br />

amylon (reviewed by Cohen, 1986). At <strong>the</strong> time, <strong>the</strong><br />

distinction between starch <strong>and</strong> cellulose, as related<br />

to <strong>the</strong> separation <strong>of</strong> <strong>the</strong> plant <strong>and</strong> animal kingdoms,<br />

was unclear. From subsequent literature (reviewed<br />

by Cohen, 1986), Virchow evidently considered <strong>the</strong><br />

amyloid substance to be starch, although <strong>the</strong>re was<br />

confusion as to whe<strong>the</strong>r amyloid denotes cellulose or<br />

starch. When, in 1859, Friedreich <strong>and</strong> Kekule (reviewed<br />

by Cohen, 1986) demonstrated both <strong>the</strong> presence<br />

<strong>of</strong> protein in a “mass” <strong>of</strong> amyloid <strong>and</strong> <strong>the</strong> apparent<br />

absence <strong>of</strong> carbohydrate based on <strong>the</strong> high<br />

nitrogen content, attention was shifted to <strong>the</strong> study<br />

<strong>of</strong> amyloid first as a protein <strong>and</strong> later as a class <strong>of</strong><br />

proteins, with a propensity to undergo changes in<br />

conformation that result in fibril formation.<br />

During <strong>the</strong> 19th <strong>and</strong> early 20th centuries, investigations<br />

on <strong>the</strong> nature <strong>of</strong> amyloid evolved from <strong>the</strong><br />

macroscopic observations <strong>of</strong> Virchow <strong>and</strong> contempo-


REVIEW: HISTORY OF THE AMYLOID FIBRIL<br />

FIG. 1. <strong>Amyloid</strong> deposits at low magnification; (a) attributed to Frerichs, 1862; (b) polarization microscopy <strong>of</strong> Congo red-stained nerve<br />

biopsy from patient with ATTR, V30M; original magnification, 6.3; (c) polarization microscopy <strong>of</strong> Congo red-stained heart tissue, original<br />

magnification, 16.<br />

89


90 REVIEW: SIPE AND COHEN<br />

raries (Fig. 1) to classifications <strong>of</strong> clinical symptoms<br />

(Table I). The term amyloid has been used to describe<br />

specific macroscopic abnormalities on autopsy<br />

<strong>of</strong> body organs <strong>and</strong> tissues <strong>of</strong> patients afflicted with<br />

a variety <strong>of</strong> clinical syndromes that were first classified<br />

as idiopathic (primary) or myeloma-associated,<br />

acquired (secondary), local, hereditary, endocrine<br />

associated, <strong>and</strong> amyloid associated with aging<br />

(Table I). While introducing <strong>the</strong> term amyloid is not<br />

one <strong>of</strong> Virchow’s widely known achievements, this<br />

distinguished scientist <strong>and</strong> advocate for public<br />

TABLE I<br />

Clinical Classification <strong>of</strong> <strong>Amyloid</strong>osis<br />

Systemic amyloidosis<br />

Primary or myeloma associated amyloidosis<br />

Secondary or reactive amyloidosis<br />

Hered<strong>of</strong>amilial amyloidosis<br />

Localized (organ-limited) amyloidosis<br />

Endocrine organ or tissue<br />

“Senile” plaques, brain<br />

Cardiovascular limited (“senile” cardiac)<br />

Lichen amyloidosis (skin)<br />

FIG. 1—Continued<br />

health’s use <strong>of</strong> a unifying nomenclature has led to a<br />

fascinating field <strong>of</strong> biomedical research.<br />

MICROSCOPIC STUDIES OF AMYLOID STRUCTURE<br />

Our underst<strong>and</strong>ing <strong>of</strong> amyloid structure has advanced<br />

with <strong>the</strong> available technology. Thus, investigators<br />

began to utilize, as <strong>the</strong>y became available,<br />

light microscopy <strong>and</strong> histopathologic dyes such as<br />

thi<strong>of</strong>lavin <strong>and</strong> Congo red. Initially, <strong>the</strong>y concluded<br />

that amyloid <strong>of</strong> a variety <strong>of</strong> sources was structurally<br />

amorphous. However, subsequent polarization light<br />

microscopic studies (Divry <strong>and</strong> Florkin, 1927) demonstrated<br />

that unstained <strong>and</strong> Congo red-stained accumulations<br />

<strong>of</strong> amyloid in a variety <strong>of</strong> tissues exhibited<br />

positive birefringence with respect to <strong>the</strong> long<br />

axis <strong>of</strong> <strong>the</strong> deposits. Thus, Congo red staining imparts<br />

a marked anisotropy to <strong>the</strong> inherent birefringence<br />

exhibited by amyloid fibrils in situ (Missmahl<br />

<strong>and</strong> Hartwig, 1953). Congophilia with apple green<br />

birefringence was <strong>the</strong> first criterion <strong>of</strong> amyloid to be<br />

adopted (Fig. 2).<br />

The possibility <strong>of</strong> an ordered submicroscopic<br />

structure, as indicated by <strong>the</strong> birefringence associated<br />

with amyloid deposits, led Cohen <strong>and</strong> col-


REVIEW: HISTORY OF THE AMYLOID FIBRIL<br />

FIG. 2. (a) Congo red-stained amyloidotic uterus. (b) Birefringence <strong>of</strong> section A showing <strong>the</strong> vascular amyloid <strong>and</strong> amyloid in <strong>the</strong><br />

muscle wall.<br />

91


92 REVIEW: SIPE AND COHEN<br />

FIG. 3. An electron micrograph <strong>of</strong> amyloid fibrils in section <strong>of</strong> human amyloidotic spleen (after Shirahama <strong>and</strong> Cohen, 1967). The<br />

fibrils are arranged in r<strong>and</strong>om array <strong>and</strong> are nonbranching. The measurements <strong>of</strong> <strong>the</strong> width <strong>of</strong> individual fibrils show moderate variation,<br />

with <strong>the</strong> 100-Å width most common (dark arrows), but 200–300 width also observable (open arrows). The section was stained with uranyl<br />

acetate <strong>and</strong> lead citrate <strong>and</strong> photographed 100 000 original magnification. Reproduced from The Journal <strong>of</strong> Cell Biology, 1967, Vol. 33,<br />

pp. 679–708, by copyright permission <strong>of</strong> The Rockefeller University Press.<br />

leagues to undertake electron microscopic studies <strong>of</strong><br />

human primary <strong>and</strong> secondary amyloid tissues as<br />

well as experimental amyloid (Cohen <strong>and</strong> Calkins,<br />

1959). Their first study clearly demonstrated that<br />

all forms <strong>of</strong> amyloid studied exhibit a comparable<br />

fibrillar ultrastructure in fixed tissue sections. In<br />

<strong>the</strong>ir initial study, Cohen <strong>and</strong> Calkins analyzed sections<br />

<strong>of</strong> amyloidotic tissues from rabbit with caseininduced<br />

amyloidosis (now known to be amyloid A<br />

(AA)); skin biopsy from a patient with clinically defined<br />

primary amyloidosis (now known to be AL<br />

amyloid); <strong>and</strong> kidney <strong>of</strong> a patient with amyloidosis <strong>of</strong><br />

parenchymal organs, i.e., spleen liver <strong>and</strong> kidney<br />

(now known to be AA amyloid). It has been amply<br />

confirmed (reviewed by Cohen et al., 1982) that amyloid<br />

deposits <strong>of</strong> diverse origins in humans <strong>and</strong> animals<br />

exhibit a similar, fibrillar submicroscopic<br />

structure: bundles <strong>of</strong> straight, rigid fibrils ranging in<br />

width from 60 to 130 Å (average 75 to 100 Å) <strong>and</strong> in<br />

length from 1000 to 16,000 Å (Fig. 3). The fibrillar<br />

morphology, upon analysis <strong>of</strong> negatively stained tissue<br />

sections by electron microscopy, was adopted as<br />

<strong>the</strong> second criterion by which amyloid is defined.<br />

In <strong>the</strong>ir original paper, Cohen <strong>and</strong> Calkins<br />

suggested that human <strong>and</strong> animal amyloid possess<br />

a beaded structure. Gueft <strong>and</strong> Ghidoni (1963)<br />

suggested that amyloid fibrils are double, with<br />

an intrafibrillar space (25 Å) approximately equal<br />

FIG. 4. (a) <strong>Amyloid</strong> fibrils from human amyloidotic spleen isolated by homogenization in physiological saline followed by sucrose gradient<br />

centrifugation. Shadowed with platinum–palladium, original magnification 70 000. (b) Sucrose gradient fraction negatively stained with<br />

uranyl acetate. The amyloid fibrils usually consist <strong>of</strong> two filaments aggregated side by side. The fibrils twist occasionally (arrow). The<br />

terminations <strong>of</strong> <strong>the</strong> fibrils are smooth <strong>and</strong> rounded (circle). Original magnification 200 000. (c) <strong>Amyloid</strong> fibril fraction prior to sucrose gradient<br />

centrifugation, negatively stained with phosphotungstic acid after 5 min sonication. Both amyloid fibrils (F) <strong>and</strong> amyloid P component attached<br />

to fibrils as stacks <strong>of</strong> pentamers (R) or as individual pentamers (P) are visible (after Shirahama <strong>and</strong> Cohen, 1967). Reproduced from The Journal<br />

<strong>of</strong> Cell Biology, 1967, Vol. 33, pp. 679–708, by copyright permission <strong>of</strong> The Rockefeller University Press.


REVIEW: HISTORY OF THE AMYLOID FIBRIL<br />

93


94 REVIEW: SIPE AND COHEN<br />

to <strong>the</strong> diameter <strong>of</strong> each component (25 Å). Terry<br />

<strong>and</strong> colleagues (Terry et al., 1964), studying human<br />

cerebral amyloid in <strong>the</strong> senile plaque <strong>of</strong><br />

Alzheimer’s disease, reported that <strong>the</strong> fibrils<br />

were 70–90 Å in width <strong>and</strong> had a triple density,<br />

indicating a hollow center. Shirahama <strong>and</strong><br />

Cohen (1967) carried out extensive high-resolution<br />

studies <strong>of</strong> <strong>the</strong> amyloid fibril in tissues, <strong>and</strong><br />

from <strong>the</strong>se <strong>and</strong> o<strong>the</strong>r studies reviewed by Cohen et<br />

al. (1982), a consensus amyloid fibril structure<br />

was described, in which two ultrastructural features<br />

comprised <strong>the</strong> subunit structure <strong>of</strong> <strong>the</strong> amyloid<br />

fibril in vivo. A filamentous subunit structure,<br />

25–35 Å in diameter, longitudinally constitutes<br />

<strong>the</strong> 75- to 100-Å-wide amyloid fibril. Two or more<br />

such subunit filaments, which cross each o<strong>the</strong>r<br />

occasionally, constitute <strong>the</strong> amyloid fibril in situ<br />

(Fig. 3).<br />

IN VITRO STUDIES OF AMYLOID FIBRIL PROTEINS<br />

The identification <strong>of</strong> a fibril associated with <strong>the</strong><br />

tinctorial properties <strong>of</strong> amyloid provided a basis<br />

upon which amyloid fibrils could be isolated from<br />

tissues. Initial isolation methods, based on a gentle<br />

physical separation, homogenization in saline, followed<br />

by low-speed centrifugation, yielded a layer <strong>of</strong><br />

FIG. 4—Continued<br />

fibrils that was not present in sedimentation pellets<br />

<strong>of</strong> normal tissues (Cohen <strong>and</strong> Calkins, 1964). When<br />

stained with Congo red, <strong>the</strong> fibril layer demonstrated<br />

green birefringence with polarizing optics.<br />

Shadow casting or negative staining defined <strong>the</strong> diameter<br />

<strong>of</strong> <strong>the</strong> isolated amyloid fibril as 75–100 Å.<br />

Filamentous subunits <strong>of</strong> 25–35 Å are arranged along<br />

<strong>the</strong> long axis <strong>of</strong> <strong>the</strong> fibril in a slow twist. <strong>Amyloid</strong><br />

fibrils isolated from tissues <strong>and</strong> organs by physical<br />

separation, using sucrose gradient centrifugation<br />

(Shirahama <strong>and</strong> Cohen, 1967), showed no remarkable<br />

differences in morphology from those tissues<br />

using <strong>the</strong> two criteria <strong>of</strong> Congophilia/green birefringence<br />

<strong>and</strong> fibrillar structure (Fig. 4). The similarities<br />

in dimensions <strong>and</strong> morphology between isolated<br />

protein fibrils <strong>and</strong> amyloid fibrils in tissues has led<br />

to <strong>the</strong> widely held view that <strong>the</strong> proteinaceous fibrils<br />

studied in vitro over <strong>the</strong> past 20 years are biochemically<br />

<strong>and</strong> structurally identical to <strong>the</strong> amyloid<br />

fibrils observed in tissues.<br />

However, this view has not taken into account <strong>the</strong><br />

widespread immunohistochemical observation that<br />

o<strong>the</strong>r components <strong>of</strong> amyloid deposits in tissues do<br />

not form fibrils in vitro, but are invariably associated<br />

with all clinical/biochemical types <strong>of</strong> amyloid.<br />

These nonfibrillar components <strong>of</strong> amyloid include


TABLE II<br />

Proteins Associated with <strong>Amyloid</strong> Deposits in Humans<br />

Pathophysiology<br />

<strong>Fibril</strong> precursor<br />

(kDa)<br />

Immunity/<br />

inflammation<br />

AL Immunoglobulin<br />

light chain (23)<br />

AH Immunoglobulin<br />

heavy chain (?)<br />

AA Serum amyloid A<br />

(12)<br />

<strong>Fibril</strong> protein<br />

(kDa)<br />

Variable region or<br />

fragment (5–23)<br />

(?)<br />

<strong>Amyloid</strong> A (5–10)<br />

A 2M 2-Microglobulin (12) 2-Microglobulin<br />

(12)<br />

ACys Cystatin C (13) Cystatin C<br />

fragment (12)<br />

ALys Lysozyme (14) Lysozyme (14)<br />

AFibA Fibrinogen (66) Fibrinogen <br />

fragment (11-P3)<br />

Endocrine<br />

hormones<br />

AIAPP Islet amyloid<br />

polypeptide (9)<br />

AANF Atrial natriuretic<br />

factor<br />

Islet amyloid<br />

polypeptide<br />

fragment (4)<br />

Atrial natriuretic<br />

factor fragment<br />

(4)<br />

ACal Procalcitonin Calcitonin<br />

fragment (3–4)<br />

APro Prolactin (?)<br />

AIns<br />

Transport<br />

molecules<br />

Insulin (porcine,<br />

iatrogenic (6)<br />

Insulin (6)<br />

ATTR Transthyretin (14) Transthyretin,<br />

intact/fragment<br />

(5–14)<br />

AApoAI Apolipoprotein AI Apolipoprotein AI<br />

(28)<br />

fragment (9–11)<br />

ALac<br />

Nervous system<br />

Lact<strong>of</strong>errin (?)<br />

A A protein precursor<br />

(110–135)<br />

A (4–5)<br />

APrP Prion protein Prion protein<br />

Cell motility<br />

(30–35)<br />

fragment (27–30)<br />

AGel Gelsolin (80–90) Gelsolin fragment<br />

(9–11)<br />

serum amyloid P component; heparan sulfate proteoglycans;<br />

<strong>and</strong> apolipoprotein E, which is associated<br />

with a number <strong>of</strong> types <strong>of</strong> amyloid, including<br />

<strong>the</strong> A deposits <strong>of</strong> Alzheimer’s disease.<br />

Although <strong>the</strong> early isolation <strong>of</strong> amyloid fibrils<br />

from tissues employed differential centrifugation,<br />

after Pras introduced <strong>the</strong> water extraction method,<br />

(Pras et al., 1968), <strong>the</strong> method has been almost universally<br />

used for extraction <strong>of</strong> almost all biochemical<br />

types <strong>of</strong> amyloid, except A <strong>and</strong> APrP (Selkoe <strong>and</strong><br />

Abraham, 1986; Prusiner <strong>and</strong> DeArmond, 1995).<br />

With <strong>the</strong> low ionic strength Pras method, amyloid<br />

REVIEW: HISTORY OF THE AMYLOID FIBRIL<br />

fibrils are separated from all proteins soluble in<br />

physiological saline by extraction <strong>of</strong> homogenates <strong>of</strong><br />

amyloid-laden tissues with physiological saline followed<br />

by differential centrifugation. Then, amyloid<br />

fibrils are separated from o<strong>the</strong>r saline-insoluble tissue<br />

proteins by suspension in water. The isolated<br />

protein fibrils, insoluble in physiologic saline, exhibit<br />

Congophilia <strong>and</strong> green birefringence <strong>and</strong> are<br />

80 to 100 Å in width.<br />

The insoluble protein fibrils, extracted from tissues<br />

laden with amyloid fibrils, can be solubilized<br />

<strong>and</strong> size fractionated by gel filtration in denaturing<br />

agents, such as guanidine hydrochloride. This technological<br />

advance made possible <strong>the</strong> biochemical era<br />

<strong>of</strong> amyloid studies. Amino acid sequence determination<br />

<strong>of</strong> solubilized, denatured amyloid fibril proteins<br />

quickly led to <strong>the</strong> recognition that a biochemically<br />

unique protein was associated with each <strong>of</strong> <strong>the</strong> clinical<br />

syndromes (Tables I <strong>and</strong> II). Since 1970, <strong>the</strong><br />

number <strong>of</strong> biochemically distinct proteins, isolated<br />

from species throughout <strong>the</strong> animal kingdom, that<br />

satisfy <strong>the</strong> three criteria for amyloid is approaching<br />

20 (Table II, Sipe, 1992).<br />

Studies from <strong>the</strong> Benditt <strong>and</strong> Glenner laboratories<br />

provided <strong>the</strong> first indication <strong>of</strong> <strong>the</strong> biochemical heterogeneity<br />

<strong>of</strong> amyloid. Glenner <strong>and</strong> co-workers<br />

(1971a) reported, on <strong>the</strong> basis <strong>of</strong> amino acid sequence<br />

analysis, that primary amyloidosis (now<br />

known as AL, but initially termed amyloid B by <strong>the</strong><br />

Benditt laboratory (Benditt, 1976)) is <strong>the</strong> result <strong>of</strong><br />

<strong>the</strong> deposition <strong>of</strong> fragments <strong>of</strong> immunoglobulin light<br />

chains. Benditt <strong>and</strong> colleagues identified, on <strong>the</strong> basis<br />

<strong>of</strong> amino acid sequence analysis <strong>of</strong> fibrils isolated<br />

from tissues <strong>of</strong> patients with chronic <strong>and</strong> recurrent<br />

acute inflammatory diseases (secondary amyloidosis),<br />

a protein termed amyloid A (Benditt et al.,<br />

1971). The existence <strong>of</strong> a third biochemical class <strong>of</strong><br />

amyloid protein was revealed when Costa et al.<br />

(1978) <strong>and</strong> Skinner <strong>and</strong> Cohen (1981) reported that<br />

prealbumin, now known as transthyretin, is <strong>the</strong><br />

fibril protein associated with familial amyloid polyneuropathy.<br />

Subsequent biochemical analyses, including<br />

amino acid sequencing, led to <strong>the</strong> identification<br />

<strong>of</strong> fragments (usually) <strong>of</strong> about 20 normally<br />

soluble proteins deposited in a specific anatomic pattern<br />

associated with specific clinical manifestation<br />

<strong>of</strong> disease (Table II). In some cases, only a single<br />

organ <strong>of</strong> <strong>the</strong> body is affected, such as <strong>the</strong> pancreas in<br />

diabetes or <strong>the</strong> brain in Alzheimer’s disease. In<br />

o<strong>the</strong>r cases, termed systemic amyloidoses, amyloid<br />

fibrils are deposited in multiple organs <strong>and</strong> tissues<br />

<strong>of</strong> <strong>the</strong> body <strong>and</strong> are thought to be derived from<br />

circulating plasma protein precursors (Sipe, 1992,<br />

1994).<br />

Today, amyloid is known to be a large biochemically<br />

<strong>and</strong> clinically heterogenous group <strong>of</strong> disorders<br />

95


96 REVIEW: SIPE AND COHEN<br />

<strong>of</strong> protein folding. Investigations <strong>of</strong> amyloid have led<br />

to <strong>the</strong> identification <strong>of</strong> previously unknown peptides<br />

<strong>and</strong> proteins implicated in a large number <strong>of</strong> important<br />

life processes. These proteins include <strong>the</strong> amyloid<br />

A protein <strong>and</strong> its precursor, <strong>the</strong> cytokine-regulated<br />

serum amyloid A (SAA); more than 70<br />

transthyretin (TTR) variants; <strong>the</strong> cerebral amyloid <br />

protein, A, <strong>and</strong> its precursor protein APP; <strong>the</strong><br />

previously unrecognized diabetes-associated islet<br />

amyloid polypeptide (IAPP); <strong>and</strong> prion proteins that<br />

cause scrapie <strong>and</strong> o<strong>the</strong>r neurodegenerative diseases<br />

(APrP) (Cohen, 1967; Glenner, 1980; Sipe, 1992;<br />

Westermark et al., 1996; Benson <strong>and</strong> Uemechi,<br />

1996; Prusiner <strong>and</strong> DeArmond, 1995).<br />

X-ray diffraction analyses <strong>of</strong> isolated amyloid protein<br />

fibrils (Bonar et al., 1969; Glenner et al., 1974;<br />

Sunde <strong>and</strong> Blake, 1998) revealed that all proteinaceous<br />

amyloid fibrils were ordered in secondary<br />

structure, with <strong>the</strong> polypeptide backbone assuming<br />

<strong>the</strong> beta pleated sheet conformation <strong>and</strong> oriented<br />

perpendicular to <strong>the</strong> fibril axis. It is particularly<br />

noteworthy that, in <strong>the</strong> case <strong>of</strong> TTR, which is<br />

present in <strong>the</strong> vitreous fluid <strong>of</strong> FAP patients as well<br />

as throughout <strong>the</strong> body, <strong>the</strong> structure <strong>of</strong> <strong>the</strong> TTR<br />

fibril from <strong>the</strong> vitreous (for which <strong>the</strong> Pras extraction<br />

method would not have to be used) <strong>and</strong> <strong>the</strong> TTR<br />

fibril extracted from kidney (for which <strong>the</strong> Pras<br />

method or a modification <strong>of</strong> it would be used) are<br />

identical (Inouye et al., 1998).<br />

The association <strong>of</strong> <strong>the</strong> pentraxin SAP with tissue<br />

amyloid deposits was revealed by electron microscopic<br />

studies <strong>of</strong> fractions <strong>of</strong> tissue homogenates<br />

that had been separated on <strong>the</strong> basis <strong>of</strong> solubility<br />

<strong>and</strong> sedimentation under conditions <strong>of</strong> decreasing<br />

ionic strength (reviewed by Cohen et al., 1982). It<br />

soon became apparent that, in addition to <strong>the</strong> fibrilrich<br />

fraction, ano<strong>the</strong>r, pentagonal, particle, composed<br />

<strong>of</strong> five identical globular subunits, was invariably<br />

associated with amyloid deposits in tissues. The<br />

pentagonal component (P component) was identified<br />

immunochemically by Cathcart <strong>and</strong> co-workers<br />

(1965) as an alpha globulin. Bladen <strong>and</strong> co-workers<br />

(1966), using a differential centrifugation technique,<br />

identified a small pentagonal structure about 90 Å<br />

in diameter (Fig. 4c). Subsequent studies (reviewed<br />

by Pepys et al., 1997) led to <strong>the</strong> identification <strong>of</strong> a<br />

circulating blood protein nearly if not identical to<br />

<strong>the</strong> tissue protein. That blood protein is now called<br />

serum amyloid P component or SAP.<br />

Despite <strong>the</strong> fact that SAP <strong>and</strong> heparan sulfate<br />

proteoglycans (HSPG) are always present in tissues<br />

with amyloid fibrils (Pepys et al., 1997; Snow <strong>and</strong><br />

Wight, 1989; Magnus <strong>and</strong> Stenstad, 1997), amyloid<br />

fibrils can be formed in vitro in <strong>the</strong>ir absence from<br />

several natural polypeptides, including insulin, glucagon,<br />

calcitonin, immunoglobulin light chains, <strong>and</strong><br />

2-microglobulin (reviewed in Sipe, 1994). <strong>Fibril</strong>s,<br />

similar in dimension <strong>and</strong> appearance to tissue-derived<br />

amyloid proteins, have also have been formed<br />

from syn<strong>the</strong>tic peptides corresponding to part or all<br />

<strong>of</strong> AA, A, <strong>and</strong> IAPP. <strong>Amyloid</strong>ogenic polypeptides<br />

can be influenced by heparin to assume <strong>the</strong> beta<br />

pleated sheet ra<strong>the</strong>r than <strong>the</strong> alpha helical conformation<br />

(McCubbin et al., 1988). Studies <strong>of</strong> syn<strong>the</strong>tic<br />

peptides led to <strong>the</strong> identification <strong>of</strong> amyloidogenic<br />

sequences in some precursors, such as AA, A, <strong>and</strong><br />

IAPP, but not in o<strong>the</strong>rs, such as AL <strong>and</strong> ATTR<br />

(reviewed in Sipe, 1994). Some amyloidogenic precursors,<br />

such as TTR <strong>and</strong> immunoglobulin light<br />

chains, are predisposed to fibril formation by structural<br />

changes (amino acid substitutions) in disparate<br />

parts <strong>of</strong> <strong>the</strong> polypeptide, changes that promote<br />

unfolding. O<strong>the</strong>rs, such as A, AA, <strong>and</strong> AIAPP precursor,<br />

contain contiguous regions <strong>of</strong> amino acids<br />

that predispose to amyloid fibril formation. Recently,<br />

amyloid fibrils were generated by acid <strong>and</strong><br />

heat denaturation <strong>of</strong> monellin, a sweet tasting plant<br />

protein (Konno et al., 1999). De novo amyloid proteins<br />

have been generated from designed combinatorial<br />

libraries, in which all sequences were designed<br />

to share an identical pattern <strong>of</strong> alternating<br />

polar <strong>and</strong> nonpolar residues, <strong>and</strong> <strong>the</strong> side chains<br />

were varied combinatorially (West et al., 1999). The<br />

resultant aggregates have fibrillar morphology, are<br />

Congophilic, <strong>and</strong> assume <strong>the</strong> beta pleated sheet secondary<br />

structure. The X-ray diffraction patterns <strong>of</strong><br />

amyloid fibrils prepared in vitro <strong>and</strong> those extracted<br />

from tissues are similar (Glenner et al., 1974). Apparently,<br />

<strong>the</strong> formation <strong>of</strong> proteinaceous amyloid<br />

fibrils is <strong>the</strong> result <strong>of</strong> a combination <strong>of</strong> factors, including<br />

<strong>the</strong> primary structure <strong>of</strong> <strong>the</strong> polypeptide <strong>and</strong><br />

<strong>the</strong> <strong>the</strong>rmodynamic parameters <strong>of</strong> <strong>the</strong> environment<br />

<strong>of</strong> <strong>the</strong> polypeptide (Kelly, 1998).<br />

<strong>Fibril</strong>s indistinguishable from those isolated from<br />

tissues were created by proteolytic digestion <strong>of</strong><br />

Bence-Jones immunoglobulin light chain proteins<br />

(Glenner et al., 1971b; Linke et al., 1973; Shirahama<br />

et al., 1973). The in vitro properties <strong>of</strong> fibrillar morphology,<br />

Congophilia, <strong>and</strong> green birefringence led to<br />

<strong>the</strong> conclusion that <strong>the</strong> AL fibrils were identical to<br />

those deposited in tissues in vivo. Thus, <strong>the</strong> in vitro<br />

observation <strong>of</strong> <strong>the</strong> properties <strong>of</strong> insolubility <strong>and</strong> resistance<br />

to proteolysis led to <strong>the</strong> assumption that<br />

amyloid fibrils exhibit <strong>the</strong> same properties in vivo<br />

<strong>and</strong> in vitro <strong>and</strong> that insolubility <strong>and</strong> resistance to<br />

proteolytic clearance lead to replacement <strong>and</strong> destruction<br />

<strong>of</strong> vital organs that are clinically affected<br />

with amyloidosis.<br />

CURRENT MODEL OF AMYLOID FIBRILS IN TISSUES<br />

As evidenced by o<strong>the</strong>r reports in this volume, we<br />

are currently entering a new molecular <strong>and</strong> <strong>the</strong>ra-


peutic era <strong>of</strong> amyloid, one in which prevention or<br />

reversal <strong>of</strong> amyloid fibril formation <strong>and</strong> <strong>the</strong> ensuing<br />

pathology in <strong>the</strong> body may be anticipated. Future<br />

multidisciplinary investigations <strong>of</strong> amyloid must be<br />

based upon a clear underst<strong>and</strong>ing <strong>of</strong> how <strong>the</strong> biophysical<br />

properties <strong>of</strong> proteins isolated from organs<br />

<strong>of</strong> patients afflicted with amyloidosis relate to <strong>the</strong><br />

actual secondary structure <strong>of</strong> <strong>the</strong>se fibrils in tissues.<br />

Inoue <strong>and</strong> colleagues (Inoue et al., 1998a) have<br />

compared experimental mouse AA amyloid fibrils in<br />

situ <strong>and</strong> after isolation by <strong>the</strong> Pras method (Pras et<br />

al., 1968). On <strong>the</strong> basis <strong>of</strong> <strong>the</strong>ir observation on AA<br />

amyloid <strong>and</strong> on o<strong>the</strong>r forms, including A, 2-microglobulin,<br />

<strong>and</strong> ATTR (Inoue et al., 1999, 1998b, 1997),<br />

<strong>the</strong>y have proposed an alternative to <strong>the</strong> protein<br />

only, beta pleated sheet, prot<strong>of</strong>ilament amyloid fibril<br />

structure. Inoue, in collaboration with Kisilevsky<br />

<strong>and</strong> co-workers, has suggested that <strong>the</strong> 80- to 100-Å<br />

fibrillar structure in tissue is different from <strong>the</strong><br />

structure <strong>of</strong> isolated protein in vitro. He proposes, on<br />

<strong>the</strong> basis <strong>of</strong> high-resolution electron microscopic<br />

studies, that in tissues, <strong>the</strong> amyloid fibril core is<br />

composed <strong>of</strong> helically wound 30-Å chondroitin sulfate<br />

proteoglycan (CSPG) double tracked structures<br />

enclosing a pentamer <strong>of</strong> amyloid P component (AP).<br />

Outside <strong>the</strong> CSPG are 45- to 50-Å heparan sulfate<br />

proteoglycans, to which AA protein filaments are<br />

attached. Presumably, <strong>the</strong> filaments are ordered to<br />

account for Congophilia <strong>and</strong> o<strong>the</strong>r specific histochemical<br />

staining, although <strong>the</strong> molecular basis for<br />

Congophilia is not clearly defined. Inoue <strong>and</strong> colleagues<br />

have proposed that <strong>the</strong> protein only fibril is<br />

formed artifactually as a result <strong>of</strong> <strong>the</strong> decreased<br />

ionic strength employed as part <strong>of</strong> <strong>the</strong> Pras extraction<br />

procedure. The Inoue model would consistently<br />

incorporate <strong>the</strong> reported HSPG binding sites in SAA<br />

(reviewed by Urieli-Shoval et al., 2000) <strong>and</strong> o<strong>the</strong>r<br />

amyloid proteins as well as <strong>the</strong> consistent changes<br />

in proteoglycan syn<strong>the</strong>sis that precede amyloidogenesis<br />

(Kisilevsky <strong>and</strong> Fraser, 1997).<br />

The structure <strong>of</strong> <strong>the</strong> amyloid fibril in tissues has<br />

not been as rigorously analyzed as <strong>the</strong> protein only<br />

model <strong>of</strong> <strong>the</strong> amyloid fibril that has been derived<br />

from <strong>the</strong> past 3 decades <strong>of</strong> in vitro studies. In particular,<br />

a revisitation <strong>of</strong> <strong>the</strong> practice <strong>of</strong> isolation <strong>of</strong><br />

amyloid fibrils from tissues, using <strong>the</strong> early “gentle<br />

amyloid fractionation techniques,” is probably warranted.<br />

Future biochemical investigations will be<br />

more fruitful if heparan sulfate proteoglycan <strong>and</strong> AP<br />

content are analyzed, in addition to <strong>the</strong> major fibrilassociated<br />

protein. In particular, comparison <strong>of</strong> TTR<br />

fibrils from vitreous fluid <strong>and</strong> organs such as kidney<br />

from FAP patients, where <strong>the</strong> composition <strong>of</strong> extracellular<br />

matrix <strong>of</strong> <strong>the</strong>se tissues <strong>and</strong> <strong>the</strong> fibril isolation<br />

procedures applied to <strong>the</strong>m are distinct (Inouye<br />

REVIEW: HISTORY OF THE AMYLOID FIBRIL<br />

et al., 1998), may greatly enhance our underst<strong>and</strong>ing<br />

<strong>of</strong> <strong>the</strong> nature <strong>of</strong> amyloid fibril structure in vivo.<br />

REFERENCES<br />

Benditt, E. P. (1976) The structure <strong>of</strong> amyloid protein AA <strong>and</strong><br />

evidence for a transmissible factor in <strong>the</strong> origin <strong>of</strong> amyloidosis,<br />

in Wegelius, O., <strong>and</strong> Pasternack, A. (Eds.), <strong>Amyloid</strong>osis, pp.<br />

323–337, Academic Press, London.<br />

Benditt, E. P., Eriksen, N., Hermodson, M. A., et al. (1971) The<br />

major proteins <strong>of</strong> human <strong>and</strong> monkey amyloid substance: Common<br />

properties including unusual N-terminal amino acid sequences,<br />

FEBS Lett. 19, 169–173.<br />

Benson, M. D., Uemichi, T. (1996) Transthyretin amyloidosis,<br />

<strong>Amyloid</strong>. Int. J. Exp. Clin. Invest. 3, 44–56.<br />

Bladen, H. A., Nylen, M. U., <strong>and</strong> Glenner, G. G. (1966) The<br />

ultrastructure <strong>of</strong> human amyloid as revealed by <strong>the</strong> negative<br />

staining technique, J. Ultrastruct. Res. 14, 449–459.<br />

Bonar, L., Cohen, A. S., <strong>and</strong> Skinner, M. M. (1969) Characterization<br />

<strong>of</strong> <strong>the</strong> amyloid fibril as a cross-beta protein, Proc. Soc. Exp.<br />

Biol. Med. 131, 1373–1375.<br />

Cathcart, E. S., Comerford, F. R., <strong>and</strong> Cohen, A. S. (1965) Immunologic<br />

studies on protein extracted from human secondary<br />

amyloid, N. Eng. J. Med. 273, 143–146.<br />

Cohen, A. S. (1967) Medical progress. <strong>Amyloid</strong>osis, N. Eng. J.<br />

Med. 277, 522–530.<br />

Cohen, A. S. (1986) General introduction <strong>and</strong> a brief history <strong>of</strong> <strong>the</strong><br />

amyloid fibril, in Marrink, J., <strong>and</strong> Van Rijswijk, M. H. (Eds.),<br />

<strong>Amyloid</strong>osis, pp. 3–19, Nijh<strong>of</strong>f, Dordrecht.<br />

Cohen, A. S., Shirahama, T., <strong>and</strong> Skinner, M. (1982) Electron<br />

Microscopy <strong>of</strong> <strong>Amyloid</strong>, in Harris, J. R., (Ed.), Electron Microscopy<br />

<strong>of</strong> Proteins, Vol. 3, pp. 165–205, Academic Press, New<br />

York.<br />

Cohen, A. S., <strong>and</strong> Calkins, E. (1959) Electron microscopic observation<br />

on a fibrous component in amyloid <strong>of</strong> diverse origins,<br />

Nature 183, 1202–1203.<br />

Cohen, A. S., <strong>and</strong> Calkins, E. (1964) Isolation <strong>of</strong> amyloid fibrils<br />

<strong>and</strong> study <strong>of</strong> effect <strong>of</strong> collagenase <strong>and</strong> hyaluronidase, J. Cell<br />

Biol. 21, 481–486.<br />

Costa, P. P., Figueira, A. S., <strong>and</strong> Bravo, F. R. (1978) <strong>Amyloid</strong> fibril<br />

protein related to prealbumin in familial amyloidotic polyneuropathy,<br />

Proc. Natl. Acad. Sci. USA 75, 4499–4503.<br />

Divry, P., <strong>and</strong> Florkin, M. (1927) Sur les proprietes optiques de<br />

l’amyloide, C. R. Soc. Biol. 97, 1808–1810.<br />

Glenner, G. G. (1980) <strong>Amyloid</strong> deposits <strong>and</strong> amyloidosis. The beta<br />

fibrilloses, N. Engl. J. Med. 302, 1283–1333.<br />

Glenner, G. G., Terry, W., Harada, M., Isersky, C., <strong>and</strong> Page, D.<br />

(1971a) <strong>Amyloid</strong> fibrils proteins: pro<strong>of</strong> <strong>of</strong> homology with immunoglobulin<br />

light chains by sequence analysis, Science 172,<br />

1150–1153.<br />

Glenner, G. G., Ein, D., Eanes, E. D., Bladen, H. A., Terry, W.,<br />

<strong>and</strong> Page, D. (1971b) Creation <strong>of</strong> “<strong>Amyloid</strong>” fibrils from Bence<br />

Jones proteins in vitro, Science 174, 712–714.<br />

Glenner, G. G., Eanes, E. D., Bladen, H. A., et al. (1974) Betapleated<br />

sheet fibrils A comparison <strong>of</strong> native amyloid with syn<strong>the</strong>tic<br />

protein fibrils, J. Histochem. Cytochem. 22, 1141–1158.<br />

Gueft, B., <strong>and</strong> Ghidoni, J. J. (1963) The site <strong>of</strong> formation <strong>and</strong><br />

ultrastructure <strong>of</strong> amyloid, Am. J. Pathol. 43, 837–854.<br />

Inoue, S., Kuroiwa, M., Tan, R., <strong>and</strong> Kisilevsky, R. (1998) A high<br />

resolution ultrastructural comparison <strong>of</strong> isolated <strong>and</strong> in situ<br />

murine amyloid fibrils, <strong>Amyloid</strong>. Int. J. Exp. Clin. Invest. 5,<br />

99–110.<br />

Inoue, S., Kuroiwa, M., <strong>and</strong> Kisilevsky, R. (1999) Basement mem-<br />

97


98 REVIEW: SIPE AND COHEN<br />

branes, micr<strong>of</strong>ibrils <strong>and</strong> beta amyloid fibrillogenesis in Alzheimer’s<br />

disease: High resolution ultrastructural findings, Brain<br />

Res. Rev. 29, 218–231.<br />

Inoue, S., Kuroiwa, M., Saraiva, M. J., et al. (1998a) Ultrastructure<br />

<strong>of</strong> familial amyloid polyneuropathy amyloid fibrils: Examination<br />

with high-resolution electron microscopy, J. Struct.<br />

Biol. 124, 1–12.<br />

Inoue, S., Kuroiwa, M., Ohashi, K., et al. (1997) Ultrastructural<br />

organization <strong>of</strong> hemodialysis-associated beta(2)-microglobulin<br />

amyloid fibrils, Kidney Int. 52, 1543–1549.<br />

Inouye, H., Domingues, F. S., Damas, A. M., Saraiva, M. J.,<br />

Lundgren, E., S<strong>and</strong>gren, O., <strong>and</strong> Kirschner, D. A. (1998b) Analysis<br />

<strong>of</strong> x-ray diffraction patterns from amyloid <strong>of</strong> biopsied vitreous<br />

humor <strong>and</strong> kidney <strong>of</strong> transthyretin (TTR) Met30 familial<br />

amyloidotic polyneuropathy (FAP) patients: Axially arrayed<br />

TTR monomers constitute <strong>the</strong> prot<strong>of</strong>ilament, <strong>Amyloid</strong> Int. J.<br />

Exp. Clin. Invest. 5, 163–174.<br />

Kelly, J. W. (1998) The environmental dependency <strong>of</strong> protein<br />

folding best explains prion <strong>and</strong> amyloid diseases, Proc. Natl.<br />

Acad. Sci. USA 95, 930–932.<br />

Kisilevsky, R., <strong>and</strong> Fraser, P. E. (1997) A beta amyloidogenesis:<br />

Unique, or variation on a systemic <strong>the</strong>me? J. Crit. Rev. Biochem.<br />

Mol. 32, 361–404.<br />

Konno, T., Murata, K., <strong>and</strong> Nagayama, K. (1999) <strong>Amyloid</strong>-like<br />

aggregates <strong>of</strong> a plant protein: A case <strong>of</strong> a sweet-tasting protein,<br />

monellin, FEBS Lett. 454, 122–126.<br />

Linke, R. P., Tischendorf, F. W., Zucker-Franklin, D., et al. (1973)<br />

The formation <strong>of</strong> amyloid-like fibrils in vitro from Bence Jones<br />

proteins <strong>of</strong> <strong>the</strong> lambda I subclass, J. Immunol. 111, 24–26.<br />

Magnus, J. H., <strong>and</strong> Stenstad, T. (1997) Proteoglycans <strong>and</strong> <strong>the</strong><br />

extracellular matrix in amyloidosis, <strong>Amyloid</strong>. Int. J. Exp. Clin.<br />

Invest. 4, 121–134.<br />

McCubbin, W. D., Kay, C. M., Narindrasorasak, S., <strong>and</strong> Kisilevsky,<br />

R. (1988) Circular dichroism <strong>and</strong> fluorescence studies<br />

on two murine serum amyloid A proteins, Biochem. J. 256,<br />

775–783.<br />

Missmahl, H. P., <strong>and</strong> Hartwig, M. (1953) Polarisationsoptische<br />

untersuchungen an der amyloidsubstance, Virchows. Arch.<br />

Pathol. Anat. 324, 489–508.<br />

Pepys, M. B., Booth, D. R., Hutchinson, W. L., Gallimore, J. R.,<br />

Collins, P. M., <strong>and</strong> Hohenester, E. (1997) <strong>Amyloid</strong> P component.<br />

A critical review, <strong>Amyloid</strong>. Int. J. Exp. Clin. Invest. 4,<br />

274–295.<br />

Pras, M., Schubert, M., Zucker-Franklin, D., et al. (1968) The<br />

characterization <strong>of</strong> soluble amyloid prepared in water, J. Clin.<br />

Invest. 47, 924–933.<br />

Prusiner, S. B., <strong>and</strong> DeArmond, S. J. (1995) Prion protein amyloid<br />

<strong>and</strong> neurodegeneration, <strong>Amyloid</strong> Int. J. Exp. Clin. Invest. 2,<br />

39–65.<br />

Selkoe, D. J., <strong>and</strong> Abraham, C. R. (1986) Isolation <strong>of</strong> paired<br />

helical filaments <strong>and</strong> amyloid fibers from human brain, Methods<br />

Enzymol. 134, 388–404.<br />

Shirahama, T., Benson, M. D., Cohen, A. S., <strong>and</strong> Tanaka, A.<br />

(1973) <strong>Fibril</strong>lar assemblage <strong>of</strong> variable segments <strong>of</strong> immunoglobulin<br />

light chains: An electron microscopic study, J. Immunol.<br />

110, 21–26.<br />

Shirahama, T., <strong>and</strong> Cohen, A. S. (1967) High resolution electron<br />

microscopic analysis <strong>of</strong> <strong>the</strong> amyloid fibril. J. Cell Biol. 33,<br />

679–708.<br />

Sipe, J. D. (1992) <strong>Amyloid</strong>osis, Annu. Rev. Biochem. 61, 947–975.<br />

Sipe, J. D. (1994) <strong>Amyloid</strong>osis, in Hindmarsh, J. T., <strong>and</strong> Goldberg,<br />

D. M. (Eds.), Critical <strong>Review</strong>s in Clinical Laboratory Sciences,<br />

Vol. 31, pp. 325–354. CRC Press, Boca Raton.<br />

Skinner, M., <strong>and</strong> Cohen, A. S. (1981) The prealbumin nature <strong>of</strong><br />

<strong>the</strong> amyloid protein in familial amyloid polyneuropathy (FAP),<br />

Biochem. Biophys. Res. Commun. 99, 1326–1332.<br />

Snow, A. D., <strong>and</strong> Wight, T. N. (1989) Proteoglycans in <strong>the</strong> pathogenesis<br />

<strong>of</strong> Alzheimer’s disease <strong>and</strong> o<strong>the</strong>r amyloidosis, Neurobiol.<br />

Aging 10, 481–497.<br />

Sunde, M., <strong>and</strong> Blake, C. C. F. (1998) From <strong>the</strong> globular to <strong>the</strong><br />

fibrous state: protein structure <strong>and</strong> structural conversion in<br />

amyloid formation, Q. Rev. Biophys. 31, 1–39.<br />

Terry, R. D., Nicholas, K. G., <strong>and</strong> Weiss, M. (1964) Ultrastructural<br />

studies in Alzheimer’s presenile dementia, Am. J. Pathol.<br />

44, 269–297.<br />

Urieli-Shoval, S., Linke, R. P., <strong>and</strong> Matzner, Y. (2000) Expression<br />

<strong>and</strong> function <strong>of</strong> serum amyloid A, a major acute phase protein,<br />

in normal <strong>and</strong> disease states, Curr. Opin. Hematol. 7, 64–69.<br />

West, M. W., Wang, W. X., Patterson, J., Mancias, J. D., Beasley,<br />

J. R., <strong>and</strong> Hecht, M. H. (1999) De novo amyloid proteins from<br />

designed combinatorial libraries, Proc. Natl. Acad. Sci. USA<br />

96, 11211–11216.<br />

Westermark, P., Sletten, K., <strong>and</strong> Johnson, K. H. (1996) Ageing<br />

<strong>and</strong> amyloid fibrillogenesis: Lessons from apolipoprotein AI,<br />

transthyretin <strong>and</strong> islet amyloid polypeptide, CIBA Found.<br />

Symp. 199, 205–222.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!