characterization of the i4399m variant of apolipoprotein(a

113
University of Windsor University of Windsor Scholarship at UWindsor Scholarship at UWindsor Electronic Theses and Dissertations Theses, Dissertations, and Major Papers 2016 Characterization of the I4399M Variant of Apolipoprotein(a): Characterization of the I4399M Variant of Apolipoprotein(a): Implications for Altered Prothrombotic Properties of Implications for Altered Prothrombotic Properties of Lipoprotein(a) Lipoprotein(a) Jackson McAiney University of Windsor Follow this and additional works at: https://scholar.uwindsor.ca/etd Recommended Citation Recommended Citation McAiney, Jackson, "Characterization of the I4399M Variant of Apolipoprotein(a): Implications for Altered Prothrombotic Properties of Lipoprotein(a)" (2016). Electronic Theses and Dissertations. 5848. https://scholar.uwindsor.ca/etd/5848 This online database contains the full-text of PhD dissertations and Masters’ theses of University of Windsor students from 1954 forward. These documents are made available for personal study and research purposes only, in accordance with the Canadian Copyright Act and the Creative Commons license—CC BY-NC-ND (Attribution, Non-Commercial, No Derivative Works). Under this license, works must always be attributed to the copyright holder (original author), cannot be used for any commercial purposes, and may not be altered. Any other use would require the permission of the copyright holder. Students may inquire about withdrawing their dissertation and/or thesis from this database. For additional inquiries, please contact the repository administrator via email ([email protected]) or by telephone at 519-253-3000ext. 3208.

Upload: others

Post on 16-May-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Characterization of the I4399M Variant of Apolipoprotein(a

University of Windsor University of Windsor

Scholarship at UWindsor Scholarship at UWindsor

Electronic Theses and Dissertations Theses, Dissertations, and Major Papers

2016

Characterization of the I4399M Variant of Apolipoprotein(a): Characterization of the I4399M Variant of Apolipoprotein(a):

Implications for Altered Prothrombotic Properties of Implications for Altered Prothrombotic Properties of

Lipoprotein(a) Lipoprotein(a)

Jackson McAiney University of Windsor

Follow this and additional works at: https://scholar.uwindsor.ca/etd

Recommended Citation Recommended Citation McAiney, Jackson, "Characterization of the I4399M Variant of Apolipoprotein(a): Implications for Altered Prothrombotic Properties of Lipoprotein(a)" (2016). Electronic Theses and Dissertations. 5848. https://scholar.uwindsor.ca/etd/5848

This online database contains the full-text of PhD dissertations and Masters’ theses of University of Windsor students from 1954 forward. These documents are made available for personal study and research purposes only, in accordance with the Canadian Copyright Act and the Creative Commons license—CC BY-NC-ND (Attribution, Non-Commercial, No Derivative Works). Under this license, works must always be attributed to the copyright holder (original author), cannot be used for any commercial purposes, and may not be altered. Any other use would require the permission of the copyright holder. Students may inquire about withdrawing their dissertation and/or thesis from this database. For additional inquiries, please contact the repository administrator via email ([email protected]) or by telephone at 519-253-3000ext. 3208.

Page 2: Characterization of the I4399M Variant of Apolipoprotein(a

Characterization of the I4399M Variant of Apolipoprotein(a):

Implications for Altered Prothrombotic Properties of Lipoprotein(a)

By

Jackson Thomas McAiney

A Thesis

Submitted to the Faculty of Graduate Studies

Through the Department of Chemistry and Biochemistry

in Partial Fulfillment of the Requirements for

the Degree of Master of Science at the

University of Windsor

Windsor, Ontario, Canada

© 2016 Jackson Thomas McAiney

Page 3: Characterization of the I4399M Variant of Apolipoprotein(a

Characterization of the I4399M variant of Apolipoprotein(a): Implications

for Altered Prothrombotic Properties of lipoprotein(a)

by

Jackson McAiney

APPROVED BY:

______________________________________________

A. Hubberstey

Department of Biological Sciences

______________________________________________

B. Mutus

Department of Chemistry and Biochemistry

______________________________________________

M. Koschinsky, Advisor

Department of Chemistry and Biochemistry

September 12, 2016

Page 4: Characterization of the I4399M Variant of Apolipoprotein(a

iii

Declaration of Co-Authorship

I hereby declare that this thesis incorporates material that is the result of joint research, as

follows:

The experiments presented herein were the work of the author under the guidance

and direction of Dr. Marlys Koschinsky and Dr. Michael Boffa. Figures 1.2, 1.3, and 1.4

were designed by and used with the permission of Dr. Marlys Koschinsky and Dr.

Michael Boffa.

The recombinant-apo(a) stable lines were provided by Dr. Marlys Koschinsky and

generated with the assistance of Corey Scipione, Ph. D. candidate at the University of

Windsor, as outlined in Chapter 2.

Confocal experiments were performed with the assistance of the Zainab Bazzi, Ph.

D. candidate at the University of Windsor. Fiber number measurements of patient sample

confocal micrographs were conducted through ImageJ, using a macro designed by Dr.

Fraser Macrae from the University of Leeds. The scanning electron microscopy

experiments were performed with the assistance of Corey Scipione and Sharon Lackey,

Facility Technician at ESEM Facility at the GLIER Institute at the University of Windsor.

Pulse-chase assay was optimized and performed with assistance of Matt Gemin,

Masters candidate at the University of Windsor, as outlined in Chapter 2.

Molecular dynamic simulations were performed by Dr. James Gauld, Associate

Professor at the University of Windsor, and his Master’s student, Dan Simard. Mass

spectroscopy spectra were generated by Corey Scipione with the assistance of Janeen

Auld, Mass Spectroscopy Instrument Specialist at the University of Windsor.

Page 5: Characterization of the I4399M Variant of Apolipoprotein(a

iv

I am aware of the University of Windsor Senate Policy on Authorship and I

certify that I have properly acknowledged the contribution of other researchers to my

thesis, and have obtained written permission from each of the co-author(s) to include the

above material(s) in my thesis.

I certify that, with the above qualification, this thesis, and the research to which it

refers, is the product of my own work.

Page 6: Characterization of the I4399M Variant of Apolipoprotein(a

v

Abstract

Elevated plasma levels of lipoprotein(a) (Lp(a)) are a causal risk factor for CHD.

Lp(a) closely resembles LDL, but contains an additional glycoprotein apolipoprotein(a)

(apo(a)) that is structurally homologous to the fibrinolytic proenzyme plasminogen. This

has led to speculation that Lp(a) can oppose the fibrinolytic functions of plasminogen. A

single nucleotide polymorphism (SNP) in the LPA gene encoding apo(a) results in an Ile

to Met substitution at position 4399 in the protease-like domain. In population studies,

this variant has been correlated with elevated plasma Lp(a) levels and with higher CHD

risk. We undertook a functional characterization of the effect of the I4399M substitution

in apo(a). Molecular dynamics simulations of wild-type (wt) apo(a) and the Met variant

revealed a shift from a buried (Ile) to slightly exposed (Met) environment. Subsequent

MALDI-TOF mass spectrometry analysis demonstrated the presence of a methionine

sulfoxide moiety at this position in the Met variant. When 17-kringle recombinant forms

of apo(a) were included in a plasma clot lysis assay, both the wt apo(a) and Met variant

inhibited lysis, but the Met variant had a 50% greater effect. However, the Met variant

was equally as efficient as wt apo(a) in inhibiting plasminogen activation on a fibrin

surface. The Met variant was also able to significantly shorten coagulation time and result

in greater turbidity for clots made from either purified fibrin or lipoprotein-deficient

plasma compared to wild-type. In agreement with these findings, SEM and confocal

microscopy of clots made from fibrin and citrated plasma showed that compared to wt

apo(a), the Met variant resulted in significant alteration of the fibrin network. Together,

our data suggest that the Met4399 variant differs structurally from wt apo(a), which may

underlie key differences related to its effects on fibrin clot architecture and fibrinolysis.

Page 7: Characterization of the I4399M Variant of Apolipoprotein(a

vi

Acknowledgements

First and foremost, I would like to most sincerely thank my supervisor Dr. Marlys

Koschinsky for her patience, her support, her feedback, and mentorship during the

challenges and achievements of my graduate studies. The last two years have been a very

rewarding and incredible learning experience and I am very grateful for the opportunity to

be able to work in your research laboratory.

I would also like to extend my thanks Dr. Michael Boffa for his assistance in

optimizing the various plate based assays, as well as his guidance and insightful advice

during my studies.

Special thanks to Dr. Gauld for producing the molecular dynamic simulations, Dr.

Macrae for allowing us to use his designed macro, and Sharon Lacke for her help with

SEM imaging.

I would like to thank Corey Scipione for his mentorship during my undergraduate

and graduate studies. He was always available for support and guidance, to help with

various experiments, and I am very grateful to have had the opportunity to work with

him. I would also like to thank all members of the MLK and MBB labs for countless

enjoyable moments and support during my studies.

Finally, I would like to thank my mother, father, and brother for their ongoing

love, support, and dedication throughout the pursuit of all of my goals, including the

completion of my studies and this thesis.

Page 8: Characterization of the I4399M Variant of Apolipoprotein(a

vii

Table of Contents

Declaration of Co-Authorship…………………………………………………………..iii

Abstract………………………………………………………………………………….v

Acknowledgements……………………………………………………………….…..…vi

List of Figures……………………………………………………………………………x

Abbreviations……………………………………………………………………………xii

Chapter 1: Introduction…………………………………………………………………..1

1.1 Relationship Between Atherosclerosis and Thrombosis

in Atherosclerotic Plaque Formation…………………………………….1

1.2 Fibrin Clot Formation/Lysis and Plasminogen Structure

and Function……………………………………………………………...2

1.3 Fibrinogen Structure and Conversion to Fibrin………………………………5

1.4 Lipoprotein(a): A Potential Link Between the Processes of

Atherosclerosis and Thrombosis…………………………………………6

1.5 Evolution and Characteristics of LPA Gene…………………………………12

1.6 Plasma Lipoprotein(a) Levels - Relationship With Apo(a)

Isoform Size……………………………………………………………...13

1.7 Transcriptional Regulation of LPA…………………………………………..14

1.8 Identification of Two SNPs that Contribute Significantly to

Plasma Lp(a) Levels and Risk…………………………………………...15

1.9 Pathogenicity of Lipoprotein(a)……………………………………………...15

1.10 Inhibition of Plasminogen Activation and Fibrinolysis by

Lp(a)/apo(a)………………………………………………………………. 18

Page 9: Characterization of the I4399M Variant of Apolipoprotein(a

viii

1.11 Lipoprotein(a) and Oxidation……………………………………………..19

1.12 LPA rs3798220 SNP and Thrombotic Risk…………………………….....20

1.13 Objectives………………………………………………………………... 21

1.14 Hypothesis………………………………………………………………...22

Chapter 2: Methods…………………………………………………………………….23

2.1 Cell Culture…………………………………………………………………23

2.1.1 Construction and expression of recombinant apo(a) variants…….23

2.1.2 Propagation of cell lines…………………………………………..23

2.1.3 Purification of recombinant apo(a) variants………………………24

2.2 Characterization of lipoprotein(a) in human plasma samples………………25

2.3 Lipoprotein-Deficient Plasma………………………………………………26

2.4 Plasminogen activation on fibrin or cell surfaces…………………………..26

2.5 Assessment of plasma clot permeability……………………………………28

2.6 In Vitro fibrinolysis and coagulation assay…………………………………29

2.7 Analyzing the interaction of I4399M apo(a) with soluble fibrinogen………29

2.8 Imaging of fibrin clots with Confocal and Scanning Electron

Microscopy………………………………………………………………30

2.9 Pulse Chase Secretion Assay……………………………………………...…31

2.10 Molecular dynamics simulations and protease domain modeling………….32

2.11 Detection of methionine sulfoxide residue on KIV10-PI4399M………..…..33

2.12 Statistical methods……………………………………………………….....33

Chapter 3: Results…………………………………………………………………….…34

3.1 Structural Differences in Apo(a) Variants Predicted by Molecular

Dynamic Simulations……………………………………………………34

Page 10: Characterization of the I4399M Variant of Apolipoprotein(a

ix

3.2 Secretion of Wild-Type and Mutated r-apo(a) Variants from HepG2 cells…36

3.3 Effect of Mutant I4399M Variant on Prothrombotic Mechanisms

Compared to Wild-Type Apo(a)………………………………………...40

3.4 The Interaction Between the I4399M Variant and Fibrinogen/Fibrin Clot

Architecture……………………………………………………………...47

3.5 Analysis of Clot Structure and Properties of Carrier Patient Plasma

Samples………………………………………………………………….50

3.6 Identification of a Modification on the Substituted Methionine Residue in

the I4399M Variant……………………………………………………...57

Chapter 4: Discussion……………………………………………………………………59

4.1 The I4399M Variant Secretion Rate Does Not Differ from Wild-Type……..59

4.2 Analysis of the I4399M Variant Effect on Plasminogen Activation on

Different Surfaces………………………………………………………..60

4.3 I4399M Variant Alters Fibrinolysis Time and Coagulation Compared to

Wild-Type………………………………………………………………..62

4.4 The I4399M Alters Fibrin Clot Architecture and Causes Precipitation of

Fibrinogen………………………………………………………………..64

4.5 Lp(a) from Homozygous Carriers of the SNP Affect Fibrin Clot and Its

Properties Similar to I4399M r-apo(a) Variant…………………………..66

4.6 The I4399M Variant possesses a Methionine Sulfoxide Moiety which may

Contribute to the Prothrombotic Potential of this Variant……………….70

4.7 Conclusions………………………………………………………………….71

References……………………………………………………………………………….73

Vita Auctoris………………………………………………………………………….....98

Page 11: Characterization of the I4399M Variant of Apolipoprotein(a

x

List of Figures

Figure 1.1. Schematic of fibrin clot formation from the fibrinogen zymogen………...…7

Figure 1.2. Structure of Lipoprotein(a)…………………………………………..………10

Figure 1.3. Schematic of the structural homology between the domains of

plasminogen and apolipoprotein(a)……………………………………………....11

Figure 1.4. Potential pathogenic mechanisms of Lipoprotein(a)………………………...17

Figure 3.1. Molecular dynamic simulations of wild-type of mutant apo(a) variants…....35

Figure 3.2 Schematic representation of the recombinant apo(a) variants utilized

in the proposed studies………………………………………………………..….38

Figure 3.3 Pulse-chase analysis of the secretion of 17K versus 17K I4399M from HepG2

cells via pulse-chase……………………………………………………………...39

Figure 3.4 tPA-mediated plasminogen activation assay on a fibrin surface in

presence of r-apo(a) variants……………………………………………………..42

Figure 3.5 Plasminogen activation assay on the surface of THP-1 monocytes………….43

Figure 3.6 Absorbance-based fibrinolysis assay with LPDP in presence and absence

of apo(a)..…………………………………………………………………….…..45

Figure 3.7 Absorbance-based coagulation assay with LPDP in presence and absence

of apo(a)..………………………………………………………………………...46

Figure 3.8 SEM imaging of fibrin clots formed in absence or presence of either I4399M

or wild-type apo(a)….…………………………………………………………………....47

Figure 3.9 Imperial stain of the samples of I4399M and wild-type apo(a) variants in

solution with fibrinogen……………………………………………………….………....49

Figure 3.10 Permeation and fibrinolysis values of patient plasma samples……………...52

Page 12: Characterization of the I4399M Variant of Apolipoprotein(a

xi

Figure 3.11 Confocal microscopy and analysis of clot formation using homozygous and

non-carrier patient samples………………………………………………………………55

Figure 3.12 Scanning electron microscopy and fiber width measurements in clots formed

from homozygote and non-carrier patient samples………………………………………56

Figure 3.13 Mass spectroscopy of truncated apo(a) variant containing the I4399M

mutation…………………………………………………………………………..58

Page 13: Characterization of the I4399M Variant of Apolipoprotein(a

xii

Abbreviations

10-P truncated recombinant apo(a) variant containing KIV10, KV, and

protease domain

10-P I4399M 10-P r-apo(a) variant possessing Ile to Met mutation

17K wild type, 17 kringle-containing recombinant form of apo(a)

17K I4399M 17K recombinant apo(a) variant with Ile to Met mutation

ε-ACA epsilon-aminoaprotic acid

apo(a) apolipoprotein(a)

apo B-100 apolipoprotein B-100

CAD coronary artery disease

CVD cardiovascular disease

Da Daltons

DMEM Dulbecco’s modified eagle medium

EC endothelial cells

ECM extracellular matrix

EDTA ethylenediaminetetraacetic acid

ELISA enzyme-linked immunosorbent assay

FBS fetal bovine serum

FDP fibrin degradation products

Fg fibrinogen

FPA fibrinogen fibrinopeptide A

FXR farnesoid X receptor

Page 14: Characterization of the I4399M Variant of Apolipoprotein(a

xiii

FPB fibrinogen fibrinopeptide B

HBS HEPES buffered saline

HNF4α hepatocyte nuclear factor 4α

HEK293 human embryonic kidney cells

HepG2 human hepatoma cells

K kringle domain

Ks Darcy constant (units: cm2)

IIa activated thrombin

LBS lysine binding site

LDL low-density lipoprotein

LDLR low-density lipoprotein receptor

Lp(a) lipoprotein(a)

LPA gene encoding apo(a)

LPDP lipoprotein-deficient plasma

MEM minimal essential media

Opti-MEM serum free, conditioned media

OxPLs oxidized phospholipids

PAI-1 plasminogen activator inhibitor-1

PBS phosphate buffered saline

PDGF platelet-derived growth factor

Pg Plasminogen

r-apo(a) recombinant apo(a)

RMSD root mean squared deviation

SDS-PAGE sodium dodecyl sulfate polyacrylamide gel electrophoresis

Page 15: Characterization of the I4399M Variant of Apolipoprotein(a

xiv

SMC smooth muscle cells

SNP single nucleotide polymorphism

TFPI tissue factor pathway inhibitor

TGF-β transforming growth factor-β

tPA tissue-type plasminogen activator

uPA urokinase-type plasminogen activator

VCAM-1 vascular-cell adhesion molecule-1

VTE venous thromboembolism

WBC white blood cells

Page 16: Characterization of the I4399M Variant of Apolipoprotein(a

1

Chapter 1

Introduction

1.1 Relationship Between Atherosclerosis and Thrombosis in Atherosclerotic Plaque

Formation

Atherosclerosis a progressive, inflammatory disease affecting the large arteries of the

body and is the primary underlying cause of cardiovascular diseases (CVD), including

coronary artery disease (CAD), ischemic stroke, and heart failure (1-4). According to the

World Health Organization, over 17 million people die each year from CVDs and 80% of

all CVD deaths are due to myocardial infarctions and strokes (5). Atherosclerosis is

characterized as the thickening of the arterial intima due to the build up of lipids and

fibrous materials, resulting in the formation of atherosclerotic lesions within vasculature

(6).

Atherogenesis begins at the site of endothelial damage within an artery, caused by

factors such as oxidative stress, environmental irritants like tobacco, bacterial or viral

infection, and turbulent blood flow or shear stress on the arterial wall (7). Atherogenesis

is stimulated by the invasion of lipid containing molecules, such as low-density

lipoprotein (LDL), into the intimal layer (8). Movement of these lipid-containing

molecules across the endothelial cell layer triggers cellular activation, resulting in the

production of cell surface adhesion molecules on the ECs, including vascular-cell

adhesion molecule-1 (VCAM-1) (6). These receptors promote the recruitment of

lymphocytes and monocytes to endothelial cells at the inflammatory site. Recruitment of

these WBCs produces chemokines and growth factors within the intima, allowing the

Page 17: Characterization of the I4399M Variant of Apolipoprotein(a

2

monocytes to transmigrate into the intimal space and differentiate into inflammatory

macrophages (8). The accumulated LDL in the intima is exposed to oxidative and

enzymatic modification; subsequent phagocytosis of these molecules by the macrophages

results in the formation of “foam cells” (6, 9). Differentiation into foam cells results in

ER stress and apoptosis of the macrophage, resulting in cytokine release and increased

recruitment of monocytes to the lesion, increasing the size of the plaque (10). As the

lesion grows in size, circulating platelets deposit at the site of injury and release platelet-

derived growth factor (PDGF) and other mitogenic factors that penetrate the vascular wall

(11). These factors cause smooth muscle cells (SMC) to proliferate and migrate into the

intima, producing an extracellular matrix (ECM) that encloses the necrotic core of lesion,

consisting of lipid components and necrotic debris (6). The smooth muscle cells and

macrophages that constitute the fibrous connective tissue enclosing the necrotic core is

referred to as the “fibrous cap” (12).The lesions will continue to increase in size and may

rupture, releasing their contents into the lumen, causing arterial blockage than can lead to

myocardial infarctions/stroke (13). Enzymes and cytokines, produced by immune cells

within the core, weaken the fibrous cap surrounding the core and render the plaque

vulnerable to rupture. Upon rupture, tissue factor and platelet-adhesion factor within the

intima come into contact with the flowing blood in the arterial lumen, which results in

immediate thrombosis formation (8).

1.2 Fibrin Clot Formation/Lysis and Plasminogen Structure and Function

Fibrinogen is a ~340kDa zymogen that plays a role in blood clotting, fibrinolysis,

inflammatory response, and wound healing. Fibrinogen is composed of three homologous

peptide chains (Aα, Bβ, and γ) that are encoded in a cluster in a region of approximately

Page 18: Characterization of the I4399M Variant of Apolipoprotein(a

3

50kb on chromosome 4 and is synthesized in hepatocytes to form a hexamer composed of

two sets of the three polypeptides (14). In response to injury or vascular disease,

fibrinogen is cleaved by the active enzyme thrombin (lla), exposing binding sites to form

insoluble fibrin clots (15). When the fibrinogen is cleaved, fibrin fibers begin to assemble

into half-staggered double-stranded structures called protofibrils and will continue to

grow by associating laterally with other protofibrils (16, 17), forming a network of fibrin

fibers. Multiple factors affect the strength and stability of fibrin clots and various arterial

and venous diseases have been shown to alter the fibrin network, leading to altered

fibrinolysis and clot stability (18, 19).

There is an important balance between clot formation in the body and subsequent

lysis to prevent blockage of vasculature. The protein responsible for breakdown of fibrin

clots is called plasminogen. Plasminogen (Pg) is a single-chained circulating zymogen

that becomes cleaved to form the activated serine protease plasmin by the physiological

activators: urokinase-type Pg activator (uPA) or tissue-type Pg activator (tPA) (20).

While uPA is immobilized on the surface of endothelial cells, the circulating tPA is

synthesized in endothelial cells of the vascular wall and is secreted as a single chain

polypeptide, released in response to numerous stimuli (21). One such stimuli is thrombin,

the serine protease responsible for the conversion of soluble fibrinogen to insoluble fibrin

to form a clot (22). This conversion by thrombin causes a conformational change in

fibrin’s structure, exposing tPA and plasminogen binding sites (23, 24). A tertiary

complex is formed between tPA, plasminogen, and fibrin at these locations, resulting in

efficient plasminogen activation (25). Similarly, immobilization of plasminogen on cell

surfaces allows for more efficient activation by tPA or uPA (26). The primary in vivo

Page 19: Characterization of the I4399M Variant of Apolipoprotein(a

4

function of activated plasmin is to degrade the fibrin networks into individual components

called fibrin degradation products (FDP).

Plasminogen is composed of 791 amino acids, arranged into a N-terminal pre-

activation peptide, 5 tandem kringle domains (KI-KV), and an active protease domain

containing the catalytic triad His603, Asp646, and Ser741 (20). The kringle domains are

composed of a triple looped structure, held by 3 disulfide bonds, and are found in many

other proteins, particularly those associated with fibrinolysis and blood coagulation (27,

28). Kringles I, IV and V in plasminogen are characterized as possessing lysine-binding

sites (LBS), which are clefts or pockets within the larger domain that allow for protein-

protein interactions. The LBS possesses cationic and anionic residues at opposite ends

deep within the cleft, with a region of highly conserved aromatic residues that creates a

hydrophobic environment (29). These cationic and anionic residues within the cleft form

ion pair interactions with the carboxyl and ammonium groups of the ω-amino acid, such

as lysine (30). Plasminogen is cleaved to form the active plasmin by a cleavage of the

Arg561-Val562 peptide bond in the C-terminal domain by plasminogen activators.

Following cleavage, the heavy chain of plasminogen, containing the kringle domains, and

the light chain, containing the catalytic domain, through two disulfide bonds (31). The

activated plasminogen then subsequently degrades the fibrin clots into fibrin degradation

products (32).

There are two main forms of plasminogen: the most abundant type is called Glu1-

Pg, which is resistant to activation due to its tight, spiraled structure that blocks tPA from

reaching the activation cleavage site (33, 34). The second type is the result of plasmin

cleavage of the Lys77-Lys78 peptide bond in native Glu1-Pg, releasing the N-terminal

activation peptide and forms Lys78-Pg. Release of the N-terminal activation peptide

Page 20: Characterization of the I4399M Variant of Apolipoprotein(a

5

results in a conformational change that exposes the activation cleavage site, making the

Lys78-Pg a better substrate for tPA and accelerating plasminogen activation. Active

plasmin can also cleave the Arg275-Ile276 bond of the single chain tPA, resulting in a

double chained version, which has increased plasminogen activation ability (35).

Likewise, initial cleavage of fibrin by plasmin reveals C-terminal lysine residues that act

like high affinity binding sites for plasminogen and tPA (36). The LBS in the kringle

domains of plasminogen can interact with these exposed lysine residues, meaning this

modified fibrin is a better cofactor for tPA-mediated plasminogen activation.

1.3 Fibrinogen Structure and Conversion to Fibrin

A fibrinogen molecule is an elongated 45nm structure consisting of two outer D

domains connected to a central E domain connected by a coiled-coil segment (37). It is

composed of two sets of three polypeptide chains, termed Aα, Bβ, and γ, which are joined

together at the N-terminal E domain with 5 symmetrical disulfide bridges, while non-

symmetrical disulfide bridges form a ring structure in this area (38-40). The Aα, Bβ, and

γ chains are 610, 461, and 411 residues long, respectively, with the γ chain possessing a

minor and major form arising from alternative splicing of the mRNA transcript (37). Each

Aα subunit contains a N-terminal fibrinopeptide A (FPA) that is cleaved by the enzyme

thrombin (lla), exposing a polymerization site, termed EA, that results in a structure called

des-AA fibrin (Fig 1.1) (41-43). Each EA site combines with a complementary-binding

pocket on neighboring fibrinogen molecules in the D domain, termed Da, located on the γ

chain between the 337 and 379 residues (44, 45). The interaction of EA and DA results in

the alignment of fibrin molecules in a staggered, overlapping arrangement, forming a

double-stranded, twisted structure called a fibrin protofibril (16, 17, 46). To strengthen

Page 21: Characterization of the I4399M Variant of Apolipoprotein(a

6

this structure, these protofibrils undergo lateral associations, allowing the creation of

multi-stranded fibers (47, 48). These protofibrils can form two different types of junctions

with other fibrils; 1) laterally with another fibril to form a four-stranded fibril, called a

“bilateral” junction, and 2) “equilateral” junctions, formed between the three fibrin

molecules, giving rise to three double-stranded fibrils (49). Equilateral junctions occur

more frequently when FPA cleavage is slower, resulting in more branches and a tighter

fibrin matrix (50). Similarly, another fibrinopeptide, fibrinopeptide B (FPB), is slower

than the release of FPA and results in des-BB fibrin (41-43). Cleavage by thrombin

exposes an independent polymerization site, termed EB, that interacts with a

complementary Db site in the β-chain segment of the D domain (51). Polymerization of

des-BB fibrin results in the same type of protofibril structure as with des-AA fibrin (50),

but the clot strength is much weaker than that of des-AA fibrin (51). The resulting clot

architecture plays an important role in its stability and lysis potential; any abnormalities

can lead to disrupted thrombolysis (52, 53). Two of the most common measures of fibrin

clot structure are fibrin diameter and pore sizes within the clot network. Dense clots are

characterized as possessing smaller diameter fibers with increased stiffness, as well as

small pores, leading to increased clot lysis times (54). It has been shown that small

diameter fibers are dissolved at faster rates than thicker fibers, suggesting that the rate of

lysis is largely a function of pore size (53).

Page 22: Characterization of the I4399M Variant of Apolipoprotein(a

7

Figure 1.1. Schematic of fibrin clot formation from the fibrinogen zymogen. Fibrinogen molecules are cleaved by thrombin (lla), resulting in the release of

fibrinopeptides FPA and FPB, forming des-AA or des-BB fibrin monomers. Fibrin

monomers align in a staggered, over-lapping pattern and cross-link to form clusters of

monomers called protofibrils. The newly formed protofibrils form junctions to other

fibrils and the collection of these cross-linked fibrils results in an insoluble fibrin clot.

[Modified from Mega, JL (55)]

Page 23: Characterization of the I4399M Variant of Apolipoprotein(a

8

1.4 Lipoprotein(a): A Potential Link Between the Processes of Atherosclerosis and

Thrombosis

Lipoprotein(a) [Lp(a)] was discovered by Kåre Berg in 1963 (56). It is a unique

lipoprotein that circulates in human plasma. It is similar to LDL in terms of lipid content,

containing about 45% cholesterol, and the presence of apolipoprotein B-100

(apoB-100), but differs by the presence of a unique glycoprotein moiety called

apolipoprotein(a) [apo(a)] (57). Apo(a) is covalently attached to the apoB-100 component

of LDL by a single disulfide bond (Fig 1.2). Apo(a) differs from other apolipoproteins in

that it contains extensive carbohydrate modification, and is hydrophilic, lacking

amphipathic alpha helices typically found in other apolipoproteins (58, 59). The heavy

glycosylation is approximately 30-35% by weight and consists of N- and O-linked

glycans attached to free serine, threonine or asparagine residues in the interkringle

regions of apo(a). Cloning of apo(a) from human liver cDNA library revealed a striking

homology between the amino acid sequence of apo(a) and the human serine protease

zymogen plasminogen (60). This is an interesting result as plasminogen is responsible for

the degradation of insoluble fibrin clots, indicating Lp(a)/apo(a) may exerts its pathogenic

effects by interfering with plasminogen’s normal functions.

Apo(a) is structurally homologous to plasminogen in that it possesses domains

resembling the plasminogen kringle IV, kringle V, and protease domain, the last of which

is inactive in apo(a) due to a series of mutations (Fig 1.3). Apo(a) possesses 10 subclasses

of plasminogen-like kringle IV domains, denoted KIV1 through KIV10, that differ from

each other by changes in amino acid sequence (60). These domains in apo(a) have been

found to exhibit between 75-84% amino acid identity with the kringle IV domain in

plasminogen. Similar to plasminogen, the kringle domains in apo(a) exhibit the

Page 24: Characterization of the I4399M Variant of Apolipoprotein(a

9

characteristic tri-looped structure, composed of 80 amino acids held together by three

conserved disulphide bonds (61). The kringle domains are joined by a sequence of

approximately 30 amino acids, referred to as the interkringle region. This region may

provide flexibility to the quaternary structure of apo(a) (62). The plasminogen-like

kringle KIV1, and KIV3-KIV10 are each present in one copy in apo(a) (63). However, the

KIV2 domain is present in a varying number of identical copies, ranging from 3 to >40

repeats (64, 65). The variable number of KIV2 repeats results in Lp(a) isoform size

heterogeneity, a result of allelic variation in the number of repeated sequences encoding

the KIV2 domain in the LPA gene (66). Several of the other plasminogen-like kringle IV

domains in apo(a) are also of interest in terms of their properties. KIV5-KIV8 domains

each possess weak lysine binding sites (LBS); the LBS in KIV7 and KIV8 are thought to

mediate the initial non-covalent interactions between apo(a) and the N-terminal lysine

resides of apoB-100 (Lys680 and Lys690) prior to covalent Lp(a) assembly (67-69) (Figure

1.3). The plasminogen-like KIV9 domain in apo(a) contains a free cysteine which forms a

covalent linkage with the apoB-100 of LDL. However, the identity of the extracellular

oxidase enzyme that has been postulated to catalyze this reaction has yet to be determined

(57, 70). The plasminogen-like KIV10 domain possesses a stronger LBS comparable to

the one in plasminogen KIV and is thought to mediate interactions with a key biological

substrates including fibrin (71, 72), which has been shown to influence plasminogen

activation on the fibrin surface (71). The plasminogen-like protease domain in apo(a) is

catalytically inactive due to an arginine to serine substitution in apo(a) that removes the

cleavage site that is recognized by plasminogen activators (73).

Page 25: Characterization of the I4399M Variant of Apolipoprotein(a

10

Figure 1.2. Structure of Lipoprotein(a). Lp(a) is composed of an LDL-like moiety

attached to the heavily glycosylated, unique protein apolipoprotein(a) through a single

disulfide bond with apoB-100. There are also lysine dependent, non-covalent interactions

between the N-terminal lysine residues of apoB-100 of LDL and the lysine binding sites

(LBS) of the KIV7-KIV8 domains of apolipoprotein(a) that play a role in Lp(a) particle

assembly. Analysis of a liver cDNA library by McLean and colleagues (60) revealed a

striking homology between apo(a) and the serine protease zymogen, plasminogen.

[Adapted from Koschinsky ML and Marcovina SM (74)]

Page 26: Characterization of the I4399M Variant of Apolipoprotein(a

11

Figure 1.3. Schematic of the structural homology between the domains of

plasminogen and apolipoprotein(a). Plasminogen consists of an amino-terminal tail

domain (T), five kringle domains (numbered sequentially from I through V) and a serine

protease domain (P). Apo(a) possesses 10 copies of sequences that resembles the

plasminogen KIV domain, sharing between 75-84% amino acid identity, followed by a

single copy of sequences resembling the plasminogen KV domain and protease domain.

Each KIV domain is present in a single copy, except for KIV2, which can be repeated

between 3 and >40 times, accounting for the size heterogeneity of Lp(a). KIV5-KIV8

containing weak lysine binding sites (LBS), and the LBS on KIV7 and KIV8 mediate the

interaction with apoB-100 on LDL prior to Lp(a) assembly. KIV9 possesses a free

cysteine residue that is involved in the covalent linkage between apo(a) and apoB-100 on

LDL. KIV10 possesses a strong LBS and has been shown to interact with biological

substrates, such as fibrin. Unlike in plasminogen, the protease domain in apo(a) contains

critical amino acid substitutions that renders the catalytic domain inactive. [Adapted from

Koschinsky ML (59)]

Page 27: Characterization of the I4399M Variant of Apolipoprotein(a

12

The inactive protease domain in apo(a) possesses a nine amino acid deletion within a

highly conserved region in serine proteases, indicating that even in the absence of

ArgSer substitution, the domain would still be inactive (73).

1.5 Evolution and Characteristics of LPA Gene

Lp(a) is present in Old World monkeys, apes, and humans. Previous comparison

of the untranslated regions of human apo(a) and plasminogen cDNA sequences reveal

that the two genes arose from a duplication event about 40 million years ago, roughly

around the time of the divergence of the Old World and New World monkeys (60).

Therefore, it was proposed that the apo(a) gene arose from a duplication of the

plasminogen gene, followed by exon deletions, multiplication, and single base

substitutions, such as the ones present in the plasminogen-like protease domain.

Interestingly, however, an Lp(a)-like particle was found in European hedgehogs (75),

which suggests either that the LPA gene arose before the major divergence of the Old

World apes, or that the apo(a) in hedgehogs arose from an independent duplication of the

plasminogen gene that is distinct from that which gave rise to apo(a) in Old World

monkeys and humans; the latter is supported by the work by Lawn et al. (76).

The human gene encoding apo(a) is located on chromosome 6, within the

cytogenetic band 6q26 (77). As stated above, alleles of LPA contain variable numbers of

identical sequences encoding the apo(a) KIV2 domain. This gives rise to apo(a) of

different isoform sizes, which forms the basis of Lp(a) size heterogeneity in the

population, and ranging from a minimum of 3 KIV2 repeats to greater than 40 (64, 65).

Null alleles have also been identified in which one allele with an excessively large

amount number of KIV2 repeats does not produce a secreted protein (78) or the allele

Page 28: Characterization of the I4399M Variant of Apolipoprotein(a

13

possesses a mutation that alters a splicing site, preventing normal mRNA splicing and

resulting in a truncated, defective polypeptide that becomes degraded (79).

1.6 Plasma Lipoprotein(a) Levels - Relationship With Apo(a) Isoform Size

Plasma Lp(a) levels in the human population vary 1000-fold, ranging from <0.1 to

>100 mg/dL (80). Work by Rader et al. showed through radiolabeling of Lp(a) that

isoform size does not influence the rate of Lp(a) catabolism (80). Therefore, the inverse

association of isoform size and Lp(a) concentrations is due to differences in Lp(a)

production rate. Unlike LDL, plasma Lp(a) levels are primarily genetically determined

(81). In Caucasians, the LPA gene accounts for ~90% of the observed variation in plasma

Lp(a) levels (82), while only ~78% of the variation in the levels in African Americans can

be explained by the gene itself (83). A general inverse correlation exists in the human

population between apo(a) isoform size and Lp(a) concentrations, with smaller isoforms

correlated with higher Lp(a) protein levels (84, 85). This has been suggested to arise due

to less efficient secretion of larger apo(a) isoforms from hepatocytes, a result of longer

retention times in the endoplasmic reticulum and increased intracellular degradation (86).

Alternatively, mRNA transcript stability, translational efficiency and efficiency of Lp(a)

assembly may also contribute to the inverse correlation, although each individual

contribution to this process is still undetermined (59). Studies have shown that in

Caucasians, elevated Lp(a) levels were found to be associated with small isoforms in over

80% of subjects; while African Americans, elevated Lp(a) levels are distributed over a

broad range of isoform sizes (87, 88). As well, African Americans were shown to have

higher Lp(a) levels than Caucasians who posses the same apo(a) isoform size (88).

Smaller apo(a) isoforms (<22 KIV domains) have been linked by association to increased

Page 29: Characterization of the I4399M Variant of Apolipoprotein(a

14

instances of atherosclerosis and CVD due to elevated Lp(a) levels in these individuals

(87, 89, 90).

1.7 Transcriptional Regulation of LPA

In addition to size of the LPA gene, cis-acting elements may influence

transcription of apo(a), thereby modulating Lp(a) levels in plasma (91). For example,

work done by Chennamsetty et al. has shown that farnesoid X receptor (FXR), which is a

bile acid-activated receptor that is expressed in the liver, strongly controls LPA by

binding to a negative control element located at the -826-bp region of the gene (92). It

was also shown to interfere with hepatocyte nuclear factor 4α (HNF4α) mediated

activation of LPA transcription.

Owing to the strong regulation of Lp(a) plasma concentrations by the LPA gene

itself, levels of this lipoprotein are relatively resistant to conventional lipid-lowering

approaches, such as diet, exercise, and pharmaceutical interventions, including statin

therapy (93). Niacin, or vitamin B3, was able to reduce Lp(a) levels up to 35% after the

26th week of treatment in a clinical trial (94), but recent studies have shown that niacin

therapy is also associated with moderately increased risk for developing diabetes (95) and

a clinical study showed it trended towards increased development of new-onset diabetes

(96). Recent therapeutic strategies have focused on inhibitors of proprotein convertase

subtilisin/kexin type 9 (PCSK9), which has been shown to decrease plasma Lp(a) by up

to 30% (97) potentially through the up-regulation of the LDL receptor (LDLR) (98).

Page 30: Characterization of the I4399M Variant of Apolipoprotein(a

15

1.8 Identification of Two SNPs that Contribute Significantly to Plasma Lp(a) Levels

and Risk

Three consecutive case-control studies were conducted of men and women who

had undergone coronary angiography and were screened for 12,077 single nucleotide

polymorphisms (SNPs) in over 7,000 genes for association with severe CAD (99). After

screening all three study groups, an SNP was discovered in the LPA gene encoding

apolipoprotein(a) that was associated with severe CAD. This SNP, denoted rs3798220,

causes a missense mutation in the inactive protease domain of apo(a), resulting in an Ile

to Met substitution at the 4399 residue of apo(a). Carriers of this SNP accounted for 5.2%

of the sample population and had increased adjusted odds ratio for serve CAD compared

to non-carriers and carriers of the I4399M allele had significantly higher levels of Lp(a)

compared to the non-carriers. Further analysis of this mutation revealed that carriers of

the SNP had smaller apolipoprotein(a) isoform sizes and significantly higher levels of

OxPLs/apoB levels compared to non-carriers (100). However, the mechanism or cause of

its advanced pathogenic properties is still unknown. Another SNP in LPA (rs10455872)

has been identified in intron 25, resulting in an AG substitution, which was also found

to be associated with elevated Lp(a) levels and increased CAD risk (101). Recent meta-

analysis of other SNPs outside the LPA locus showed that no candidate genes have an

effect on Lp(a) levels (102). Interestingly the effects of the rs3798220 SNP were shown

to be present in carriers of European background, it has been shown that carriers of East

and Southeast Asian background had no association with smaller isoforms or elevated

Lp(a) concentrations and the variant was not found in African participates (103). While

the association of these two SNPs with high Lp(a) levels has been well documented, it has

Page 31: Characterization of the I4399M Variant of Apolipoprotein(a

16

been postulated that the increased risk of CAD in carriers of these SNPs is abolished

when adjusting for Lp(a) levels (104).

1.9 Pathogenicity of Lipoprotein(a)

Owing to key genetic studies performed within the last decade, Lp(a) is now

conclusively classified as an independent and causal risk factor for cardiovascular disease

(87, 105-108). High plasma Lp(a) concentrations have been determined to be a risk factor

for numerous atherosclerotic diseases, which includes coronary artery disease (CAD) (93,

109), peripheral vascular disease (110, 111), as well as thrombotic disorders that are not

secondary to atherosclerosis. In this regard, clinical studies have shown that high Lp(a)

levels (>30mg/dL) can be directly correlated to instances of ischemic stroke in children

(112, 113) and venous thromboembolism (114), although some studies have questioned

this direct relationship to instances of VTE (115, 116). Using immunohistochemistry,

Lp(a) has also been observed in the arterial intima at a site of an atherosclerotic plaque

(117), and it has been shown to preferentially accumulate within the lesion compared to

LDL alone (118). The extent of Lp(a) accumulation within the plaque was directly related

to the plasma levels of this lipoprotein(a). Due to the structural similarities between Lp(a)

and both LDL and plasminogen, it has been hypothesized that Lp(a) can contribute to

both proatherosclerotic and prothrombotic processes, both of which contribute to

cardiovascular events (Fig 1.4) (119). Mechanisms that are potentially proatherogenic

include increases in macrophage foam cell formation, smooth muscle cell (SMC)

proliferation and migration, monocyte chemo-attractant activity and endothelial cell

adhesion molecule expression (93, 120). Likewise, mechanisms that could be potentially

prothrombotic include increased platelet responsiveness (121), inhibition of fibrin clot

Page 32: Characterization of the I4399M Variant of Apolipoprotein(a

17

Figure 1.4. Potential pathogenic mechanisms of Lipoprotein(a). Various in vitro and

animal studies have been conducted with demonstrates Lp(a) can contribute to both

proatherogenic (left side) and prothrombotic mechanisms (right side) by which Lp(a)

exerts its pathogenic effects. EC, endothelial cells; SMC, smooth muscle cells; PAI-1,

plasminogen activator inhibitor-1; TFPI, tissue factor pathway inhibitor [Adapted from

Koschinsky ML (59)]

Page 33: Characterization of the I4399M Variant of Apolipoprotein(a

18

lysis (122, 123) and tPA-mediated plasminogen activation (124), which have been

observed in vitro and in vivo.

Although Lp(a) has been identified as an independent risk factor for CAD, the

exact mechanism of action of Lp(a) is still not fully understood. It is postulated that in an

atherosclerotic lesion, Lp(a) can interact with the plaque elements through binding of the

apo(a) moiety to matrix components such as fibrin, fibronectin, and laminin (125). The

strong LBS of KIV10 of apo(a) has been shown to mediate increased endothelial cell

contraction and permeability by a Rho/Rho kinase-dependent pathway (126). In addition,

the inflammatory environment of arterial lesions containing oxidized lipid moieties and

cellular debris may potentially interact with the Lp(a) molecule, enhancing its

pathogenetic effects through its effect on fibrinogen and other blood proteins (109, 127).

It is also hypothesized that the apo(a) component of Lp(a) upregulates plasminogen

activator inhibitor-1 (PAI-1) expression in ECs, preventing the cleavage of the

plasminogen pro-enzyme (128). As well, inhibition of plasmin formation prevents the

plasmin-mediated activation of transforming growth factor-β (TGF-β), resulting in the

inhibition of SMC migration and proliferation (129, 130). Since plasminogen activation

and fibrinolysis are critical mechanisms in many cellular processes, knowledge of the

exact mechanisms by which Lp(a)/apo(a) exert its downstream effects are crucial in

disease prevention.

1.10 Inhibition of Plasminogen Activation and Fibrinolysis by Lp(a)/apo(a)

Inhibition of fibrinolysis has also been citied as a risk factor for atherosclerosis,

leading to increased fibrin deposition in vasculature and the persistence of mural thrombi

within atherosclerotic plaques (131, 132). Lp(a)/apo(a) has been shown to inhibit

Page 34: Characterization of the I4399M Variant of Apolipoprotein(a

19

fibrinolysis, contributing to its prothrombotic potential (133). Due to the structural

similarities between apo(a) and plasminogen, Lp(a) is therefore able to inhibit fibrinolysis

through binding to the free C-terminal lysine residues on fibrin, creating a quaternary

complex with plasminogen on the fibrin surface, preventing tPA-mediated plasminogen

activation (71). Hancock et al. investigated the effect of the domains of apo(a) on

plasminogen activation on the fibrin surface and determined that the strong LBS in KIV10,

the KV domain and the amino terminus of apo(a) were required for maximum inhibition

(71). Similarly, it was also determined that apo(a) isoform size does not affect degree of

inhibition of activation (71). Apo(a) is also able to inhibit pericellular plasminogen

activation on endothelial cells and THP-1 monocytes in vitro through binding of cell

surface plasminogen receptors by the strong LBS in KIV10 and KV domain, preventing

binding by plasminogen (134). Finally, it has been shown that apo(a) is able to interfere

with the positive feedback mechanism of Glu1-Pg to Lys78-Pg conversion, with the KIV5-9

domains being critical in this effect and effect was shown to be independent of isoform

size (135).

1.11 Lipoprotein(a) and Oxidation

The mechanisms underlying the atherogenic action of Lp(a)/apo(a) has been

attributed, in part, to the presence of oxidative modification on both the lipid and protein

component (apo(a) and apoB-100) of Lp(a) (136-138). This modification on Lp(a) causes

diverse biochemical (139) and functional changes that promote atherogenesis (140). For

example, oxidation of Lp(a) has been reported to: increase PAI-1 production in vascular

endothelial cells (141), impair endothelium-dependent vasodilation (142), allow Lp(a)

uptake by macrophages though scavenger receptors (143), stimulate SMC proliferation

Page 35: Characterization of the I4399M Variant of Apolipoprotein(a

20

(144), and induce adhesion of monocytes onto endothelial cells (145). Lp(a) functions in

a similar way to LDL to enter the intima at the site of an atherosclerotic plaque through

leaky junctions in the endothelial layer (146), however Lp(a) has a longer half-life than

LDL and therefore results in larger accumulation of Lp(a) within an atherosclerotic

plaque compared to LDL (137). This prolonged accumulation of Lp(a) within the sub-

intima relative to LDL is hypothesized to be due to Lp(a)’s affinity for the extracellular

matrix (ECM) and fibrin (72, 147). Furthermore, there are lower antioxidants associated

with Lp(a) than LDL, which together with the increased accumulation, promotes

preferential Lp(a) oxidative modification compared with LDL (148).

More recently, Lp(a) has been shown to be a preferential carrier of oxidized

phospholipids (OxPLs) (149). OxPLs can covalent modify proteins, such as apo(a),

through a process by which the phospholipid oxidizes to form aldehydes, which can then

react with ε-amino groups on lysine residues to form Schiff base adducts (150) or through

Michael addition (151). It has been shown that apo(a) is modified by the addition of

oxidized phosphatidylcholine (OxPtdC), however the exact location of the modification

and the functional consequences have not been fully investigated (152). It has been shown

that apo(a) elicits a dose-dependent increase in interleukin-8 (IL-8) mRNA and protein

production in macrophages through signaling pathways with CD36 and TLR2 and the

OxPL on ap(a); mutating apo(a) in a way that abolishes the OxPL blunts this

inflammatory response (153).

1.12 LPA rs3798220 SNP and Thrombotic Risk

As stated above, carriers of the rs3798220 SNP are associated with elevated Lp(a)

levels and increased adjusted odds ratio for severe CAD (99). A recent randomized trial

Page 36: Characterization of the I4399M Variant of Apolipoprotein(a

21

of low-dose aspirin therapy was conducted on a group of 25,131 healthy Caucasian

women to deduce whether aspirin reduced cardiovascular risk in minor allele carriers

(154). They determined that carriers of the allele possessed elevated Lp(a) levels and

increased cardiovascular risk, but the carriers benefitted over two-fold compared to non-

carriers following the trial. More recently, work done by Rowland et al. showed that the

clot properties of carriers of the SNP depended on ethnicity, where Caucasians had

decreased permeability and longer lysis time, and non-Caucasians had increased

permeability and shorter lysis times (155).

1.13 Objectives

Clinical data has shown that carriers of the rs3798220 SNP have altered

permeability and lysis times compared to non-carriers, and more effectively responds to

aspirin therapy. However, the mechanism and biochemical rationale behind this

observation has not been fully investigated. The purpose of this thesis will be to

characterize the I4399M variant apo(a) using a variety of biochemical approaches. The

Met variant will be introduced into a physiologically relevant apo(a) variant containing 17

KIV domains (17K) and compared to wild-type 17K without the mutation. Specific

objectives of this research project were developed as follows:

1) Perform pulse-chase assays in HepG2 cells to determine whether an altered

secretion rate of the Met variant contributes to the elevated Lp(a) plasma

concentrations

2) Analyze the effect of the apo(a) Met variant on fibrin clot properties compared

to wild-type apo(a), such as fibrinolysis, coagulation, and fibrin clot

architecture

Page 37: Characterization of the I4399M Variant of Apolipoprotein(a

22

3) Determine if the Met variant causes a structural change in apo(a) or possible

modification of apo(a) through sulfoxidation

1.14 Hypothesis

The 17K I4399M variant is structurally and therefore functionally different from

wild-type 17K, resulting in the increased CAD risk with enhanced prothrombotic

potential associated with the Met variant.

Page 38: Characterization of the I4399M Variant of Apolipoprotein(a

23

Chapter 2

Methods

2.1 Cell Culture

2.1.1 Construction and expression of recombinant apo(a) variants

Plasmids encoding a 17-kringle-containing form of recombinant apo(a) (17K r-

apo(a)), as well as KIV10-P, a truncated version of 17K containing the plasminogen-like

kringle IV10, kringle V, and the inactive protease domain; were constructed and expressed

in human embryonic kidney (HEK 293) cells. 17K I4399M and KIV10-P I4399M

variants, corresponding to the rs3798220 single nucleotide polymorphism, were generated

by site directed mutagenesis using the QuikChange II-XL Site-Directed Mutagenesis Kit

(Agilent Technologies, Santa Clara, CA, USA) with the mutagenic primer pair; sense 5’-

CAT GTT CAG GAA ATg GAA GTG TCT AGG CTG-3’ and antisense 5’-CAG CCT

AGA CAC TTC cAT TTC CTG AAC ATG-3’. The constructed vectors were co-

transfected with MegaTran 1.0 transfection reagent (OriGene, Rockville, MD, USA) and

pRSV-Neo vector (ATCC, Manassas, VA, USA), and dilution cloned to create stable cell

lines.

2.1.2 Propagation of cell lines

Cell culture materials, including culture dishes, 96 well plates, and tubes (Sarstedt,

Montreal, Quebec, Canada) were used in a sterile class-II type Biosafety Cabinet

(NuAire, Plymouth, MN, USA). Human embryonic kidney (HEK 293) cells (ATCC)

were grown with minimum essential media (MEM, Thermo Fischer Scientific, Waltham,

Page 39: Characterization of the I4399M Variant of Apolipoprotein(a

24

MA, USA), completed with 10% fetal bovine serum (FBS, Thermo Fischer) and 1% (v/v)

antibiotic solution (100X Anti-Anti, Thermo Fischer). Human hepatoma (HepG2) cells

(ATCC) were grown in MEM (Thermo Fischer), completed with 10% ATCC FBS

(ATCC), and 1% (v/v) antibiotic solution (Thermo Fischer). All cells were grown at

370C in 5% CO2 in a CO2 incubator with a HEPA filter (Forma Scientific, Marietta, Ohio,

USA), with new media being supplemented every 2 days.

2.1.3 Purification of recombinant apo(a) variants

All recombinant apo(a) variants were purified from conditioned medium of stably

expressing HEK293 cell lines using lysine-Sepharose affinity chromatography (GE

Healthcare, Mississauga, ON, Canada), as previously described (Hancock 2003). Briefly,

conditioned media (Opti-MEM, Thermo Fischer) was collected every three days from the

stably transfected cells in 2L Corning roller bottles (Sigma-Aldrich, Oakville, ON,

Canada). The media was loaded in a 50mL column containing lysine-Sepharose CL-4B

(GE Healthcare), pre-equilibrated with phosphate-buffered saline (PBS, 137mM NaCl,

2.68mM KCl, 10.14mM Na2HPO4, 1.38mM KH2PO4, pH 7.4). The loaded column was

washed with 10 column volumes of PBS and eluted with 200mM ε-aminocaproic acid (ε-

ACA, Sigma-Aldrich) in PBS buffer. The fractions were collected and analyzed

spectrophotometrically at 280nm and protein-containing fractions were pooled. The

pooled fractions were dialyzed three times in Hepes-buffered saline (HBS, 150mM NaCl,

20mM HEPES, pH 7.4) at 40C and dehydrated using 20,000 MW polyethylene glycol

(PEG, Sigma-Aldrich). The concentration was determined spectrophotometrically for all

apo(a) variants (corrected for Rayleigh scattering), using the following molecular

weights and molar extinction coefficients: 17K and 17K I4399M (Mr ~ 278,219; ε280nm =

Page 40: Characterization of the I4399M Variant of Apolipoprotein(a

25

2.07); KIV10-P and KIV10-P I4399M (Mr ~ 52,040; ε280nm = 2.16). All proteins were

assessed for purity by SDS-PAGE and silver staining, under reduced (containing 10mM

dithiothretiol) and non-reduced conditions. Aliquots of the purified proteins were stored

at -70oC.

2.2 Characterization of lipoprotein(a) in human plasma samples

Blood samples were collected in citrated tubes and DNA was extracted using the

GeneJET Whole Blood Genomic DNA Purification Mini Kit (Thermo Scientific). An

additional blood sample was collected, with minimal stasis, into an 8.5 ml Vacutainer®

Glass citrated tube (Product Number 364606; BD, Franklin Lakes, NJ). This tube was

centrifuged at 3500 g for 15 min and plasma immediately frozen at −80 °C in aliquots of

0.5 ml. rs3798220 genotype was determined by sequencing a 427 base pair amplicon of

the LPA gene performed by the London Regional Genomics Centre (Robarts Research

Institute, London, Ontario). Sequences were analyzed using DNAman 3.0 software

(Lynnon Biosoft, Quebec, Canada). Primers for PCR were 5-GAA GGG GCT GGA CCA

TAT TT-3 and 5-CGG GTA ACA GGG TAT CTC TTT-3 (156). The PCR conditions

used were: at 98 °C followed by 36 cycles of 98 °C for 30 sec; 53 °C for 30 sec; 72 °C

for 40 sec, and a final extension for 10 min at 72 °C. The final reaction volume was 25 μl

and contained 10 ng genomic DNA, and were carried out using Q5® High-Fidelity DNA

Polymerase (New England Biolabs, Ipswitch, MA, USA) as per manufacturer’s

instructions. Plasma samples were reduced and assessed for lp(a) isoform size using a

SDS-Agarose gel electrophoresis, and were compared to a ladder of recombinant apo(a)

isoform standards. Lp(a) plasma levels were determined using the Macra® Lp(a) Enzyme

Page 41: Characterization of the I4399M Variant of Apolipoprotein(a

26

Linked Immunosorbant Assay (Mercodia, Uppsala, Sweden) as per manufacturer’s

protocol.

2.3 Lipoprotein-Deficient Plasma

Lipoprotein-deficient plasma (LPDP) was prepared from human plasma collected

from healthy individuals that did not possess the rs3798330 genotype. Full blood was

spun at 15,000rpm at 40C for 15 minutes, the plasma was removed from the cells and the

density was adjusted from 1.006 g/mL to 1.21 g/mL with NaBr according to the following

equation:

NaBr (g) = [mL Plasma x (ρFinal – ρInitial)]/(1 - 0.245 x ρFinal)

where ρFinal = 1.21 g/mL, ρInitial = 1.006 g/mL and 0.245 mL/g is the partial specific

volume of NaBr. Sealable centrifuge tubes (Beckman Coulter, Mississauga, ON, Canada)

were filled with the plasma, sealed shut and spun at 45,000rpm for 20h at 40C in a

Beckman Ultracentrifuge with the type 70.1 Ti Rotor. The plasma separated into two

distinct layers and the top layer; containing LDL, VLDL, Chylomicrons, and Lp(a), was

carefully removed with a syringe. The lipoprotein-free layer was dialyzed against HBS

buffer at 40C and aliquots were stored at -200C prior to use.

2.4 Plasminogen activation on fibrin or cell surfaces

THP-1 monocytes were seeded at a density of 200,000 cells per well. Cells were

washed three times with HBS containing 0.4% (w/v) BSA prior to adding reaction

mixture. Reaction mixtures contained: 600 nM plasma derived plasminogen, 15 nM tPA

Page 42: Characterization of the I4399M Variant of Apolipoprotein(a

27

(Alteplase), H-D-Val-Leu-Lys-7-amido-4-methylcoumarin (Bachem), and 400nM 17K or

17KI4399M protein. Plasmin formation was monitored over 1 hours at 37°C at an

excitation wavelength of 370 nm and an emission wavelength of 470 nm and emission

cutoff filter of 455 nm using a plate reading fluorescence spectrometer (SpectraMax M5e,

Molecular Devices). The rate of plasminogen activation was taken as the initial slope of

the plot of RFU against min2 from 10 to 40 mins as performed previously (157). Rates

obtained for each individual experiment performed in triplicate were normalized to the

rate of plasminogen activation in the absence of apo(a), or to 17K treatment.

Similarly, plasminogen activation on the surface of a fibrin clot was conducted by

methods previously published (71). Recombinant plasminogen containing an serine to

cysteine mutation in the active site (S741C) was expressed and purified from baby

hamster kidney cells (BHK-21) and was labeled with 5-iodoacetamidoflurocein, as

described previously (34). Cleavage of Flu-Plasminogen was analyzed in a 96-well

Microfluor2 black plates (Costar) pre-coated with HBST (150mM NaCl, 20mM HEPES,

pH 7.4, 1% (v/v) Tween 80). Flu-Plasminogen was titrated in triplicate at 6 different

concentrations (0, 0.3, 0.6, 0.9, 1.2, and 1.5 μM) and added to HBST (0.02% (v/v) Tween

80) containing human fibrinogen (1mg/mL final) and wild-type or I4399M variant. This

mix was added to wells containing small aliquots of lla, CaCl2, and tPa, and final

concentrations of 1U/mL, 10mM, and 5nM, respectively. The quenching of fluorescence

was measured under the following conditions: 370C, excitation wavelength, 495nm;

emission wavelength, 535nm; cutoff wavelength, 530nm; sensitivity normal; PMT

setting, low; run-time, 1h; 36 second intervals. Raw fluorescence measurements were

corrected for the buffer blank and corrected for internal filter effects. Linear regression

Page 43: Characterization of the I4399M Variant of Apolipoprotein(a

28

was used to determine the initial rates of fluorescence decrease and the rate of plasmin

formation was given by the equation:

d[Pn]/dt = (dI/dt)*(1/(r*I0))*([P]0/[A]0)

where d[Pn]/dt represents the rate of plasmin formation per mole of tPa (s-1), dI/dt is the

initial rate of fluorescence decrease, r is the relative maximum change in fluorescence

intensity (0.5 for fibrin surface (34)), I0 is the initial fluorensence intensity, [P]0 is the

initial Flu-plasminogen concentration, and [A]0 is the initial tPA concentration. The rates

of plasminogen activation are the average of triplicate measurements.

2.5 Assessment of plasma clot permeability

Fibrin clot permeability was determined using a pressure-driven system, briefly,

20mM CaCl2 and 1 U/mL human thrombin (Haemotological Technologies Inc.) were

added to 120 μl citrated plasma and transferred to clear glass tubes to hold the clot. After

2 hours of incubation in a wet chamber, tubes containing the clots were connected via

plastic tubing to a reservoir of a buffer (0.01 M Tris, 0.1 M NaCl, pH 7.5) and the

volumes flowing through the fibrin gels were measured within 60 minutes. A permeation

coefficient, referred to as a Darcy constant, (Ks), was calculated from the equation:

Ks = (Q x L x η)/(t x A x Δp)

where Q was the flow rate in time t; L, the length of a fibrin gel; η, the viscosity of liquid

(in poise); A, the cross-sectional area (in cm2), and Δp, the pressure differential (in

Page 44: Characterization of the I4399M Variant of Apolipoprotein(a

29

dyne/cm). The resulting Darcy constants were expressed in units of cm2, representing the

intrinsic permeability of the clot.

2.6 In Vitro fibrinolysis and coagulation assay

Clot lysis assays were performed in similar manner to that previously published

with some modification (158). Pooled lipoprotein-deficient plasma (LPDP) was diluted

1:1 with TBST solution (0.02M Tris, 0.15M NaCl, 0.01% Tween 20, pH 7.4) and spiked

with 3 different amount of recombinant wild-type and I4399M apo(a). The reaction

mixture was added to wells containing small, separated aliquots of CaCl2, human

thrombin (Sigma-Aldrich), and tPa (Alteplase) at a final concentration of 10mM, 1U/mL

and 3nM, respectively. The kinetics of lysis was analyzed spectrophotometrically at

405nm at 370C for 3 h to measure the change in turbidity in the well. The time required

for 50% decrease in clot turbidity (t50%) was reported. Fibrinolysis assay was conducted

again for human plasma samples that were homozygous carriers and non-carriers of the

SNP by clotting with 20mM CaCl2 and 1U/mL of lla; the t1/2 values were also reported.

Plasma coagulation assays were performed in similar manner to the clot lysis assay, but

was clotted with CaCl2 and human thrombin (Sigma) at a final concentration of 20mM

and 1U/mL, respectively, in the absence of tPa. The time required to reach ½ max

absorbance in clot turbidity (t½) and the extent of clot formation (ΔAbsorbance) were

reported.

2.7 Analyzing the interaction of I4399M apo(a) with soluble fibrinogen

3mg/mL Fibrinogen (Calbiochem, Mississauga, ON, Canada) was mixed together

with the wild type and I4399M variant to a final concentration of 3μM, vortexed, and

Page 45: Characterization of the I4399M Variant of Apolipoprotein(a

30

incubated for 10 mins at room temperature. The samples were then prepared with the

addition of 4X SDS reducing sample buffer and boiled at 90C for 10 mins. Samples

were analyzed on a 4-15% SDS-PAGE gel and stained using Imperial Stain (Thermo

Scientific) for visualization of proteins.

2.8 Imaging of Fibrin Clots with Confocal and Scanning Electron Microscopy

For scanning electron microscopy (SEM), clots were generated by clotting

3mg/mL fibrinogen (Calbiochem, Mississauga, ON, Canada) with 1U/mL thrombin

(Haemotologic Technologies) and 20mM CaCl2, in the absence and presence of wild-type

and mutant I4399M apo(a), at a final concentration of 1μM. A mixture of Fg and apo(a)

was added to a microdrop of 20mM CaCl2 and 1U/mL IIa on a SEM peg with conductive

tape and the clot matured for 3h in a moist chamber box. The clots were fixed in 2.5%

Glutaraldehyde (Thermo Fischer) for 30min, followed by four 30 minute washes in dH2O.

For human plasma samples, plasma was clotted with 20mM CaCl2 and 1U/mL IIa in a

glass tube and matured for 2 hours. The clots were permeated for 1h, removed from glass

tube, fixed for 30min, and washed four times in dH20 for 30min. The clots were imaged

using field-emission environmental scanning electron microscope (ESEM, FEI

QUANTA-200 FEG type, Holland FEI Company). Images were taken at a pressure of

70kPa, operating at an accelerating voltage of 10kV and approximately a 10mm working

distance. The images were observed at 15,000X and 30,000X magnification. ImageJ was

used to count fiber widths on the resulting images

In the case of confocal microscopy images, clots were formed with plasma from

patients without the rs3798220 SNP and homozygous carriers of the genotype with

Page 46: Characterization of the I4399M Variant of Apolipoprotein(a

31

20mM CaCl2 and 1U/mL of IIa. The mixtures were supplemented with 0.3mg/mL Alexa

Fluor® 568 labelled fibrinogen (Calbiochem, Mississauga, Canada), fluorescent labelling

was performed using Alexa Fluor® 568 Protein Labeling Kit (Invitrogen). Clots were

made in glass tubes and matured for 2h. Following maturation, the clots were permeated

for 1h and placed on microscope slides. The resulting fluorescent clots were imaged on an

Olympus FV1000 confocal microscope at 60X magnification. ImageJ was used with a

macro to count the number of fibers per 100μm.

2.9 Pulse Chase Secretion Assay

HepG2 cells were transiently transfected with an apo(a) expressing construct

(pRK5ha17) encoding 17 KIV repeats without the mutation and containing the I4399M

mutation. The cells were grown to 90% confluency and split into a 6 well flask (Sarstedt)

the day before the experiment. On the day of the experiment, the cells were starved for 1h

with methionine- and cysteine-free DMEM without serum, pulse-labeled for 1h in the

same media containing 200 μCi/mL [S35]methionine, and chased in complete media

containing 10nM methionine. The media was collected off the cells at 0, 30, 60, 120, 240,

and 480 min. Following removal of the media, the cells were washed twice with PBS and

the cellular extract was collected after treating the cells with cold lysis buffer (100mM

Tris, pH 8.0, 100mM NaCl, 10mM EDTA, 1% Triton X-100, 0.1% SDS, 100mM ε-

aminoaproic acid) containing protease inhibitors. Cellular debris was removed from

media and cellular extract through centrifugation, and both samples were pre-cleared by

incubation with gelatin-agarose (Sigma) for 2h at 40C. After centrifugation, the

supernatants were incubated with an anti-apo(a) primary antibody overnight at 40C. The

following day, r-protein-G (Invitrogen) beads were added to the samples and incubated

Page 47: Characterization of the I4399M Variant of Apolipoprotein(a

32

for another 2h at 40C. Following centrifugation, the resulting pellets were washed twice in

PBS (137mM NaCl, 2.7mM KCl, 10.1mM Na2HPO4, 1.8mM KH2PO4, pH 7.4) and once

in TE buffer (10mM Tris, 1mM EDTA, pH 7.0). The pellets were then boiled for 10mins

in SDS sample buffer (4% SDS, 20% glycerol, 0.001% bromophenol blue, 125mM Tris,

100mM dithiothreitol, pH 6.8). Immunoprecipitated samples were run on 7% SDS-PAGE

gels, dried, and exposed using autoradiography (Fujifilm). After 4 days, the image plate

was then imaged using a Molecular Imager® FX (Bio-RAD) and the appearance and

intensity of the bands were analyzed.

2.10 Molecular dynamics simulations and protease domain modeling

The Molecular Operating Environment (MOE) software was used for model

preparation, energy minimizations, and assessment of the generated trajectories. The

molecular dynamics (MD) simulations were performed using the NAMD program. All

chemical models were derived from the crystal structure obtained by Gabel et al (73).

More specifically, water molecules present in the crystallographic structure were deleted.

Hydrogens and protons were then added to the model using the default method in MOE

with water molecules then added using the TIP3P water model. The latter added water

molecules throughout the protein, where appropriate, and extending outwards from the

edge of the protein to form a 6 Å layer of water molecules. The resulting solvated model

was then minimized with the AMBER:12EHT force field(6) until the root-mean-square

gradient of the total energy was below 0.05 kcal mol–1 Å–1. The same procedure was used

to generate both initial solvated structures; the wild-type apo(a) and the I4399M variant.

For all MD simulations the AMBER12:EHT force field was used as was the default

velocity Verlet integration method to obtain the trajectories. The solvated systems were

Page 48: Characterization of the I4399M Variant of Apolipoprotein(a

33

first equilibrated for 0.15 ns at 150 K, while the subsequent production MDs ran for 4.90

ns with a time step of 2 fs.

2.11 Detection of methionine sulfoxide residue on KIV10-PI4399M

To characterize recombinant protein samples, 15g of either KIV10-P or KIV10-

PI4399M protein was reduced and labeled with 100mM iodoacetamide (Sigma) for 60

min, shaking, at room temperature in the dark. Samples were then subjected to a buffer

exchange with ice cold trichloroacetic acid (15% v/v), 20 min incubation at 4C, followed

by 3X acetone washes of the pellet and re-suspended 4X reducing sample buffer. Samples

were then purified using SDS-PAGE followed by visualization of proteins by Imperial

stain, as per manufacturer’s protocol, excision of the protein band, and subjected to a 16

hour in-gel tryptic digest with a ratio of ~80:1 apo(a):Trypsin Gold (Promega) (159).

Tryptic peptides were then extracted, spotted in a 1:1 ratio with alpha-cyano-4-

hydroxycinnamic acid and subjected to MALDI-TOF mass spectrometry using a Waters

SYNAPT G2-Si instrument. The resulting spectra were analyzed using MassLynx

Software (Waters).

2.12 Statistical methods

Statistical analyses were performed with the use of SPSS software, version 22.0

(SPSS Inc., Chicago, Illinois). Comparisons between samples were performed by one-

way ANOVA using a Tukey post-hoc analysis. Statistical significance was assumed at p

< 0.05.

Page 49: Characterization of the I4399M Variant of Apolipoprotein(a

34

Chapter 3

Results

3.1 Structural Differences in Apo(a) Variants Predicted by Molecular Dynamic

Simulations

The rs3798220 SNP in the LPA gene, first identified by Luke et al. (99) causes a

missense mutation, resulting in an isoleucine to methionine substitution at the 4399

position of the protease domain of apo(a). However, resulting conformational changes of

the protease domain caused this mutated residue have yet to be studied. Therefore, we

began our investigation by using Molecular Operating Equipment (MOE) software to

conduct molecular dynamic simulations, using apo(a) variants, with and without the

mutation, derived from the crystal structure of the protease-like domain of apo(a)

obtained from Gabel et al (73).

Average occupancy images of the wild-type and I399M variant were generated to

determine if any structural differences exist between the Met and Ile variants (Fig 3.1).

Comparison of the root-mean-square deviation (RMSD) of the coordination sphere of the

two variants showed no gross conformational changes of the protease domain in either

protein. However, it was observed that in the wild-type apo(a) variant, the 4399Ile residue

is buried within the inactive protease domain, shielded from the solvent by an adjacent

Arg residue at position 4420. With the introduction of the 4399Met residue, the adjacent

4420Arg residue no longer shields the residue and is in fact pushed into the solvent. This

allows the mutant Met residue to be exposed to the environment, making it accessible to

possible modifications.

Page 50: Characterization of the I4399M Variant of Apolipoprotein(a

35

Figure 3.1. Molecular dynamic simulations of wild-type and mutant apo(a) variants. A. To determine any structural differences between the wild-type and mutant variants,

Molecular Operating Environment (MOE) software was used to produce average

occupancy images using the NAMD program for the two apo(a) variants derived from the

crystal structure reported by Gabel et al (73). B. Root mean squared deviation (RMSD)

analysis was conducted on the gamma carbon of Ile and gamma sulfur of Met versus the

proximal guanidino nitrogen of the adjacent Arg. The average distance between the two

atoms over the course of the simulation was reported.

3

5

7

9

11

13

0 1000 2000 3000 4000 5000

Dis

tan

ce (

Å )

Time (sec)

IleMet

Ile

Met

Avg. Distance

=3.39±0.03Å

Avg. Distance

=4.37±0.12Å

17K 17K I4399M

A

B

Page 51: Characterization of the I4399M Variant of Apolipoprotein(a

36

Looking further into this result, RMSD analysis was conducted on the average

distance (Å) between the gamma carbon (4399Ile residue) or sulfur (4399Met

residue) and the proximal guanidino nitrogen (4420Arg residue) (Fig 3.1). Over the

course of the simulation, 4399Ile and 4420Arg have no significant interaction, seen by the

fluctuation in the distance between the two atoms (Average distance; 4.37+ 0.12Å), while

the 4399Met and 4420Arg stay in relatively close proximity (Average distance; 3.39 +

0.03Å), even with both residues exposed to the environment, and may coordinate to

remain in this low energy state.

Despite the observation that the 4399Met residue does not cause any gross

conformational changes in the protease-like domain, the residue becomes exposed to the

external environment and could be potentially modified. Furthermore, the Met residue

may coordinate with a proximal Arg residue, a structural conformation not seen in wild-

type apo(a). Therefore, the effect of the structural difference of I4399M variant on the

processes of thrombosis and fibrinolysis, compared to wild-type apo(a), were further

studied.

3.2 Secretion of Wild-Type and Mutated r-Apo(a) Variants from HepG2 cells

Previous studies on the I4399M apo(a) variant have demonstrated that carriers of

the corresponding rs3798220 SNP are associated with smaller apo(a) isoform sizes, as

well as high plasma Lp(a) concentrations (100). It has also been reported that apo(a)

variants with smaller isoform sizes secrete more efficiently than larger apo(a) isoform

sizes (160). This suggests that increased secretion of apo(a) containing the mutation from

hepatic cells could be contributing to the high Lp(a) levels in carriers of this SNP. The

purpose of this assay was to determine if this observation was the result of the association

Page 52: Characterization of the I4399M Variant of Apolipoprotein(a

37

of this mutation with smaller isoform sizes or if the mutated residue is affecting its

secretion.

To address this, we conducted an in vitro pulse-chase secretion assay in cultured

human hepatoma (HepG2) cells transfected with plasmids containing the cDNA of

recombinant apo(a) variants, containing 17 KIV domains, with and without the mutation

(Figure 3.2). This experiment was conducted by incubating the transfected HepG2 cells

with 200 μCi/mL [S35]methionine and chased in complete media, followed by the

collection of the media and lysates of the cells at 6 increasing time points. Samples were

immunoprecipitated, run on an SDS-PAGE gel and exposed to a screen for 3 days. Before

conducting the assay, it was first important to determine the time points to collect samples

to ensure that the decrease in apo(a) in the lysates and subsequent increase in apo(a) in the

media could be visualized. It was concluded that the time points 0, 30, 60, 120, 240, and

480 minutes effectively showed this trend in the secretion profiles. Following the

exposure to the screen, the intensity of bands for the 17K and 17K I4399M variants in the

media and lysates were normalized against the maximum band intensity, and plotted as a

function of time (Figure 3.3). As anticipated, the intensity of apo(a) in the lysates

decreased, while the intensity of apo(a) in the media increased as a function of time.

However, compared to the 17K wild-type apo(a), there was no significant difference in

the intensity of apo(a) bands at any time point in the media or lysates compared to the

17K I4399M variant. Both variants showed maximum apo(a) intensity in the media at 30

minutes, and at 240 minutes in the lysates following pulse.

Page 53: Characterization of the I4399M Variant of Apolipoprotein(a

38

Figure 3.2 Schematic representation of the recombinant apo(a) variants utilized in

the proposed studies. The 17K recombinant apo(a) construct shown on the top line

represents a physiologically relevant apo(a) isoform that contains 17 plasminogen-like

kringle IV domains, denoted 1-10, and 8 KIV2 repeats. KV represents the plasminogen-

like kringle V domain and P is the inactive protease domain. The white circle within the

second construct, labeled 17K I4399M, represents the substitution in the protease domain

of apo(a) introduced using site directed mutagenesis.

Page 54: Characterization of the I4399M Variant of Apolipoprotein(a

39

Figure 3.3 Pulse-chase analysis of the secretion of 17K versus 17K I4399M from

HepG2 cells. Transfected HepG2 cells were pulse labeled with [S35]methionine and

chased in completed media. Media and lysates were collected at 0, 30, 60, 120, 240, and

480 minutes, run on SDS-PAGE gel and exposed on a screen for 3 days before imaging.

A. Representative images of the apo(a) intensity in the media and lysates of the apo(a)

variants at the six different time points. B. The intensity of apo(a) for the media and

lysates were normalized to maximum intensity and plotted as a function of time. The

results represent the mean + SEM of four independent experiments.

A

B

Page 55: Characterization of the I4399M Variant of Apolipoprotein(a

40

Therefore, this result suggests that the presence of the mutation does not alter the

rate of secretion in apo(a) variants of the same length. Using this knowledge, these two

apo(a) variants of same length were used in subsequent assays to deduce the specific

affect of this mutation on the coagulation and fibrinolytic pathways.

3.3 Effect of I4399M Variant on Prothrombotic Mechanisms Compared to Wild-

Type Apo(a)

Previous studies have shown that Lp(a)/apo(a) has the capability to inhibit in vitro

fibrinolysis through a mechanism involving binding to lysine residues of fibrin,

preventing tPA-mediated plasminogen activation on the fibrin surface (134, 161) and

therefore inhibiting lysis of the fibrin network (140, 162). Similarly, it has also been

shown that apo(a) is able to inhibit pericellular plasminogen activation on multiple cell

types through binding to cell surface receptors and preventing the binding of plasminogen

(134). Therefore, we investigated the effect of the I4399M variant on these processes

compared to the wild-type apo(a) variant.

We began by performing a tPA-mediated plasminogen activation assay on a fibrin

surface with 0.9μM of catalytically-inactive, fluorescently-labeled plasminogen in the

presence of wild-type and mutant apo(a). The rate of fluorescence decrease measured by

spectrophotometry was determined and inserted into the derived equation to determine

the rate of plasmin formation per mole of tPA. These values were then normalized against

the rate with no apo(a) and presented as the relative % change at different concentrations

of apo(a) (Figure 3.4). At an apo(a) concentration of 0.5μM, the wild-type variant

inhibited the rate of plasmin formation by 25% compared to no apo(a) control, while the

I4399M variant inhibited the rate by 27%. When the concentration was increased to 1μM,

Page 56: Characterization of the I4399M Variant of Apolipoprotein(a

41

the wild-type and I4399M variant inhibited plasmin formation rate by 44% and 46%

respectively. As expected, increasing concentration of apo(a) results in greater inhibition

of plasmin formation on the fibrin surface (71). However, there was no statistical

difference between the percentage of inhibition caused by wild-type and mutant apo(a)

variants. To examine the effect of the mutant on pericellular plasminogen activation, an

assay was conducted where the purified apo(a) variants were incubated with THP-1

monocytes and a fluorescent substrate of plasmin. The rate of plasminogen activation was

taken as the initial slope of RFU change per min2 and normalized against the rate with no

apo(a) present (Figure 3.5). As expected, the wild-type apo(a) variant inhibited

plasminogen activation on the cell surface by ~30%. Surprisingly, the I4399M variant

showed no inhibition of pericellular activation compared to the normal and resulted

significantly less inhibition compared to the wild-type variant.

To expand on the preceding result, we investigated how the I4399M variant

affects fibrinolysis compared to wild-type. A fibrinolysis assay was conducted, where

lipoprotein-deficient plasma (LPDP) was incubated with increasing concentrations of the

wild-type and I4399M apo(a) variants and the change of absorbance was measured by

spectrometry at 370C for 3h. The absorbance profiles were generated and the ½ time

required for clot lysis was reported (t50%) (Figure 3.6). At an apo(a) concentration of

0.5μM, the t50% lysis of the wild-type apo(a) was determined to be 2182 seconds, ~3X

greater than control (740 seconds), while the I4399M variant had a t50% of 3020 seconds,

~4X greater than the control. At 1μM apo(a) concentration, the t50% times for the 17K and

I4399M variant were ~5.5X greater (3987 seconds) and ~7X greater (5023 seconds) than

control, respectively.

Page 57: Characterization of the I4399M Variant of Apolipoprotein(a

42

Figure 3.4 tPA-mediated plasminogen activation assay on a fibrin surface in

presence of r-apo(a) variants. Purified fibrinogen was clotted with 1U/mL lla, 10mM

CaCl2, and 5nM tPa in the presence and absence of wild-type and mutant apo(a). Samples

were spiked with a catalytically inactive variant of plasminogen labeled with 5-

iodoacetamide. The quenching of fluorescence was measured by spectroscopy at 370C;

excitation: 495nm, emission: 535nm, cutoff: 530nm. The rate of plasmin formation per

mole of tPa was calculated and was represented as a relative percentage change of the rate

without apo(a). *: p < 0.05 by two tailed ANOVA. The data represent the means ± SEM

of four independent experiments each conducted in triplicate.

-60.00

-50.00

-40.00

-30.00

-20.00

-10.00

0.00

0.5 1

Rel

ativ

e %

chan

ge

in p

lasm

in f

orm

atio

n

com

par

ed t

o c

on

tro

l

Concentration of apo(a) (μM)

17K 17K I4399M

*

*

Page 58: Characterization of the I4399M Variant of Apolipoprotein(a

43

Figure 3.5 Plasminogen activation assay on the surface of THP-1 monocytes. THP-1

monocytes, seeded at a density of 200,000 cells per well, were incubated with 0.6μM

plasminogen, 15nM tPa, and a fluorescence substrate of plasmin (H-D-Val-Leu-Lys-7-

amido-4-methylcoumarin) in the presence and absence of wild-type and mutant apo(a).

Plasmin formation was monitored for 1h at 370C by spectrometry; excitation: 370 nm,

emission: 470 nm, and cutoff: 455 nm. The rate of plasminogen activation was

represented as the initial slope of the plot of RFU against min2 and normalized against

control. *: p < 0.05 by two-tailed ANOVA. The data represents the means ± SEM of three

independent experiments each conducted in triplicate.

Page 59: Characterization of the I4399M Variant of Apolipoprotein(a

44

Finally, at an apo(a) concentration of 1.5μM, the t50% for the wild-type and I4399M

variant were ~8X greater (5867 seconds) and ~9.5X greater (6623 seconds) than control,

respectively. As previously observed, increasing concentration of apo(a) results in a

greater inhibition of fibrinolysis compared to no apo(a) control. Interestingly, at all three

concentrations of apo(a), the I4399M variant significantly increases the inhibition of

fibrinolysis compared to wild-type.

As the previous results suggest, the increase in inhibition of fibrinolysis seen with

the I4399M mutant compared to wild-type can not be explained by the inhibition of

plasminogen activation on the fibrin surface. Therefore, we investigated how the apo(a)

variants affect clot formation through an absorbance based coagulation assay. The assay

was conducted in a similar manner to the previous assay in the absence of tPA and the

change of absorbance was measured by spectrometry at 370C for 1h. The absorbance

profiles were generated and the difference between the maximum and minimum

absorbance, representing the turbidity of the sample, was calculated. The ½ time required

for clot formation were reported (Figure 3.7). Looking at the sample turbidity, at all three

concentrations of apo(a), the wild-type variant shows ~11% increase in absorbance

compared to no apo(a) control, although the difference is not statistically significant.

However, the I4399M variant was shown to significantly increase the sample turbidity at

0.5μM, 1μM, and 1.5μM of apo(a) by 20%, 25%, and 44%, respectively, compared to the

wild-type apo(a) variant. Similarly, the observed ½ clotting time for the wild-type variant

at the three apo(a) concentrations significantly decreased by 23%, 30%, and 34%,

respectively, compared to the no apo(a) control.

Page 60: Characterization of the I4399M Variant of Apolipoprotein(a

45

Figure 3.6 Absorbance-based fibrinolysis assay with LPDP in presence and absence

of apo(a). To study clot formation and subsequent lysis, lipoprotein-deficient plasma

(LPDP) was clotted with 1 U/mL thrombin (lla), 20 mM CaCl2, and 3nM tPA in the

absence and presence of wild-type and mutant apo(a) and analyzed by spectroscopy at

405nm. The changes in turbidity were measured at 37oC for 3 hr, normalized against

maximum absorbance and plotted as a function of time for A. 0.5μM, B. 1μM, and C.

1.5μM of apo(a). D. The time to reach 50% lysis (t50%) at the three concentrations were

reported. *: p < 0.05 versus wild-type, **: p < 0.02 versus no apo(a) by two tailed

ANOVA. The data represent the means ± SEM of three independent experiments each

conducted in triplicate.

Page 61: Characterization of the I4399M Variant of Apolipoprotein(a

46

Figure 3.7 Absorbance-based coagulation assay with LPDP in presence and absence

of apo(a). To study clot formation, lipoprotein-deficient plasma (LPDP) was clotted with

1 U/mL thrombin (lla), 20 mM CaCl2, in the absence and presence of wild-type and

mutant apo(a) and analyzed by spectroscopy at 405nm. The changes in turbidity were

measured at 37oC for 1 hr, normalized against minimum absorbance and plotted as a

function of time for A. 0.5μM, B. 1μM, and C. 1.5μM of apo(a). D. The turbidity of the

solution and E. The time to reach 50% coagulation at the three concentrations were

reported. *: p < 0.05 versus wild-type by two tailed ANOVA, +: p < 0.05 versus no

apo(a) by t-test, ++: p < 0.05 versus wild-type by t-test. The data represent the means ±

SEM of three independent experiments each conducted in triplicate.

Page 62: Characterization of the I4399M Variant of Apolipoprotein(a

47

Interesting, it was seen that the I4399M variant significantly decreased the ½ clotting

time compared to the wild-type at all concentrations of apo(a). The resulting ½ clotting

times decreased by 27%, 35%, and 37% in the presence of the I4399M variant,

respectively, compared to wild-type.

Overall, these results suggest that while the I4399M variant does not differ in the

degree of inhibition of plasminogen activation on the fibrin surface, it does alter the clot

properties compared to wild-type apo(a), seen by the increased t50% values and increased

turbidity of the sample, as well as decreased clotting time. Therefore, the next step was to

understand how this mutant affects the fibrin clot architecture in a way that results in the

observed affects on fibrinolysis and coagulation.

3.4 The Relationship Between the I4399M Variant and Fibrinogen/Fibrin Clot

Architecture

In order to understand the increased pro-thrombotic potential of the I4399M

variant, the effect of the variant on the fibrin clot architecture was studied. This was

achieved through scanning electron microscopy (SEM) of fibrin clots formed in the

absence or presence of the I4399M and wild-type variant. Clots were formed on SEM

pegs and subjected to low-vacuum. Images were captured at 15,000X and 30,000X

magnification. The 30,000X images were analyzed using the imaging software ImageJ

and the average width of a minimum of 25 fibers were reported (Figure 3.8). It was found

that clots containing the wild-type apo(a) possessed a greater number of fibers compared

to no apo(a) control, while the fiber width was not significantly different than the control.

However, the I4399M clots had a much more disorganized structure and possessed

multiple nodules throughout the clot, resulting from aggregated fibrin fibers.

Page 63: Characterization of the I4399M Variant of Apolipoprotein(a

48

Figure 3.8 SEM imaging of fibrin clots formed in the absence or presence of either

the I4399M or wild-type apo(a). A. Purified fibrinogen in the absence and presence of

wild-type and mutant apo(a) was clotted with 1U/mL lla and 20mM CaCl2 and visualized

using environmental scanning electron microscopy (SEM). Images were taken at 15,000x

magnification with a HFW of 6.67 microns. B. Fiber widths from 30,000X images of the

clots from A were calculated using the imaging software ImageJ, with a minimum of 25

fibers measured per field of view. *: p < 0.05 by two tailed ANOVA. The data represent

the means ± SEM of four separate images from the same clot.

Page 64: Characterization of the I4399M Variant of Apolipoprotein(a

49

Figure 3.9 Imperial stain of samples of I4399M and wild-type apo(a) variant in

solution with fibrinogen. A. 3 mg/mL of purified human fibrinogen was incubated with

3μM of wild-type and I4399M apo(a) for 10 minutes at RT. Samples were vortexed and

the I4399M mixture was centrifuged. The mixtures and resulting I4399M mixture pellet

and supernatant were run on a 4-15% gradient SDS-PAGE gel and imaged with Imperial

stain.

Page 65: Characterization of the I4399M Variant of Apolipoprotein(a

50

The resulting fibers were ~46% thicker than no apo(a) control and wild-type apo(a).

When optimizing the various plate based assays with purified fibrinogen, high

concentrations of the 17K I4399M variant (~3μM) resulted in a turbid mixture, a result

not seen with the wild-type 17K apo(a) at the same concentration. Therefore, in order to

understand this observation, purified fibrinogen was incubated with 3μM of both apo(a)

variants and the resulting I4399M mixture was centrifuged to separate any insoluble

material. The supernatant and pellet were run on an SDS-PAGE gel and Imperial stained

(Figure 3.9). The 17K and 17K I4399M variants in solution with fibrinogen were

effectively visualized, showing the three bands of fibrinogen resulting from treatment

with a reducing agent. Following centrifugation, the supernatant of the 17K I4399M and

fibrinogen mixture contained a strong band for apo(a), but very weak bands for

fibrinogen. Surprisingly, most of the fibrinogen in the mixture was found in the pellet,

along with some apo(a). Overall, we have shown that the I4399M variant is able to alter

the fibrin clot network in a way that differs from wild-type and is able to precipitate

soluble fibrinogen.

3.5 Analysis of Clot Structure and Properties of Plasma from Homozygotes of the

rs3798220 SNP

While the investigation up to this point was focused on the purified mutant apo(a)

component, we then wanted to examine how a Lp(a) molecule containing the Met

mutation would affect the fibrin clot structure and its subsequent properties. We obtained

4 patient samples that are homozygous carriers of the rs3798220 SNP and the samples

were phenotyped to determine apo(a) isoform size. We then screened multiple people to

identify wild-type controls; these samples were genotyped to determine if they possessed

Page 66: Characterization of the I4399M Variant of Apolipoprotein(a

51

the SNP and phenotyped to match the isoform sizes (data not shown). We were able to

match one homozygote, which was a low Lp(a) expresser ([Lp(a)] ~ 5 mg/dL), with a

non-carrier of similar Lp(a) concentration and isoform size ([Lp(a)] ~ 4 mg/dL, Size =

23K). Similarly, we were able to match a highly expressing homozygote ([Lp(a)] ~ 40

mg/dL) to a single non-carrier with similar Lp(a) concentration and isoform size ([Lp(a)]

~ 35 mg/dL, Size = 14K).

The first experiment using these samples was a clot permeability assay to

determine the effect of the mutation on fibrin clot pore size. Previous studies have shown

that Lp(a) is able to produce thinner fibers with tighter pores, making the clot less

permeable and less susceptible to lysis (53). In our study, clots from homozygotes and

non-carriers were formed in glass tubes, hooked to a buffer reservoir, and the time

between drops and the weight of the drops were recorded. These values were placed into

a derived equation and the values are reported as Darcy constants (ks), with units of cm2,

representing the surface area of the pores in the clot (Figure 3.10). For the low [Lp(a)]

samples, the non-carrier and homozygote samples had a Darcy constant of 7.10x10-9 cm2

and 7.52x10-9 cm2, respectively. For the high [Lp(a)] samples, the non-carrier had a

Darcy constant of 4.51x10-9 cm2, while the homozygote had a Darcy constant of 5.89x10-

4.53x10-9 cm2. This suggests that there is no significant difference in pore size between

the homozygote and non-carrier clots at both the low and high Lp(a) concentrations.

Page 67: Characterization of the I4399M Variant of Apolipoprotein(a

52

Sample [Lp(a)] (mg/dL) Phenotype

Low [Lp(a)] Non-Carrier 4.55 23K

Homozygote 4.46 24K

High [Lp(a)] Non-Carrier 35.71 18K

Homozygote 46.68 17K

Figure 3.10 Permeation and fibrinolysis values of patient plasma samples. A. Plasma

from paired patients that were non-carriers as well as patients homozygous for the

polymorphism were clotted with 1U/mL lla and 20mM CaCl2 in a clear glass tube and

allowed to mature in wet box for 2hrs. The tubes were then connected via tubing to a

reservoir containing a buffer (0.01M Tris, 0.1M NaCl, pH 7.5) and the volumes flowing

through the fibrin gels were used to calculate darcy constants (ks). B. Lipoprotein-

deficient plasma (LPDP) in the presence or absence of non-carrier or homozygote plasma

samples were clotted with 1 U/mL thrombin (lla) and 20 mM CaCl2 and analyzed by

spectroscopy at 405nm, measuring the changes in turbidity at 37oC for 3 hr. The time

needed to reach ½ lysis time was compared to LPDP control and presented as relative ½

lysis time. *: p < 0.05 by two tailed ANOVA. The data represent the means ± SEM of

three independent experiments each conducted in triplicate.

Page 68: Characterization of the I4399M Variant of Apolipoprotein(a

53

We were able to show that purified I4399M mutant can inhibit fibrinolysis to a

greater degree than wild-type apo(a). Therefore, we conducted the same fibrinolysis assay

using these homozygote and non-carrier samples to determine if the same result can be

observed. Samples were spiked with tPA and clotted with CaCl2 and lla and the change of

absorbance was measured for 3 hours. The t50% times were reported, normalized against

the LPDP control with no Lp(a) and reported as relative lysis times compared to control

(Figure 3.10). As anticipated, the samples with low Lp(a) concentrations had similar lysis

times; the non-carrier and homozygote samples had relative t50% times of 1.215 and 1.219,

respectively. As for individuals with high Lp(a) concentrations, the non-carrier had a

relative t50% time of 1.354, while the homozygotes had significantly inhibited fibrinolysis

to a greater degree than the non-carrier, resulting in a relative t50% time of 1.519. Overall,

these patient samples have shown that low expressers of Lp(a), regardless of carrier

status, have similar permeation and fibrinolysis values, while homoygotes that have high

Lp(a) concentrations have similar Darcy constants than non-carriers, but are able to

significantly increase inhibition of fibrinolysis compared to wild-type. Based on these

observations, it was imperative to visualize these clots to see if the clot architecture is

altered in a way similar to the results seen with the purified apo(a) variants.

Confocal microscopy and SEM was used to visualize fibrin clot architecture in the

homozygote and non-carrier clots. For confocal microscopy, clots were made in glass

tubes similar to the permeability assay and spiked with fluorescently labeled fibrinogen.

Mature clots were washed, placed on slides and imaged at 60X. The resulting images

were analyzed using ImageJ software using a designed macro, which overlays a grid to

measure the number of intersecting fibers in a given field of view. The results are the

averages of the counts from 6 different images of the same clot and the values are

Page 69: Characterization of the I4399M Variant of Apolipoprotein(a

54

reported as the number of fibers per 100μm (Figure 3.11). At low concentrations of Lp(a),

both the non-carrier and carrier samples have nearly identical number of fibers per

100μm. As expected, at a high concentration of Lp(a), the non-carrier has a greater

number of counted fibers, resulting in a denser clot with tighter pores. However, the

homozygote samples do not follow this same type of trend and have significantly lower

numbers of fibers compared to the non-carrier samples. This would suggest that the clots

produced by the carrier plasma are not as dense and may possess larger pores. To expand

on this result, SEM was utilized to measure the fiber width of the clots. This was

conducted similarly to the confocal procedure, without the addition of labeled fibrinogen.

The paired homozygote and non-carrier samples were washed and visualized under low

vacuum at 15,000X and 30,000X magnification. Using the ImageJ software, the widths of

at least 25 fibers of 6 different images of the same clot were calculated and the average

fiber width was reported in μm (Figure 3.11). The samples with low Lp(a) concentrations

showed that the carriers had significantly thicker fibers compared to the non-carriers, with

average fiber widths of 0.066μm and 0.054μm, respectively. Surprisingly, the high

[Lp(a)] non-carrier sample was shown to have ticker fibers than the non-carrier with a

lower Lp(a) concentration. However, as anticipated, the high [Lp(a)] carrier had

significantly thicker fibers compared to both the high [Lp(a)] non-carrier and the low

[Lp(a)] carrier samples. Overall, these results would suggest that homozygous carriers of

the apo(a) mutation have the same permeability as non-carrier clots but possesses less

fibers in a given field of view, as suggested by the permeation and confocal data.

However, the homozygote clots have thicker fibers and has the capability to inhibit

fibrinolysis to a greater degree than non-carriers.

Page 70: Characterization of the I4399M Variant of Apolipoprotein(a

55

Figure 3.11 Confocal microscopy analysis of clot formation using homozygote and

non-carrier plasma samples. Plasma samples from homozygous carriers and non-

carriers of the rs3798220 SNP were clotted with 20mM CaCl2 and 1U/mL of IIa and

were supplemented with 0.3mg/mL Alexa Fluor® 568 labelled fibrinogen. Clots were

washed for one hour and transferred to a microscope slide. The resulting fluorescent

clots were imaged on an Olympus FV1000 confocal microscope at 60X magnification.

A. Representative images of the micrographs from confocal microscopy of the plasma

samples at 60X magnification. B. ImageJ software was used to analyze the images and a

collaborator-designed macro was used to determine the number of fibers per 100μm. *:

p < 0.05 by two tailed ANOVA. The data represent the means ± SEM of 6 different

images of the same clot.

A.

B.

Page 71: Characterization of the I4399M Variant of Apolipoprotein(a

56

Figure 3.12 Scanning electron microscopy and fiber width measurements in clots

formed from homozygote and non-carrier patient samples. Plasma samples from

homozygous carriers and non-carriers of the rs3798220 SNP were clotted with 20mM

CaCl2 and 1U/mL of IIa. Clots were fixed, washed for one hour and transferred to a SEM

pegs. The resulting clots were imaged on an field-emission environmental SEM under

low vacuum at 15,000X and 30,000X magnification. A. Representative images of the

micrographs from SEM of the plasma samples at 15,000X magnification. B. Fiber widths

from 30,000X images of the clots from A were calculated using ImageJ, with a minimum

of 25 fibers measured per field of view *: p < 0.05 by two tailed ANOVA, ++: p < 0.05 vs

non-carrier by two tailed ANOVA. The data represent the means ± SEM of 6 different

images of the same clot.

A.

B.

Page 72: Characterization of the I4399M Variant of Apolipoprotein(a

57

3.6 Identification of a Modification on the Substituted Methionine Residue in the

I4399M Variant

The molecular dynamics data suggested that the substituted methionine residue is

exposed to the environment, possibly becoming susceptible to potential modification.

Previous studies have shown that methionine residues can undergo oxidative modification

to methionine sulfoxide, which plays a role in many cellular functions (163). Therefore,

we investigated whether this exposed Met on the apo(a) residue can be modified to a

methionine sulfoxide moiety. We conducted mass spectroscopy of the 4399Met mutation

in a smaller variant of apo(a), containing KIV10, KV and inactive protease domains

(Figure 3.12). Two theoretical masses of the tryptic peptide containing the mutation were

calculated, possessing either an unmodified Met residue and or an oxidized Met residue

(2505.2595 and 2521.2544, respectively). The actual MS spectra contains no observable

peak for the peptide of mass 2505.2595 and a major peak that corresponds to the peptide

of mass 2521.2544, indicating that the presence of this 4399Met mutation in

apolipoprotein(a) is susceptible to oxidative modification and can result in a methionine

sulfoxide moiety on this residue.

Page 73: Characterization of the I4399M Variant of Apolipoprotein(a

58

Figure 3.13 Mass spectroscopy of a truncated apo(a) variant containing the I4399M

mutation. A truncated recombinant apo(a) was constructed containing the KIV10, KV,

and inactive protease domains, either with (denoted KIV10-P I4399M) with and without

(denoted KIV10-P) the polymorphism. The KIV10-P I4399M was subjected to in-gel

tryptic digestion and then analyzed by MALDI-TOF mass spectrometry using a Waters

SYNAPT G2-Si instrument. The spectra were observed using MassLynx software and

compared to two theoretical masses for the potential tryptic peptide.

Page 74: Characterization of the I4399M Variant of Apolipoprotein(a

59

Chapter 4

Discussion

4.1 The I4399M Variant Secretion Rate Does Not Differ from Wild-Type

In this study, we utilized the human heptoma cell line, HepG2, in order to analyze

the secretion rate of recombinant 17-kringle containing variants of apo(a). This cell line

was deemed appropriate for our in vitro assay as it has been shown that lipoproteins, and

specifically Lp(a), is synthesized by hepatic cells in the liver (164, 165). In the

hepatocytes, apo(a) is initially formed as a low molecular mass precursor, and then

modified in the endoplasmic reticulum (ER) before entering the Golgi and becoming

glycosylated (166). Therefore, using a cell line other than hepatocytes would not produce

the mature apo(a) structure with the proper modifications, which could ultimately affect

its secretion rate from that specific cell.

In our study, we found that when comparing two apo(a) variants of the same

length, the mutation itself can not alter the secretion rate from hepatocytes. When looking

at the imges for the wild-type and I4399M variant, it is seen that the apo(a) in the lysates

reaches maximum at 30 minutes and then the intensity decreases as the time of the assay

increases. This decrease in intensity in the lysates is opposed by an increase in the media

signal for both variants as the assay progressed. The maximum signal for both variants in

the media was seen at 240min, and a dip in band intensity was seen at 480min, which

most likely due to degradation of the mature apo(a). We use protease inhibitors in our

lysis buffer to prevent degradation during immunoprecipitation, but it is not used on the

media samples. During the assay, the samples are kept on ice to prevent enzymatic

Page 75: Characterization of the I4399M Variant of Apolipoprotein(a

60

degradation, so one would expect a minimal amount of degradation from the assay.

However, the last time point is collected after 8 hours and the stress caused by the assay

could cause the release of inflammatory cytokines from the hepatocytes (167) that could

cause the recruitment of proteases which would lead to degradation of apo(a).

Since it was determined that the mutation alone can not account for the elevated

Lp(a) levels in carriers, there are two likely explanations that could account for this

observation. First, carriers of this mutation have been correlated with smaller isoform

sizes (99) and smaller isoforms are associated with higher Lp(a) protein levels (84, 85).

Therefore, when comparing levels in a population, carriers possessing smaller apo(a)

isoforms would have higher levels than non-carriers with larger isoform sizes. Secondly,

the I4399M mutant may have a longer turnover time than the wild-type variant or may

not be as efficiently catabolized, resulting in increased plasma Lp(a) concentration.

Further studies must be conducted to analyze the internalization of this I4399M variant

and its subsequent turnover rate compared to wild-type. Furthermore, the mutation can be

introduced to a smaller variant and its secretion rate compared to wild-type variant of the

same size could be determined to confirm the results of our study.

4.2 Analysis of the I4399M Variant Effect on Plasminogen Activation on Different

Surfaces

Previous studies have shown that Lp(a)/apo(a) has the capability to inhibit

plasminogen activation on a fibrin surface by binding to free lysines on fibrin (71) to

prevent tPA-mediated plasminogen activation and on the cell surface by binding to cell

surface plasminogen receptors (134). Similarly, it was also shown that the strong LBS of

KIV10, the KV domain, and the amino-terminal amino acids were critical for the

Page 76: Characterization of the I4399M Variant of Apolipoprotein(a

61

inhibition of plasminogen activation in the fibrin surface (71) and removal of the strong

LBS in KIV10 or the KV domain blunts the inhibition of activation on the cell surface

(134). Therefore, one would predict that a mutation in the protease domain would not

result in a change in the inhibition of plasminogen activation on these surfaces.

As predicted, the I4399M variant inhibited activation to the same degree as wild-

type on a fibrin surface at multiple apo(a) concentrations. The assay relied on a quenching

of a fluorescent signal to calculate the rate of plasmin formation, which is much more

efficient for this use than an absorbance based assay. An absorbance assay would rely on

decreasing turbidity of the clot as it is lysed and would not give information on the rate of

plasminogen cleavage. On the other hand, using fluorescence would allow one to

calculate the amount of plasmin formed per mole of tPA (71). The derived equation takes

into account the background fluorescence within the spectrometer to try to eliminate the

error that is associated with fluorescence-based assays. While the predicted result was

seen on the fibrin surface, an unexpected result was seen on the surface of THP-1

monocytes. Wild-type apo(a) showed a 30% decrease in activation by competing with

plasminogen for binding sites, which agrees with literature (134, 168), while the I4399M

variant showed no inhibition of activation compared to the no apo(a) control. This

fluorescence-based assay relies on quenching of fluorescence as well, where a fluorescent

substrate of plasminogen is cleaved and results in quenching. There are two potential

reasons for this blunting of inhibition seen with the I4399M variant. First, the fluorescent

substrate could be binding to the I4399M variant, and thus making it unavailable to

plasminogen, or it can become post-transnationally modified, therefore preventing it from

entering plasmin’s active site. In order for this to happen, the mutated residue in this

variant would have to play a role in this interaction as the two apo(a) variants used are of

Page 77: Characterization of the I4399M Variant of Apolipoprotein(a

62

the same length. Secondly, the mutated residue could be interfering with apo(a) binding

to the cell surface receptors, leaving more receptors open for the plasminogen conversion.

Previous studies have shown that a point mutation in KIV10 domain can result in blunting

of inhibition as well (134, 168), and while the I4399M mutation is not occurring in one of

the domains needed for binding, it could be acting through a different mechanism to

prevent plasminogen binding to the cell surface. Overall, this data would suggest that the

inhibition of plasminogen activation on the fibrin surface cannot account for the increased

lysis time and altered coagulation seen with this I4399M variant.

4.3 I4399M Variant Alters Fibrinolysis Time and Coagulation Compared to Wild-

Type

It has been shown that Lp(a)/apo(a) has the ability to inhibit the rate of fibrinolysis

in vitro in a dose-dependent manner and there is evidence that isoform size may also

contribute to the degree of inhibition (161, 162, 169). It achieves this through preventing

the conversion of Glu- to Lys-plasminogen, as well as blunting inhibition of the

plasminogen activator inhibitor-1 (162). We were able to show that the I4399M variant

was able to significantly increase fibrinolysis time compared to wild-type at all three

concentrations of apo(a) in a dose dependent manner. Taking the plasminogen activation

data into account, this suggests that the I4399M variant is achieving this response in a

mechanism separate from inhibiting plasminogen activation. It has been shown that

thinner fibrin fibers lyse quicker than thick fibers, but apo(a) interacting with fibrinogen

results in thinner fibrin fibers in a denser fiber network, decreasing permeability and

resulting in increased lysis time (52, 54). However, a recent study has shown that thicker

fibers are more likely to elongate at first rather than being lysed, compared to thinner

Page 78: Characterization of the I4399M Variant of Apolipoprotein(a

63

fibers which are immediately lysed, resulting in increased lysis time (170). Therefore, it

could be suggested that the potential explanation for the effect of the I4399M variant is

that the mutant creates a tighter, denser clot that is less permeable and less susceptible to

lysis. On the other hand, the mutation could cause thicker fibrin fibers, which are less

susceptible to lysis and result in the increased lysis time. Further studies were needed on

the structure of the fibrin clot to understand this result.

To further analyze this result, coagulation assays were conducted in the absence of

tPA to understand if the I4399M variant alters the way the clot is formed. Literature has

shown that many factors influence how a fibrin clot is formed, such as thrombin

concentration, pH, ionic strength, and concentrations of calcium (171-173). Higher

concentrations of IIa result in thicker fibers and a less dense network, while low IIa

concentrations result in thinner fibers and a more dense network (170). The assay is

characterized as possessing a “lag” phase with no change in turbidity as the protofibrils

are initially formed (16). As the protofibrils laterally aggregate, the turbidity begins to

increase. The magnitude of turbidity relates to the structure of the fibrin network, with

thicker fibers causing a greater increase in final turbidity (90, 174). In our assay,

controlling the amount of IIa added and other conditions would allow us to see changes

caused by the apo(a) variants. We showed that the ½ coagulation time of the I4399M

variant was significantly shorter compared to wild-type and the clots of the I4399M

variant were significantly more turbid that the wild-type variant. The greater turbidity of

the I4399M variant clot would suggest a different clot architecture than the wild-type and

could insinuate that the clot is made of thicker fibers, causing the higher turbidity of the

sample. Overall, the results from the fibrinolysis and coagulation assay indicated that the

fibrin network may be different for the clots of the I4399M variant compared to wild-type

Page 79: Characterization of the I4399M Variant of Apolipoprotein(a

64

that could account for the differences seen. The next step was to visualize the fibrin clot

to understand how the I4399M variant affects clot architecture.

4.4 The I4399M Alters Fibrin Clot Architecture and Causes Precipitation of

Fibrinogen

As previously discussed, Lp(a)/apo(a) has been shown to result in an altered clot

structure, possessing thinner fibers in a more dense network that results in increased lysis

time (175, 176). However, our data would suggest that the increased lysis time and

increased turbidity of the clot seen with the I4399M variant may be a result of different

clot architecture compared to wild-type. SEM was used to image the clots due to its high-

resolution capabilities and clots being made in the absence of cells, allowing the clots to

be effectively washed and therefore visualized. The wild-type variant resulted in a denser

clot with thinner fiber widths (Average = ~75μm). Unexpectedly, the I4399M variant

resulted in a less dense clot with fibers that were significantly thicker than wild-type

(Average = ~312μm). As well, the I4399M variant produced a clot that did not form the

organized pore network seen with the wild-type variant. It was also characterized as

possessing multiple fibrin nodules at the branching points of the fibrin network. The

thicker fibers and nodules can account for the greater turbidity of the solution seen in the

clotting absorbance assay. Although the I4399M clot looks less dense than the wild-type

clot, the thicker fibers and nodules at the branching points could perhaps explain the

increased lysis time in the mechanism described by Bucay et al. (170), where thicker

fibers elongate and delay lysis times. Our study did not investigate the interaction of this

mutant variant with other proteins or factors in the coagulation cascade, but the data

Page 80: Characterization of the I4399M Variant of Apolipoprotein(a

65

suggests that I4399M variant is able to alter the fibrin clot network to account for its

increased prothombotic potential.

Through optimization of the previous assays, when high concentrations of 17K

I4399M are incubated with fibrinogen, the mixture became very turbid. The mixture was

centrifuged, run on a gel, and Imperial stained to visualize the protein. Imperial stain was

used over a silver stain due to its specificity and purity, while an immunoblot of the

samples would be ineffective as the precipitate was unknown. The staining showed that

after centrifugation, almost all the fibrinogen in the mixture is in the pellet while the

majority of 17K I4399M remained in the supernatant. This suggests that the apo(a) is

interacting with the soluble fibrinogen and causing it to precipitate. The apo(a) band in

the pellet could be the result of the apo(a) that is bound to the exposed lysines of the

fibrinogen molecules or potentially co-precipitation of apo(a) with the fibrinogen

molecule. Precipitation of soluble fibrinogen is usually only achieved through a rapid

heat-precipitation technique (177) or in the presence of a high concentration of Ca2+ on a

fibrin gel (178). Since the I4399M variant only differs by the single Met residue

compared to wild-type, this residue could potentially be interacting with the fibrinogen

molecule, although this has to be further analyzed. Taken together, the data suggest that

the I4399M variant interacts with fibrin that results in a unique clot network with thicker

fibers and aggregated nodules at branching points, and is able to cause precipitation of

fibrinogen at high concentrations.

Page 81: Characterization of the I4399M Variant of Apolipoprotein(a

66

4.5 Lp(a) from Homozygous Carriers of the SNP Affect Fibrin Clot and Its

Properties Similar to I4399M Apo (a) Variant

The data generated in the preceding assays were in the presence of purified 17K

I4399M apo(a). Although the effect of Lp(a) should be studied in parallel with the apo(a)

data, samples from carriers of this SNP were difficult to obtain due to the low frequency

of this mutation, and we currently do not possess a method to produce Lp(a) in vitro. We

were able to obtain 4 individual samples from homozygous carriers of the SNP to analyze

with our biochemical assays. The 2nd patient sample was very turbid upon arrival,

signifying the plasma was coagulated and therefore was not used. The other three samples

were phenotyped through immunoblotting to determine their isoform size. Previous

studies on this mutation have used patient samples in various biochemical assays and

accounted for race, gender, and ethnicity, but did not match the carriers with non-carriers

of similar Lp(a) levels and isoform sizes . The results of these studies could therefore be a

result of the association of the mutation with higher levels of Lp(a), accounting for the

increased risk, rather than observing the direct effect of the mutation itself. Therefore, we

screened multiple people for the SNP by genotyping to determine suitable non-carrier

controls. These samples were phenotyped and the Lp(a) levels were determined via

ELISA and matched to carrier samples with similar isoform sizes and Lp(a) levels. We

were able to match a non-carrier to a carrier with very low Lp(a) levels and an isoform

size of 23K. Similarly, we were able to match two carrier samples to a single non-carrier

sample with very high Lp(a) levels and an isoform size of 17K. While more patient

samples should be used to generate more complete conclusions, matching the non-carrier

and carrier samples together can provide insight into effect of the I4399M Lp(a) on the

various prothrombotic mechanisms.

Page 82: Characterization of the I4399M Variant of Apolipoprotein(a

67

We started with permeation of the clots in glass tubes to determine the Darcy

constants, representing the area of the pores in the clot. As previously described, wild-

type Lp(a)/apo(a) has been shown to result in a tighter, denser clot with a smaller Darcy

constant, which relates to its ability to inhibit fibrinolysis (19, 52, 54). The matched

homozygote and non-carrier pair with low Lp(a) levels showed no difference in Darcy

constant, indicating the Lp(a) levels were too low to impact the fiber network. At the

higher Lp(a) concentration, the homozygote and non-carrier had similar Darcy constants

that were both significantly lower than the samples with low Lp(a) levels. This would

suggest that the Met variant does not significantly alter the clot permeability compared to

wild-type, regardless of Lp(a) levels. However, images from SEM shows the I4399M

apo(a) variant causes significantly thicker fibers than clots with non-carrier plasma. We

conducted the previously described fibrinolysis assay with the paired plasma samples.

Following data generation, the samples were controlled against a no-apo(a) control and

shown as the relative increase in clot lysis time compared to the control. The non-carrier

and homozygote pair with low Lp(a) levels did not show any difference between each

other in relative lysis time, agreeing with the permeation result. The samples with high

Lp(a) levels both had increased relative lysis times compared to the low Lp(a) levels,

however, the homozygote had a significantly longer relative lysis time compared to the

non-carrier. Taking the permeation and fibrinolysis data together, the homozygote with

high Lp(a) levels agrees with the data from the assays in the presence of purified I4399M

apo(a), suggesting the homozygous clots have thicker fibers in an altered fibrin network,

contributing to the increased lysis times.

To strengthen the data set, the clots of these homozygote and non-carrier pairs

were visualized using confocal microscopy and SEM. Confocal microscopy was utilized

Page 83: Characterization of the I4399M Variant of Apolipoprotein(a

68

first to generate a larger field of view of the clot so that the fiber densities of the clots

could be determined. Labeled-fibrinogen was added to the mixture prior to clotting so that

it accounted for 1/100 of the total volume as to prevent over-saturating with fluorescent

signal. Following the generation of the images, ImageJ was utilized with the collaborator-

designed macro to determine clot density. The macro overlays a 10x10 grid over the

image and counts the number of fibers that intersects this grid, giving values of the

number of fibers per 100μm. In order to eliminate background noise from the actual

fibers, the images were first converted to binary before analysis. As anticipated from the

clot permeation results, the homozygote and non-carrier pair with low Lp(a) levels had

similar fiber counts, indicating the clots had comparable densities. On the other hand, for

the samples with high Lp(a) concentrations, the non-carrier had significantly more fibers

in a given field than the homozygote carriers. While the resulting Darcy constants of the

wild-type and mutant clots were comparable at both Lp(a) concentration intervals, the

homozygote with high Lp(a) levels clots had significantly lower fiber counts compared to

the non-carrier clot with high Lp(a) levels. While the presence of less fibers in the carrier

clots would suggest much permeable fibrin clots, the comparable Darcy constants

indicates that the fibers in the mutant clots are altered in such a way to account for this

discrepancy. It must be stated that confocal imaging of fibrin clots using fluorescently

labeled fibrinogen is somewhat flawed, as it does not completely constitute all the fibers

within the clot (only 1/100 of the volume) and it is much more difficult to determine fiber

edges using a light based method. Therefore, SEM was also used as it provides more

accurate results in terms of quantitative measurements, not only due to the difference in

how images are generated, but the increased magnification compared to confocal

microscopy. In order to take SEM images of the plasma clots, the clots needed to be

Page 84: Characterization of the I4399M Variant of Apolipoprotein(a

69

intensely cleaned to remove any debris or excess ions that may interfere with the imaging

(179). Therefore, the clots were made in glass tubes, connected to a reservoir to permeate

the clot and then removed from the tube onto a SEM peg. In keeping with the results seen

from the SEM images with purified apo(a), the homozygote with low Lp(a) levels had

significantly thicker fibers than in the non-carrier clot. The same trend was also seen in

the pair with the high Lp(a) levels, with the homozygote having significantly thicker

fibers than the non-carrier. Finally, both the homozygotes and non-carriers had thicker

fibers at higher Lp(a) levels compared to low Lp(a) levels (p < 0.05 and p < 0.01,

respectively). These data can be considered to be more accurate than the confocal as the

SEM is able to differentiate the edges of the fibers.

Overall, the data from the Lp(a) of homozygote carriers of the I4399M mutation

agree with the data from the apo(a) I4399M variant. The clots made from this variant are

characterized as having altered fibrin networks, with thicker fibers and nodules at the

branching points of the clot. The clots from the homozygotes have less fibers than the

non-carriers at high Lp(a) levels, but have comparable Darcy constants and result in

increased clot lysis times compared to wild-type. This may be explained by the fibers

elongating and delaying lysis, thereby increasing lysis time (170) or by some other

unknown mechanism. In terms of the risk of severe CAD associated with this mutant, we

suggest that the pathogenic effects seen are a result of the I4399M variant affecting the

fibrin clot network, resulting in increased clot lysis time. As well, there may be a role for

the isoform size of carriers of this SNP, which can contribute to this effect.

Page 85: Characterization of the I4399M Variant of Apolipoprotein(a

70

4.6 The I4399M Variant Possesses a Methionine Sulfoxide Moiety Which May

Contribute to the Prothrombotic Potential of this Variant

The preceding data suggests that the mutated Met residue in the I4399M variant of

apo(a) may be contributing to the prothombotic potential of this variant. Although

unmodified Met residues are not usually involved in signaling or other pathways, exposed

Met residues have the capability to become oxidized to methionine sulfoxide moieties,

which can have many physiological consequences on signaling or fibrin clot properties

(180). It has also been determined that several important proteins in vascular biology have

been shown to contain oxidative-sensitive methionine residues that have potential

regulatory roles in pathogenesis of vascular of thrombotic diseases (180). Therefore, we

investigated whether the exposed Met mutant in the I4399M variant can be modified by

oxidation.

The molecular dynamic simulations were conducted on the wild-type and I4399M

variant of apo(a); the protease domains were exposed to hydration and average occupancy

images were generated. In the wild-type apo(a), the Ile residue remained buried

throughout the simulation with a proximal Arg residue at 4420 acted as a cap. In the

I4399M variant, the 4420Arg residue opens, exposing the 4399Met residue to the

environment. The RMSD analyzes the relative change in distance between two residues

over a given time. The analysis showed that in this open conformation, there was minimal

movement between the Arg and Met residues, suggesting that they may be interacting and

coordinating this open conformation. While this coordination the 4420Arg residue was

not confirmed, it gave the basis that the Met residue is exposed for potential modification.

To determine if a methionine sulfoxide moiety is present in the I4399M variant,

we conducted mass spectroscopy on a truncated version of apo(a). This variant only

Page 86: Characterization of the I4399M Variant of Apolipoprotein(a

71

possessed the KIV10, domain, the KV domain, and the protease domain containing the

mutant Met residue. This truncated variant was require for mass spec in order to produce

a clear signal with less tryptic peptides to analyze. The peptide containing the mutation

was a good flyer for mass spectroscopy and two theoretical masses for this peptide were

generated, one for the unmodified Met residue, and one possessing the methionine

sulfoxide moiety. The spectra showed an intense peak for the theoretical mass with the

Met sulfoxide moiety and did not produce a visible peak for the other theoretical mass,

therefore confirming the ability of the I4399M mutant to possess a methionine sulfoxide

moiety. This result could help explain the pathogenic affects of this I4399M mutant, as

oxidative modification can be transferred to other proteins through redox reactions (180).

Studies have shown that fibrinogen can be modified by oxidation through environmental

factors like smoking, altering the fibrin clot properties of these individuals (181). This

suggests that the fibrinogen itself may be directly interacting with the methionine

sulfoxide residue in the I4399M variant to become oxidized, or it may affect fibrinogen in

a mechanism that differs from the previously described ones. Therefore, the mechanism

by which the I4399M alters the fibrin network is still to be deduced.

4.7 Conclusions

This present study has demonstrated that the rs3798220 SNP in LPA, which

results in an Ile to Met at the 4399 residue of the protease domain, alters fibrin clot

architecture in the vasculature, which can account for the altered prothrombotic properties

of this variant. Lp(a) from carriers of the SNP and purified apo(a) containing the mutation

results in thicker fibrin fibers and aggregated nodules at the branching points of a fibrin

clot. Clots from individuals homozygous for the mutation with high Lp(a) levels are less

Page 87: Characterization of the I4399M Variant of Apolipoprotein(a

72

dense than clots formed from comparable individuals lacking the mutation, but are

comparable with respect to permeability. As well, clots corresponding to the Met variant

are significantly more turbid and possess significantly longer fibrinolysis times compared

to wild-type. This variant was also found to cause precipitation of soluble fibrinogen. We

did not observe an effect of the Met variant on the rate of secretion of 17K from cultured

hepatocytes, suggesting that this mechanism does not explain elevated levels of Lp(a)

containing this polymorphism in the human population. Upon analysis of the mutated

residue, it was found that Met in the I4399M variant has the capability to become

modified to a methionine sulfoxide residue. Overall, our data suggests that this

modification affects fibrinogen and the formation of the fibrin clot, which accounts for

the affect of the I4399M variant on coagulation and fibrinolysis. However, the exact

mechanism for these effects remains to be discovered. Further studies are needed to

understand the full mechanism by which the Met residue becomes oxidized, as well as

how this mutated residue interacts with fibrinogen to cause the altered fibrin structure that

we observed in this study.

Page 88: Characterization of the I4399M Variant of Apolipoprotein(a

73

References

1. Falk E. Pathogenesis of Atherosclerosis. Journal of the American College of

Cardiology. 2006;47(8, Supplement):C7-C12.

2. Hansson GK, Robertson AK, Soderberg-Naucler C. Inflammation and

atherosclerosis. Annual review of pathology. 2006;1:297-329.

3. Kragel AH, Reddy SG, Wittes JT, Roberts WC. Morphometric analysis of the

composition of atherosclerotic plaques in the four major epicardial coronary arteries in

acute myocardial infarction and in sudden coronary death. Circulation. 1989;80(6):1747-

56.

4. Lusis AJ. Atherosclerosis. Nature. 2000;407(6801):233-41.

5. World Health Organization: Cardiovascular Disease [Available from:

http://www.who.int/cardiovascular_diseases/en/.

6. Hansson GK. Inflammation, atherosclerosis, and coronary artery disease. The

New England journal of medicine. 2005;352(16):1685-95.

7. Abraham D, Distler O. How does endothelial cell injury start? The role of

endothelin in systemic sclerosis. Arthritis research & therapy. 2007;9 Suppl 2:S2.

8. Ley K, Miller YI, Hedrick CC. Monocyte and macrophage dynamics during

atherogenesis. Arteriosclerosis, thrombosis, and vascular biology. 2011;31(7):1506-16.

9. Shashkin P, Dragulev B, Ley K. Macrophage differentiation to foam cells. Current

pharmaceutical design. 2005;11(23):3061-72.

Page 89: Characterization of the I4399M Variant of Apolipoprotein(a

74

10. Dickhout JG, Basseri S, Austin RC. Macrophage function and its impact on

atherosclerotic lesion composition, progression, and stability: the good, the bad, and the

ugly. Arteriosclerosis, thrombosis, and vascular biology. 2008;28(8):1413-5.

11. Marx SO, Totary-Jain H, Marks AR. Vascular smooth muscle cell proliferation in

restenosis. Circulation Cardiovascular interventions. 2011;4(1):104-11.

12. Gomez D, Owens GK. Smooth muscle cell phenotypic switching in

atherosclerosis. Cardiovascular research. 2012;95(2):156-64.

13. Arroyo LH, Lee RT. Mechanisms of plaque rupture: mechanical and biologic

interactions. Cardiovascular research. 1999;41(2):369-75.

14. Neerman-Arbez M, de Moerloose P. Mutations in the fibrinogen gene cluster

accounting for congenital afibrinogenemia: an update and report of 10 novel mutations.

Human mutation. 2007;28(6):540-53.

15. Mosesson MW. Fibrinogen and fibrin structure and functions. Journal of

thrombosis and haemostasis : JTH. 2005;3(8):1894-904.

16. Ferry JD. The Mechanism of Polymerization of Fibrinogen. Proceedings of the

National Academy of Sciences of the United States of America. 1952;38(7):566-9.

17. Fowler WE, Hantgan RR, Hermans J, Erickson HP. Structure of the fibrin

protofibril. Proceedings of the National Academy of Sciences of the United States of

America. 1981;78(8):4872-6.

18. Fogelson AL, Keener JP. Toward an understanding of fibrin branching structure.

Physical review E, Statistical, nonlinear, and soft matter physics. 2010;81(5 Pt 1):051922.

Page 90: Characterization of the I4399M Variant of Apolipoprotein(a

75

19. Undas A, Ariens RA. Fibrin clot structure and function: a role in the

pathophysiology of arterial and venous thromboembolic diseases. Arteriosclerosis,

thrombosis, and vascular biology. 2011;31(12):e88-99.

20. Castellino FJ, Ploplis VA. Structure and function of the plasminogen/plasmin

system. Thrombosis and haemostasis. 2005;93(4):647-54.

21. Lamba D, Bauer M, Huber R, Fischer S, Rudolph R, Kohnert U, et al. The 2.3 A

crystal structure of the catalytic domain of recombinant two-chain human tissue-type

plasminogen activator. Journal of molecular biology. 1996;258(1):117-35.

22. Church FC, Pratt CW, Noyes CM, Kalayanamit T, Sherrill GB, Tobin RB, et al.

Structural and functional properties of human alpha-thrombin, phosphopyridoxylated

alpha-thrombin, and gamma T-thrombin. Identification of lysyl residues in alpha-

thrombin that are critical for heparin and fibrin(ogen) interactions. The Journal of

biological chemistry. 1989;264(31):18419-25.

23. Tsurupa G, Medved L. Fibrinogen alpha C domains contain cryptic plasminogen

and tPA binding sites. Annals of the New York Academy of Sciences. 2001;936:328-30.

24. Yakovlev S, Makogonenko E, Kurochkina N, Nieuwenhuizen W, Ingham K,

Medved L. Conversion of fibrinogen to fibrin: mechanism of exposure of tPA- and

plasminogen-binding sites. Biochemistry. 2000;39(51):15730-41.

25. Hoylaerts M, Rijken DC, Lijnen HR, Collen D. Kinetics of the activation of

plasminogen by human tissue plasminogen activator. Role of fibrin. The Journal of

biological chemistry. 1982;257(6):2912-9.

Page 91: Characterization of the I4399M Variant of Apolipoprotein(a

76

26. Hajjar KA, Harpel PC, Jaffe EA, Nachman RL. Binding of plasminogen to

cultured human endothelial cells. The Journal of biological chemistry.

1986;261(25):11656-62.

27. Castellino FJ, McCance SG. The kringle domains of human plasminogen. Ciba

Foundation symposium. 1997;212:46-60; discussion -5.

28. McCance SG, Menhart N, Castellino FJ. Amino acid residues of the kringle-4 and

kringle-5 domains of human plasminogen that stabilize their interactions with omega-

amino acid ligands. The Journal of biological chemistry. 1994;269(51):32405-10.

29. Angles-Cano E, Rojas G. Apolipoprotein(a): structure-function relationship at the

lysine-binding site and plasminogen activator cleavage site. Biological chemistry.

2002;383(1):93-9.

30. Tulinsky A, Park CH, Mao B, Llinas M. Lysine/fibrin binding sites of kringles

modeled after the structure of kringle 1 of prothrombin. Proteins. 1988;3(2):85-96.

31. Madison EL, Coombs GS, Corey DR. Substrate specificity of tissue type

plasminogen activator. Characterization of the fibrin independent specificity of t-PA for

plasminogen. The Journal of biological chemistry. 1995;270(13):7558-62.

32. Nesheim M. Thrombin and fibrinolysis. Chest. 2003;124(3 Suppl):33s-9s.

33. Cockell CS, Marshall JM, Dawson KM, Cederholm-Williams SA, Ponting CP.

Evidence that the conformation of unliganded human plasminogen is maintained via an

intramolecular interaction between the lysine-binding site of kringle 5 and the N-terminal

peptide. The Biochemical journal. 1998;333 ( Pt 1):99-105.

Page 92: Characterization of the I4399M Variant of Apolipoprotein(a

77

34. Horrevoets AJ, Pannekoek H, Nesheim ME. A steady-state template model that

describes the kinetics of fibrin-stimulated [Glu1]- and [Lys78]plasminogen activation by

native tissue-type plasminogen activator and variants that lack either the finger or kringle-

2 domain. The Journal of biological chemistry. 1997;272(4):2183-91.

35. Renatus M, Engh RA, Stubbs MT, Huber R, Fischer S, Kohnert U, et al. Lysine

156 promotes the anomalous proenzyme activity of tPA: X-ray crystal structure of single-

chain human tPA. The EMBO journal. 1997;16(16):4797-805.

36. de Vries C, Veerman H, Koornneef E, Pannekoek H. Tissue-type plasminogen

activator and its substrate Glu-plasminogen share common binding sites in limited

plasmin-digested fibrin. The Journal of biological chemistry. 1990;265(23):13547-52.

37. Henschen A, Lottspeich F, Kehl M, Southan C. Covalent structure of fibrinogen.

Annals of the New York Academy of Sciences. 1983;408:28-43.

38. Blomback B, Hessel B, Hogg D. Disulfide bridges in nh2 -terminal part of human

fibrinogen. Thrombosis research. 1976;8(5):639-58.

39. Huang S, Cao Z, Davie EW. The role of amino-terminal disulfide bonds in the

structure and assembly of human fibrinogen. Biochemical and biophysical research

communications. 1993;190(2):488-95.

40. Zhang JZ, Redman CM. Identification of B beta chain domains involved in human

fibrinogen assembly. The Journal of biological chemistry. 1992;267(30):21727-32.

41. Blomback B. Studies on fibrinogen: its purification and conversion into fibrin.

Acta Physiol Scand Suppl. 1958;43(148):1-51.

Page 93: Characterization of the I4399M Variant of Apolipoprotein(a

78

42. Blomback B, Hessel B, Hogg D, Therkildsen L. A two-step fibrinogen--fibrin

transition in blood coagulation. Nature. 1978;275(5680):501-5.

43. Scheraga HA. Thrombin and its interaction with fibrinogen. Annals of the New

York Academy of Sciences. 1958;75(1):189-94.

44. Everse SJ, Spraggon G, Veerapandian L, Riley M, Doolittle RF. Crystal structure

of fragment double-D from human fibrin with two different bound ligands. Biochemistry.

1998;37(24):8637-42.

45. Shimizu A, Nagel GM, Doolittle RF. Photoaffinity labeling of the primary fibrin

polymerization site: isolation and characterization of a labeled cyanogen bromide

fragment corresponding to gamma-chain residues 337-379. Proceedings of the National

Academy of Sciences of the United States of America. 1992;89(7):2888-92.

46. Krakow W, Endres GF, Siegel BM, Scheraga HA. An electron microscopic

investigation of the polymerization of bovine fibrin monomer. Journal of molecular

biology. 1972;71(1):95-103.

47. Hewat EA, Tranqui L, Wade RH. Electron microscope structural study of

modified fibrin and a related modified fibrinogen aggregate. Journal of molecular

biology. 1983;170(1):203-22.

48. Mosesson MW, Siebenlist KR, Amrani DL, DiOrio JP. Identification of

covalently linked trimeric and tetrameric D domains in crosslinked fibrin. Proceedings of

the National Academy of Sciences of the United States of America. 1989;86(4):1113-7.

49. Mosesson MW, DiOrio JP, Siebenlist KR, Wall JS, Hainfeld JF. Evidence for a

second type of fibril branch point in fibrin polymer networks, the trimolecular junction.

Blood. 1993;82(5):1517-21.

Page 94: Characterization of the I4399M Variant of Apolipoprotein(a

79

50. Mosesson MW, DiOrio JP, Muller MF, Shainoff JR, Siebenlist KR, Amrani DL,

et al. Studies on the ultrastructure of fibrin lacking fibrinopeptide B (beta-fibrin). Blood.

1987;69(4):1073-81.

51. Medved LV, Litvinovich SV, Ugarova TP, Lukinova NI, Kalikhevich VN,

Ardemasova ZA. Localization of a fibrin polymerization site complementary to Gly-His-

Arg sequence. FEBS letters. 1993;320(3):239-42.

52. Collet JP, Allali Y, Lesty C, Tanguy ML, Silvain J, Ankri A, et al. Altered fibrin

architecture is associated with hypofibrinolysis and premature coronary atherothrombosis.

Arteriosclerosis, thrombosis, and vascular biology. 2006;26(11):2567-73.

53. Lord ST. Molecular mechanisms affecting fibrin structure and stability.

Arteriosclerosis, thrombosis, and vascular biology. 2011;31(3):494-9.

54. Collet JP, Park D, Lesty C, Soria J, Soria C, Montalescot G, et al. Influence of

fibrin network conformation and fibrin fiber diameter on fibrinolysis speed: dynamic and

structural approaches by confocal microscopy. Arteriosclerosis, thrombosis, and vascular

biology. 2000;20(5):1354-61.

55. Mega JL, Morrow DA, de Lemos JA, Mohanavelu S, Cannon CP, Sabatine MS.

Thrombus Precursor Protein and Clinical Outcomes in Patients With Acute Coronary

Syndromes. Journal of the American College of Cardiology. 2008;51(25):2422-9.

56. K B. A new serum type system in man--the LP system. Acta Pathol Microbiol

Scand. 1963;59:369-82.

57. Koschinsky ML, Cote GP, Gabel B, van der Hoek YY. Identification of the

cysteine residue in apolipoprotein(a) that mediates extracellular coupling with

apolipoprotein B-100. The Journal of biological chemistry. 1993;268(26):19819-25.

Page 95: Characterization of the I4399M Variant of Apolipoprotein(a

80

58. Fless GM, ZumMallen ME, Scanu AM. Physicochemical properties of

apolipoprotein(a) and lipoprotein(a-) derived from the dissociation of human plasma

lipoprotein (a). The Journal of biological chemistry. 1986;261(19):8712-8.

59. Koschinsky ML. Novel insights into Lp(a) physiology and pathogenicity: more

questions than answers? Cardiovascular & hematological disorders drug targets.

2006;6(4):267-78.

60. McLean JW, Tomlinson JE, Kuang WJ, Eaton DL, Chen EY, Fless GM, et al.

cDNA sequence of human apolipoprotein(a) is homologous to plasminogen. Nature.

1987;330(6144):132-7.

61. Scanu AM, Edelstein C. Kringle-dependent structural and functional

polymorphism of apolipoprotein (a). Biochimica et biophysica acta. 1995;1256(1):1-12.

62. Phillips ML, Lembertas AV, Schumaker VN, Lawn RM, Shire SJ, Zioncheck TF.

Physical properties of recombinant apolipoprotein(a) and its association with LDL to

form an LP(a)-like complex. Biochemistry. 1993;32(14):3722-8.

63. van der Hoek YY, Wittekoek ME, Beisiegel U, Kastelein JJ, Koschinsky ML. The

apolipoprotein(a) kringle IV repeats which differ from the major repeat kringle are

present in variably-sized isoforms. Human molecular genetics. 1993;2(4):361-6.

64. Lackner C, Cohen JC, Hobbs HH. Molecular definition of the extreme size

polymorphism in apolipoprotein(a). Human molecular genetics. 1993;2(7):933-40.

65. Marcovina SM, Hobbs HH, Albers JJ. Relation between number of

apolipoprotein(a) kringle 4 repeats and mobility of isoforms in agarose gel: basis for a

standardized isoform nomenclature. Clinical chemistry. 1996;42(3):436-9.

Page 96: Characterization of the I4399M Variant of Apolipoprotein(a

81

66. Koschinsky ML, Beisiegel U, Henne-Bruns D, Eaton DL, Lawn RM.

Apolipoprotein(a) size heterogeneity is related to variable number of repeat sequences in

its mRNA. Biochemistry. 1990;29(3):640-4.

67. Becker L, McLeod RS, Marcovina SM, Yao Z, Koschinsky ML. Identification of

a critical lysine residue in apolipoprotein B-100 that mediates noncovalent interaction

with apolipoprotein(a). The Journal of biological chemistry. 2001;276(39):36155-62.

68. Gabel BR, McLeod RS, Yao Z, Koschinsky ML. Sequences within the amino

terminus of ApoB100 mediate its noncovalent association with apo(a). Arteriosclerosis,

thrombosis, and vascular biology. 1998;18(11):1738-44.

69. Rahman MN, Becker L, Petrounevitch V, Hill BC, Jia Z, Koschinsky ML.

Comparative analyses of the lysine binding site properties of apolipoprotein(a) kringle IV

types 7 and 10. Biochemistry. 2002;41(4):1149-55.

70. Becker L, Nesheim ME, Koschinsky ML. Catalysis of covalent Lp(a) assembly:

evidence for an extracellular enzyme activity that enhances disulfide bond formation.

Biochemistry. 2006;45(32):9919-28.

71. Hancock MA, Boffa MB, Marcovina SM, Nesheim ME, Koschinsky ML.

Inhibition of plasminogen activation by lipoprotein(a): critical domains in

apolipoprotein(a) and mechanism of inhibition on fibrin and degraded fibrin surfaces. The

Journal of biological chemistry. 2003;278(26):23260-9.

72. Harpel PC, Gordon BR, Parker TS. Plasmin catalyzes binding of lipoprotein (a) to

immobilized fibrinogen and fibrin. Proceedings of the National Academy of Sciences of

the United States of America. 1989;86(10):3847-51.

Page 97: Characterization of the I4399M Variant of Apolipoprotein(a

82

73. Gabel BR, Koschinsky MI. Analysis of the proteolytic activity of a recombinant

form of apolipoprotein(a). Biochemistry. 1995;34(48):15777-84.

74. Koschinsky ML, Marcovina SM. Structure-function relationships in

apolipoprotein(a): insights into lipoprotein(a) assembly and pathogenicity. Current

opinion in lipidology. 2004;15(2):167-74.

75. Laplaud PM, Beaubatie L, Rall SC, Jr., Luc G, Saboureau M. Lipoprotein[a] is the

major apoB-containing lipoprotein in the plasma of a hibernator, the hedgehog (Erinaceus

europaeus). Journal of lipid research. 1988;29(9):1157-70.

76. Lawn RM, Boonmark NW, Schwartz K, Lindahl GE, Wade DP, Byrne CD, et al.

The recurring evolution of lipoprotein(a). Insights from cloning of hedgehog

apolipoprotein(a). The Journal of biological chemistry. 1995;270(41):24004-9.

77. GeneCards: LPA Gene [Internet]. Gene Cards. Available from:

http://www.genecards.org/cgi-bin/carddisp.pl?gene=LPA&keywords=LPA,gene.

78. Gaw A, Boerwinkle E, Cohen JC, Hobbs HH. Comparative analysis of the apo(a)

gene, apo(a) glycoprotein, and plasma concentrations of Lp(a) in three ethnic groups.

Evidence for no common "null" allele at the apo(a) locus. The Journal of clinical

investigation. 1994;93(6):2526-34.

79. Cox LA, Jett C, Hixson JE. Molecular basis of an apolipoprotein[a] null allele: a

splice site mutation is associated with deletion of a single exon. Journal of lipid research.

1998;39(7):1319-26.

80. Rader DJ, Cain W, Ikewaki K, Talley G, Zech LA, Usher D, et al. The inverse

association of plasma lipoprotein(a) concentrations with apolipoprotein(a) isoform size is

Page 98: Characterization of the I4399M Variant of Apolipoprotein(a

83

not due to differences in Lp(a) catabolism but to differences in production rate. The

Journal of clinical investigation. 1994;93(6):2758-63.

81. Tsimikas S, Hall JL. Lipoprotein(a) as a potential causal genetic risk factor of

cardiovascular disease: a rationale for increased efforts to understand its pathophysiology

and develop targeted therapies. J Am Coll Cardiol. 2012;60(8):716-21.

82. Boerwinkle E, Leffert CC, Lin J, Lackner C, Chiesa G, Hobbs HH.

Apolipoprotein(a) gene accounts for greater than 90% of the variation in plasma

lipoprotein(a) concentrations. The Journal of clinical investigation. 1992;90(1):52-60.

83. Mooser V, Scheer D, Marcovina SM, Wang J, Guerra R, Cohen J, et al. The

Apo(a) gene is the major determinant of variation in plasma Lp(a) levels in African

Americans. American journal of human genetics. 1997;61(2):402-17.

84. Gavish D, Azrolan N, Breslow JL. Plasma Ip(a) concentration is inversely

correlated with the ratio of Kringle IV/Kringle V encoding domains in the apo(a) gene.

The Journal of clinical investigation. 1989;84(6):2021-7.

85. Utermann G, Kraft HG, Menzel HJ, Hopferwieser T, Seitz C. Genetics of the

quantitative Lp(a) lipoprotein trait. I. Relation of LP(a) glycoprotein phenotypes to Lp(a)

lipoprotein concentrations in plasma. Human genetics. 1988;78(1):41-6.

86. White AL, Guerra B, Lanford RE. Influence of allelic variation on

apolipoprotein(a) folding in the endoplasmic reticulum. The Journal of biological

chemistry. 1997;272(8):5048-55.

87. Longenecker JC, Klag MJ, Marcovina SM, Liu YM, Jaar BG, Powe NR, et al.

High lipoprotein(a) levels and small apolipoprotein(a) size prospectively predict

Page 99: Characterization of the I4399M Variant of Apolipoprotein(a

84

cardiovascular events in dialysis patients. Journal of the American Society of Nephrology

: JASN. 2005;16(6):1794-802.

88. Marcovina SM, Albers JJ, Wijsman E, Zhang Z, Chapman NH, Kennedy H.

Differences in Lp[a] concentrations and apo[a] polymorphs between black and white

Americans. Journal of lipid research. 1996;37(12):2569-85.

89. Kraft HG, Lingenhel A, Kochl S, Hoppichler F, Kronenberg F, Abe A, et al.

Apolipoprotein(a) kringle IV repeat number predicts risk for coronary heart disease.

Arteriosclerosis, thrombosis, and vascular biology. 1996;16(6):713-9.

90. Wild SH, Fortmann SP, Marcovina SM. A prospective case-control study of

lipoprotein(a) levels and apo(a) size and risk of coronary heart disease in Stanford Five-

City Project participants. Arteriosclerosis, thrombosis, and vascular biology.

1997;17(2):239-45.

91. Huby T, Afzal V, Doucet C, Lawn RM, Gong EL, Chapman MJ, et al. Regulation

of the expression of the apolipoprotein(a) gene: evidence for a regulatory role of the 5'

distal apolipoprotein(a) transcription control region enhancer in yeast artificial

chromosome transgenic mice. Arteriosclerosis, thrombosis, and vascular biology.

2003;23(9):1633-9.

92. Chennamsetty I, Claudel T, Kostner KM, Baghdasaryan A, Kratky D, Levak-

Frank S, et al. Farnesoid X receptor represses hepatic human APOA gene expression. The

Journal of clinical investigation. 2011;121(9):3724-34.

93. Marcovina SM, Koschinsky ML, Albers JJ, Skarlatos S. Report of the National

Heart, Lung, and Blood Institute Workshop on Lipoprotein(a) and Cardiovascular

Disease: recent advances and future directions. Clinical chemistry. 2003;49(11):1785-96.

Page 100: Characterization of the I4399M Variant of Apolipoprotein(a

85

94. Illingworth DR, Stein EA, Mitchel YB, Dujovne CA, Frost PH, Knopp RH, et al.

Comparative effects of lovastatin and niacin in primary hypercholesterolemia. A

prospective trial. Archives of internal medicine. 1994;154(14):1586-95.

95. Goldie C, Taylor AJ, Nguyen P, McCoy C, Zhao XQ, Preiss D. Niacin therapy

and the risk of new-onset diabetes: a meta-analysis of randomised controlled trials. Heart

(British Cardiac Society). 2016;102(3):198-203.

96. Goldberg RB, Bittner VA, Dunbar RL, Fleg JL, Grunberger G, Guyton JR, et al.

Effects of Extended-Release Niacin Added to Simvastatin/Ezetimibe on Glucose and

Insulin Values in AIM-HIGH. The American journal of medicine. 2016;129(7):753.e13-

22.

97. Raal FJ, Giugliano RP, Sabatine MS, Koren MJ, Langslet G, Bays H, et al.

Reduction in lipoprotein(a) with PCSK9 monoclonal antibody evolocumab (AMG 145): a

pooled analysis of more than 1,300 patients in 4 phase II trials. J Am Coll Cardiol.

2014;63(13):1278-88.

98. Romagnuolo R, Scipione CA, Boffa MB, Marcovina SM, Seidah NG, Koschinsky

ML. Lipoprotein(a) catabolism is regulated by proprotein convertase subtilisin/kexin type

9 through the low density lipoprotein receptor. The Journal of biological chemistry.

2015;290(18):11649-62.

99. Luke MM, Kane JP, Liu DM, Rowland CM, Shiffman D, Cassano J, et al. A

polymorphism in the protease-like domain of apolipoprotein(a) is associated with severe

coronary artery disease. Arteriosclerosis, thrombosis, and vascular biology.

2007;27(9):2030-6.

Page 101: Characterization of the I4399M Variant of Apolipoprotein(a

86

100. Arai K, Luke MM, Koschinsky ML, Miller ER, Pullinger CR, Witztum JL, et al.

The I4399M variant of apolipoprotein(a) is associated with increased oxidized

phospholipids on apolipoprotein B-100 particles. Atherosclerosis. 2010;209(2):498-503.

101. Lu W, Cheng Y-C, Chen K, Wang H, Gerhard GS, Still CD, et al. Evidence for

several independent genetic variants affecting lipoprotein (a) cholesterol levels. Human

molecular genetics. 2015;24(8):2390-400.

102. Zabaneh D, Kumari M, Sandhu M, Wareham N, Wainwright N, Papamarkou T, et

al. Meta analysis of candidate gene variants outside the LPA locus with Lp(a) plasma

levels in 14,500 participants of six White European cohorts. Atherosclerosis.

2011;217(2):447-51.

103. Khalifa M, Noureen A, Ertelthalner K, Bandegi AR, Delport R, Firdaus WJ, et al.

Lack of association of rs3798220 with small apolipoprotein(a) isoforms and high

lipoprotein(a) levels in East and Southeast Asians. Atherosclerosis. 2015;242(2):521-8.

104. Clarke R, Peden JF, Hopewell JC, Kyriakou T, Goel A, Heath SC, et al. Genetic

variants associated with Lp(a) lipoprotein level and coronary disease. The New England

journal of medicine. 2009;361(26):2518-28.

105. Kamstrup PR, Nordestgaard BG. Lipoprotein(a) concentrations, isoform size, and

risk of type 2 diabetes: a Mendelian randomisation study. The lancet Diabetes &

endocrinology. 2013;1(3):220-7.

106. Kronenberg F, Kronenberg MF, Kiechl S, Trenkwalder E, Santer P, Oberhollenzer

F, et al. Role of lipoprotein(a) and apolipoprotein(a) phenotype in atherogenesis:

prospective results from the Bruneck study. Circulation. 1999;100(11):1154-60.

Page 102: Characterization of the I4399M Variant of Apolipoprotein(a

87

107. Marcovina SM, Koschinsky ML. Lipoprotein(a) Concentration and

Apolipoprotein(a) Size. A Synergistic Role in Advanced Atherosclerosis?

1999;100(11):1151-3.

108. Nordestgaard BG, Chapman MJ, Ray K, Boren J, Andreotti F, Watts GF, et al.

Lipoprotein(a) as a cardiovascular risk factor: current status. European heart journal.

2010;31(23):2844-53.

109. Marcovina SM, Koschinsky ML. A critical evaluation of the role of Lp(a) in

cardiovascular disease: can Lp(a) be useful in risk assessment? Seminars in vascular

medicine. 2002;2(3):335-44.

110. Cheng SW, Ting AC, Wong J. Lipoprotein (a) and its relationship to risk factors

and severity of atherosclerotic peripheral vascular disease. European journal of vascular

and endovascular surgery : the official journal of the European Society for Vascular

Surgery. 1997;14(1):17-23.

111. Price JF, Lee AJ, Rumley A, Lowe GD, Fowkes FG. Lipoprotein (a) and

development of intermittent claudication and major cardiovascular events in men and

women: the Edinburgh Artery Study. Atherosclerosis. 2001;157(1):241-9.

112. Milionis HJ, Winder AF, Mikhailidis DP. Lipoprotein (a) and stroke. Journal of

clinical pathology. 2000;53(7):487-96.

113. Nowak-Gottl U, Strater R, Heinecke A, Junker R, Koch HG, Schuierer G, et al.

Lipoprotein (a) and genetic polymorphisms of clotting factor V, prothrombin, and

methylenetetrahydrofolate reductase are risk factors of spontaneous ischemic stroke in

childhood. Blood. 1999;94(11):3678-82.

Page 103: Characterization of the I4399M Variant of Apolipoprotein(a

88

114. von Depka M, Nowak-Gottl U, Eisert R, Dieterich C, Barthels M, Scharrer I, et al.

Increased lipoprotein (a) levels as an independent risk factor for venous

thromboembolism. Blood. 2000;96(10):3364-8.

115. Danik JS, Buring JE, Chasman DI, Zee RYL, Ridker PM, Glynn RJ.

Lipoprotein(a), Polymorphisms in the LPA gene and Incident Venous Thromboembolism

Among 21,483 Women. Journal of thrombosis and haemostasis : JTH. 2013;11(1):205-8.

116. Tsai AW, Cushman M, Rosamond WD, Heckbert SR, Polak JF, Folsom AR.

Cardiovascular risk factors and venous thromboembolism incidence: the longitudinal

investigation of thromboembolism etiology. Archives of internal medicine.

2002;162(10):1182-9.

117. Reblin T, Meyer N, Labeur C, Henne-Bruns D, Beisiegel U. Extraction of

lipoprotein(a), apo B, and apo E from fresh human arterial wall and atherosclerotic

plaques. Atherosclerosis. 1995;113(2):179-88.

118. Pepin JM, O'Neil JA, Hoff HF. Quantification of apo[a] and apoB in human

atherosclerotic lesions. Journal of lipid research. 1991;32(2):317-27.

119. Hajjar KA, Nachman RL. The role of lipoprotein(a) in atherogenesis and

thrombosis. Annual review of medicine. 1996;47:423-42.

120. Anuurad E, Boffa MB, Koschinsky ML, Berglund L. Lipoprotein(a): a unique risk

factor for cardiovascular disease. Clinics in laboratory medicine. 2006;26(4):751-72.

121. Rand ML, Sangrar W, Hancock MA, Taylor DM, Marcovina SM, Packham MA,

et al. Apolipoprotein(a) enhances platelet responses to the thrombin receptor-activating

peptide SFLLRN. Arteriosclerosis, thrombosis, and vascular biology. 1998;18(9):1393-9.

Page 104: Characterization of the I4399M Variant of Apolipoprotein(a

89

122. Biemond BJ, Friederich PW, Koschinsky ML, Levi M, Sangrar W, Xia J, et al.

Apolipoprotein(a) attenuates endogenous fibrinolysis in the rabbit jugular vein

thrombosis model in vivo. Circulation. 1997;96(5):1612-5.

123. Sangrar W, Bajzar L, Nesheim ME, Koschinsky ML. Antifibrinolytic effect of

recombinant apolipoprotein(a) in vitro is primarily due to attenuation of tPA-mediated

Glu-plasminogen activation. Biochemistry. 1995;34(15):5151-7.

124. Loscalzo J, Weinfeld M, Fless GM, Scanu AM. Lipoprotein(a), fibrin binding,

and plasminogen activation. Arteriosclerosis (Dallas, Tex). 1990;10(2):240-5.

125. van der Hoek YY, Sangrar W, Cote GP, Kastelein JJ, Koschinsky ML. Binding of

recombinant apolipoprotein(a) to extracellular matrix proteins. Arteriosclerosis and

thrombosis : a journal of vascular biology / American Heart Association.

1994;14(11):1792-8.

126. Cho T, Jung Y, Koschinsky ML. Apolipoprotein(a), through its strong lysine-

binding site in KIV(10'), mediates increased endothelial cell contraction and permeability

via a Rho/Rho kinase/MYPT1-dependent pathway. The Journal of biological chemistry.

2008;283(45):30503-12.

127. Dube JB, Boffa MB, Hegele RA, Koschinsky ML. Lipoprotein(a): more

interesting than ever after 50 years. Current opinion in lipidology. 2012;23(2):133-40.

128. Etingin OR, Hajjar DP, Hajjar KA, Harpel PC, Nachman RL. Lipoprotein (a)

regulates plasminogen activator inhibitor-1 expression in endothelial cells. A potential

mechanism in thrombogenesis. The Journal of biological chemistry. 1991;266(4):2459-

65.

Page 105: Characterization of the I4399M Variant of Apolipoprotein(a

90

129. Grainger DJ, Kemp PR, Liu AC, Lawn RM, Metcalfe JC. Activation of

transforming growth factor-beta is inhibited in transgenic apolipoprotein(a) mice. Nature.

1994;370(6489):460-2.

130. Grainger DJ, Kirschenlohr HL, Metcalfe JC, Weissberg PL, Wade DP, Lawn RM.

Proliferation of human smooth muscle cells promoted by lipoprotein(a). Science (New

York, NY). 1993;260(5114):1655-8.

131. Juhan-Vague I, Collen D. On the role of coagulation and fibrinolysis in

atherosclerosis. Annals of epidemiology. 1992;2(4):427-38.

132. Kullo IJ, Gau GT, Tajik AJ. Novel risk factors for atherosclerosis. Mayo Clinic

proceedings. 2000;75(4):369-80.

133. Rouy D, Koschinsky ML, Fleury V, Chapman J, Angles-Cano E.

Apolipoprotein(a) and plasminogen interactions with fibrin: a study with recombinant

apolipoprotein(a) and isolated plasminogen fragments. Biochemistry. 1992;31(27):6333-

9.

134. Romagnuolo R, Marcovina SM, Boffa MB, Koschinsky ML. Inhibition of

plasminogen activation by apo(a): role of carboxyl-terminal lysines and identification of

inhibitory domains in apo(a). Journal of lipid research. 2014;55(4):625-34.

135. Feric NT, Boffa MB, Johnston SM, Koschinsky ML. Apolipoprotein(a) inhibits

the conversion of Glu-plasminogen to Lys-plasminogen: a novel mechanism for

lipoprotein(a)-mediated inhibition of plasminogen activation. Journal of thrombosis and

haemostasis : JTH. 2008;6(12):2113-20.

Page 106: Characterization of the I4399M Variant of Apolipoprotein(a

91

136. Naruszewicze M, Giroux LM, Davignon J. Oxidative modification of Lp(a)

causes changes in the structure and biological properties of apo(a). Chemistry and physics

of lipids. 1994;67-68:167-74.

137. Nielsen LB. Atherogenecity of lipoprotein(a) and oxidized low density

lipoprotein: insight from in vivo studies of arterial wall influx, degradation and efflux.

Atherosclerosis. 1999;143(2):229-43.

138. Palinski W, Rosenfeld ME, Yla-Herttuala S, Gurtner GC, Socher SS, Butler SW,

et al. Low density lipoprotein undergoes oxidative modification in vivo. Proceedings of

the National Academy of Sciences of the United States of America. 1989;86(4):1372-6.

139. Beaudeux JL, Cesarini ML, Gardes-Albert M, Maclouf J, Merval R, Esposito B, et

al. Native and gamma radiolysis-oxidized lipoprotein(a) increase the adhesiveness of

rabbit aortic endothelium. Atherosclerosis. 1997;132(1):29-35.

140. Steinberg D. At last, direct evidence that lipoxygenases play a role in

atherogenesis. The Journal of clinical investigation. 1999;103(11):1487-8.

141. Ren S, Man RY, Angel A, Shen GX. Oxidative modification enhances

lipoprotein(a)-induced overproduction of plasminogen activator inhibitor-1 in cultured

vascular endothelial cells. Atherosclerosis. 1997;128(1):1-10.

142. Galle J, Bengen J, Schollmeyer P, Wanner C. Impairment of endothelium-

dependent dilation in rabbit renal arteries by oxidized lipoprotein(a). Role of oxygen-

derived radicals. Circulation. 1995;92(6):1582-9.

143. Zioncheck TF, Powell LM, Rice GC, Eaton DL, Lawn RM. Interaction of

recombinant apolipoprotein(a) and lipoprotein(a) with macrophages. The Journal of

clinical investigation. 1991;87(3):767-71.

Page 107: Characterization of the I4399M Variant of Apolipoprotein(a

92

144. Riches K, Franklin L, Maqbool A, Peckham M, Adams M, Bond J, et al.

Apolipoprotein(a) acts as a chemorepellent to human vascular smooth muscle cells via

integrin alphaVbeta3 and RhoA/ROCK-mediated mechanisms. The international journal

of biochemistry & cell biology. 2013;45(8):1776-83.

145. Ragab MS, Selvaraj P, Sgoutas DS. Oxidized lipoprotein (a) induces cell adhesion

molecule Mac-1 (CD 11b) and enhances adhesion of the monocytic cell line U937 to

cultured endothelial cells. Atherosclerosis. 1996;123(1-2):103-13.

146. Nielsen LB, Nordestgaard BG, Stender S, Niendorf A, Kjeldsen K. Transfer of

lipoprotein(a) and LDL into aortic intima in normal and in cholesterol-fed rabbits.

Arteriosclerosis, thrombosis, and vascular biology. 1995;15(9):1492-502.

147. Pillarisetti S, Paka L, Obunike JC, Berglund L, Goldberg IJ. Subendothelial

retention of lipoprotein (a). Evidence that reduced heparan sulfate promotes lipoprotein

binding to subendothelial matrix. The Journal of clinical investigation. 1997;100(4):867-

74.

148. Sattler W, Kostner GM, Waeg G, Esterbauer H. Oxidation of lipoprotein Lp(a). A

comparison with low-density lipoproteins. Biochimica et biophysica acta.

1991;1081(1):65-74.

149. Bergmark C, Dewan A, Orsoni A, Merki E, Miller ER, Shin MJ, et al. A novel

function of lipoprotein [a] as a preferential carrier of oxidized phospholipids in human

plasma. Journal of lipid research. 2008;49(10):2230-9.

150. Friedman P, Horkko S, Steinberg D, Witztum JL, Dennis EA. Correlation of

antiphospholipid antibody recognition with the structure of synthetic oxidized

Page 108: Characterization of the I4399M Variant of Apolipoprotein(a

93

phospholipids. Importance of Schiff base formation and aldol condensation. The Journal

of biological chemistry. 2002;277(9):7010-20.

151. van der Valk FM, Bekkering S, Kroon J, Yeang C, Van den Bossche J, van Buul

JD, et al. Oxidized Phospholipids on Lipoprotein(a) Elicit Arterial Wall Inflammation and

an Inflammatory Monocyte Response in Humans. Circulation. 2016;134(8):611-24.

152. Edelstein C, Pfaffinger D, Yang M, Hill JS, Scanu AM. Naturally occurring

human plasminogen, like genetically related apolipoprotein(a), contains oxidized

phosphatidylcholine adducts. Biochimica et biophysica acta. 2010;1801(7):738-45.

153. Scipione CA, Sayegh SE, Romagnuolo R, Tsimikas S, Marcovina SM, Boffa MB,

et al. Mechanistic insights into Lp(a)-induced IL-8 expression: a role for oxidized

phospholipid modification of apo(a). Journal of lipid research. 2015;56(12):2273-85.

154. Chasman DI, Shiffman D, Zee RYL, Louie JZ, Luke MM, Rowland CM, et al.

Polymorphism in the Apolipoprotein(a) Gene, Plasma Lipoprotein(a), Cardiovascular

Disease, and Low-dose Aspirin Therapy. Atherosclerosis. 2009;203(2):371-6.

155. Rowland CM, Pullinger CR, Luke MM, Shiffman D, Green L, Movsesyan I, et al.

Lipoprotein (a), LPA Ile4399Met, and Fibrin Clot Properties. Thrombosis

research.133(5):863-7.

156. Rowland CM, Pullinger CR, Luke MM, Shiffman D, Green L, Movsesyan I, et al.

Lipoprotein (a), LPA Ile4399Met, and fibrin clot properties. Thromb Res.

2014;133(5):863-7.

157. Romagnuolo R, Marcovina SM, Boffa MB, Koschinsky ML. Inhibition of

plasminogen activation by apo(a): role of carboxyl-terminal lysines and identification of

inhibitory domains in apo(a). The Journal of Lipid Research. 2014;55(4):625-34.

Page 109: Characterization of the I4399M Variant of Apolipoprotein(a

94

158. Sangrar W, Gabel BR, Boffa MB, Walker JB, Hancock MA, Marcovina SM, et al.

The solution phase interaction between apolipoprotein(a) and plasminogen inhibits the

binding of plasminogen to a plasmin-modified fibrinogen surface. Biochemistry.

1997;36(34):10353-63.

159. Shevchenko A, Tomas H, Havlis J, Olsen JV, Mann M. In-gel digestion for mass

spectrometric characterization of proteins and proteomes. Nat Protoc. 2006;1(6):2856-60.

160. Bonen DK, Nassir F, Hausman AM, Davidson NO. Inhibition of N-linked

glycosylation results in retention of intracellular apo[a] in hepatoma cells, although

nonglycosylated and immature forms of apolipoprotein[a] are competent to associate with

apolipoprotein B-100 in vitro. Journal of lipid research. 1998;39(8):1629-40.

161. Angles-Cano E. Overview on fibrinolysis: plasminogen activation pathways on

fibrin and cell surfaces. Chemistry and physics of lipids. 1994;67-68:353-62.

162. Edelberg JM, Pizzo SV. Lipoprotein (a) in the regulation of fibrinolysis. Journal

of atherosclerosis and thrombosis. 1995;2 Suppl 1:S5-7.

163. Levine RL, Moskovitz J, Stadtman ER. Oxidation of methionine in proteins: roles

in antioxidant defense and cellular regulation. IUBMB life. 2000;50(4-5):301-7.

164. Davis RA. Lipoprotein synthesis and secretion by cultured hepatocytes. Methods

in enzymology. 1986;129:272-83.

165. Malaguarnera M, Giugno I, Trovato BA, Panebianco MP, Siciliano R, Ruello P.

Lipoprotein(a) concentration in patients with chronic active hepatitis C before and after

interferon treatment. Clinical therapeutics. 1995;17(4):721-8.

Page 110: Characterization of the I4399M Variant of Apolipoprotein(a

95

166. White AL, Rainwater DL, Lanford RE. Intracellular maturation of

apolipoprotein[a] and assembly of lipoprotein[a] in primary baboon hepatocytes. Journal

of lipid research. 1993;34(3):509-17.

167. Dong W, Simeonova PP, Gallucci R, Matheson J, Fannin R, Montuschi P, et al.

Cytokine expression in hepatocytes: role of oxidant stress. Journal of interferon &

cytokine research : the official journal of the International Society for Interferon and

Cytokine Research. 1998;18(8):629-38.

168. Hajjar KA, Gavish D, Breslow JL, Nachman RL. Lipoprotein(a) modulation of

endothelial cell surface fibrinolysis and its potential role in atherosclerosis. Nature.

1989;339(6222):303-5.

169. Knapp JP, Herrmann W. In vitro inhibition of fibrinolysis by apolipoprotein(a)

and lipoprotein(a) is size- and concentration-dependent. Clinical chemistry and laboratory

medicine. 2004;42(9):1013-9.

170. Bucay I, O'Brien ET, 3rd, Wulfe SD, Superfine R, Wolberg AS, Falvo MR, et al.

Physical determinants of fibrinolysis in single fibrin fibers. PloS one.

2015;10(2):e0116350.

171. Carr ME, Jr., Gabriel DA, McDonagh J. Influence of Ca2+ on the structure of

reptilase-derived and thrombin-derived fibrin gels. The Biochemical journal.

1986;239(3):513-6.

172. Glover CJ, McIntire LV, Brown CH, 3rd, Natelson EA. Rheological properties of

fibrin clots. Effects of fibrinogen concentration, Factor XIII deficiency, and Factor XIII

inhibition. The Journal of laboratory and clinical medicine. 1975;86(4):644-56.

Page 111: Characterization of the I4399M Variant of Apolipoprotein(a

96

173. Nair CH, Shah GA, Dhall DP. Effect of temperature, pH and ionic strength and

composition on fibrin network structure and its development. Thrombosis research.

1986;42(6):809-16.

174. Carr ME, Jr., Shen LL, Hermans J. Mass-length ratio of fibrin fibers from gel

permeation and light scattering. Biopolymers. 1977;16(1):1-15.

175. Undas A, Ciesla-Dul M, Drazkiewicz T, Sadowski J. Altered fibrin clot properties

are associated with residual vein obstruction: effects of lipoprotein(a) and

apolipoprotein(a) isoform. Thrombosis research. 2012;130(3):e184-7.

176. Undas A, Stepien E, Tracz W, Szczeklik A. Lipoprotein(a) as a modifier of fibrin

clot permeability and susceptibility to lysis. Journal of thrombosis and haemostasis : JTH.

2006;4(5):973-5.

177. Millar HR, Simpson JG, Stalker AL. An evaluation of the heat precipitation

method for plasma fibrinogen estimation. Journal of clinical pathology. 1971;24(9):827-

30.

178. Carr ME, Gabriel DA, McDonagh J. Influence of Ca2+ on the structure of

reptilase-derived and thrombin-derived fibrin gels. Biochemical Journal.

1986;239(3):513-6.

179. Fischer ER, Hansen BT, Nair V, Hoyt FH, Dorward DW. Scanning Electron

Microscopy. Current Protocols in Microbiology. 2012;CHAPTER:Unit2B.-UnitB.

180. Gu SX, Stevens JW, Lentz SR. Regulation of thrombosis and vascular function by

protein methionine oxidation. Blood. 2015;125(25):3851-9.

Page 112: Characterization of the I4399M Variant of Apolipoprotein(a

97

181. Martinez M, Weisel JW, Ischiropoulos H. Functional impact of oxidative post-

translational modifications on fibrinogen and fibrin clots. Free radical biology &

medicine. 2013;65:10.1016/j.freeradbiomed.2013.06.039.

Page 113: Characterization of the I4399M Variant of Apolipoprotein(a

98

Vita Auctoris

Name: Jackson McAiney

Place of Birth: Windsor, ON, Canada

Year of Birth: 1992

Education: University of Windsor, Windsor, ON, Canada

B. Sc. Honours Biochemistry and Biotechnology

2010-2014

University of Windsor, Windsor, ON, Canada

M. Sc. Chemistry and Biochemistry

2014-2016

Conferences: Southern Ontario Undergraduate Student Chemistry Conference

March 2014

Authors: Jackson McAiney1, Corey Scipione1, Dr. James Gauld1,

Marlys L. Koschinksy1,2, Michael B. Boffa1,2.

(1University of Windsor, 2Robarts Research Institute)

Title: Functional Characterization of the I4399M mutant of

apolipoprotein(a)

(Oral Presentation)

AHA/ATVB Conference 2016

May 2016

Authors: Jackson McAiney1 and Corey Scipione1, Janeen Auld1,

Dan Simard1, Dr. James Gauld1, Robert A. Hegele2, Marlys L.

Koschinksy1,2,

Michael B. Boffa1,2.

(1University of Windsor, 2Robarts Research Institute)

Title: Characterization of the I4399M variant of apolipoprotein(a):

implications for altered prothrombotic properties of lipoprotein(a)

(Poster Presentation)