peptide-modified polymer for endothelialization

96
Peptide-Modified Polymer For Endothelialization A Thesis Submitted to the Faculty Of Drexel University By Jamie Lyn Ostroha In partial fulfillment of the degree Of Masters of Science In Materials Engineering December 2001

Upload: others

Post on 28-Nov-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Peptide-modified polymer for endothelialization

Peptide-Modified Polymer For Endothelialization

A Thesis

Submitted to the Faculty

Of

Drexel University

By

Jamie Lyn Ostroha

In partial fulfillment of the degree

Of

Masters of Science

In

Materials Engineering

December 2001

Page 2: Peptide-modified polymer for endothelialization

ii

Dedications I would like to dedicate this book to my husband and my parents for their support and encouragement.

Page 3: Peptide-modified polymer for endothelialization

iii

Acknowledgements I would like to thank my advisor, Dr. T. Twardowski, for providing me with the opportunity to work on this project. It has been an interesting and enjoyable project. I would like to thank John Kemnitzer and Integra LifeSciences Corporation for the donation of polymer. Special thanks are due to Dr. James San Antonio from Thomas Jefferson University for providing the materials and lab space necessary for most of my work and for valuable discussions on the direction of this project. Special thanks to Dr. Markus Germann, also of Thomas Jefferson University, for the valuable NMR education and providing resources to run NMR. This work was funded in part by a joint Thomas Jefferson University and Drexel University grant.

Page 4: Peptide-modified polymer for endothelialization

iv

Table of Contents

List of Tables .................................................................................................................... vii

List of Figures .................................................................................................................. viii

Abstract ............................................................................................................................... x

1 Introduction................................................................................................................. 1

2 Background ................................................................................................................. 3

2.1 Cardiovascular Diseases ..................................................................................... 3

2.1.1 Atherosclerosis................................................................................................ 3

2.1.2 Stenosis ........................................................................................................... 4

2.1.3 Thrombosis ..................................................................................................... 4

2.2 Biology................................................................................................................ 5

2.2.1 Artery and Vein Biology................................................................................. 5

2.2.2 Luminal Endothelial Cell Biology.................................................................. 6

2.2.2.1 Endothelial Cell Surface ......................................................................... 7

2.2.2.2 Endothelial Cell Injury............................................................................ 7

2.3 Current Treatments of Cardiovascular Disease .................................................. 8

2.3.1 Non-Surgical Therapy..................................................................................... 8

2.3.2 Minimally Invasive Surgical Methods............................................................ 9

2.3.2.1 Angioplasty ............................................................................................. 9

2.3.2.2 Stents..................................................................................................... 10

2.3.3 Invasive Surgical Methods............................................................................ 11

2.3.3.1 Grafting ................................................................................................. 11

2.3.3.2 Synthetic Grafts .................................................................................... 13

Page 5: Peptide-modified polymer for endothelialization

v

2.3.3.3 Biodegradable Synthetic Grafts ............................................................ 14

2.3.3.4 Alternative Materials ............................................................................ 15

2.4 Polymer Modification ....................................................................................... 16

2.4.1 RGD Containing Peptides............................................................................. 17

2.4.2 Endothelial Seeding ...................................................................................... 18

3 Design Concept ......................................................................................................... 19

3.1 Material Selection ............................................................................................. 19

3.2 Heparin-Binding Peptides................................................................................. 20

4 Experimental ............................................................................................................. 23

4.1 Polymer Derivatization ..................................................................................... 23

4.2 NMR Spectroscopy........................................................................................... 25

4.3 Cell Culture....................................................................................................... 26

4.4 EC Attachment.................................................................................................. 26

4.5 EC Growth ........................................................................................................ 28

5 Results and Discussion ............................................................................................. 30

5.1 NMR ................................................................................................................. 30

5.2 Cell Attachment ................................................................................................ 36

5.3 Cell Growth....................................................................................................... 38

6 Conclusions............................................................................................................... 40

6.1 Recommendations: Characterization ................................................................ 41

6.1.1 Peptide Attachment....................................................................................... 41

6.1.2 Peptide Structure........................................................................................... 42

6.1.3 Cell Affinity .................................................................................................. 42

Page 6: Peptide-modified polymer for endothelialization

vi

6.1.4 Polymer Selection ......................................................................................... 42

6.2 Recommendations: Fabrication ........................................................................ 43

6.2.1 Processing ..................................................................................................... 43

6.2.2 Sterilization................................................................................................... 44

6.2.3 Storage .......................................................................................................... 44

List of References ............................................................................................................. 46

Appendix A- Tables.......................................................................................................... 58

Appendix B- Figures......................................................................................................... 66

Appendix C- Nomenclature .............................................................................................. 82

Appendix D- Abbreviations.............................................................................................. 84

Page 7: Peptide-modified polymer for endothelialization

vii

List of Tables 1. Patency rates for ASV and IMA grafting of the coronary artery of varying

diameters, from [142].................................................................................................. 58

2. Patency data for representative vascular grafts, adapted from Ku [51]. ..................... 59

3. Improving hemocompatibility of artificial biomaterials [73]. .................................... 60

4. Properties of native tissue and several synthetic polymers......................................... 61

5. NMR peak integration values for p(DTEC-c-X%DTC), X=0, 13, and 35, peaks...... 62

6. Integration values corresponding to polyalanine peaks for attachment to p(DTEC-c-X%DTC), X=0, 23, 50............................................................................................. 63

7. The corresponding attachment percentages have been calculated.............................. 63

8. Cell attachment strength as assessed using a centrifugation assay. Note that the poly(DTE carbonate) shows no significant improvement in attachment over the negative control, but attachment increases with additional functional groups. .......... 64

9. Comparative cell growth data at 7 days for various experimental and common vascular implant materials. ......................................................................................... 65

Page 8: Peptide-modified polymer for endothelialization

viii

List of Figures

1. Cross-Section of a typical artery and vein [143]. ........................................................66

2. a) Molecular model of poly(DTEC) repeat unit showing bond angles and bond lengths. b) Repeat unit of poly(DTE-co-X%DT carbonate). When R = CH2CH3 the repeat unit is desaminotyrosil-tyrosine-ethyl ester (DTE); when R = H the free unit is desaminotyrosil-tyrosine (DT). X is the molar fraction of repeat units with pendant acid groups. (Note: H’s are labeled α, β, γ, δ, ε, for identification and quantification by NMR spectra. The hydrogens in the pendant R groups are labeled by proximity to the ester group. Thus, the hydrogen on the free acid is R1, while the hydrogens in the ethyl ester are labeled R2 and R3.) c) Repeat unit of poly(DTEC-co-X%DTC) and L-lysine ethyl ester. Arrows indicate both possible attachment locations.....................................................................................................67

3. 2D TOCSY of poly(DTEC) in CDCl3. Peaks are labeled as detailed in Figure 2. Solid lines indicate three different correlation systems with interactions between neighboring hydrogen. .................................................................................................68

4. 1D NMR spectra of a) P(DTEC-c-0%DTC) and b)P(DTEC-c-13%DTC) with peak integration values. The peak labels correspond to hydrogen as marked in Figure 2. .......................................................................................................................69

5. 1-D and 2-D COSY NMR spectrum of L-lysine ethyl ester. Solid lines indicate correlation systems. Below is the corresponding molecule. ........................................70

6. TOCSY spectrum of P(DTEC-c-13%DTC) with attached L-lysine ethyl ester. The original correlation systems from the poly(DTE-co-X%DT carbonate) system are indicated by lines. The solid lines indicate correlation systems from the L-lysine. Note that the system has ( ) ( ) two related correlation systems, corresponding to α– and ε–amine attachment of the lysine.........................................71

7. NMR spectrum of P(DTEC-c-13%DTC) with attached L-lysine ethyl ester. Integration values are below peaks. The ratio of the α poly(DTE carbonate)peak, which should be constant at a concentration of 1 across chemistries, to the lysine spectral group at ca. 4.38 shows an approximately 11.5±0.5% lysine attachment......72

8. TOCSY spectrum of p(DTEC-c-13%DTC) with attached L-lysine ethyl ester and polylysine. Note that there are no additional peaks, only intensification of existing peaks. ...........................................................................................................................73

9. a) 1D NMR spectrum of p(DTEC-c-50%DTC). Peak integration values are shown below select peaks. The integration value for the peak at 4.7 ppm corresponds to the α peak in the polymer. Characteristic polyalanine peaks are also indicated. b) Structure of polyalanine...............................................................................................74

Page 9: Peptide-modified polymer for endothelialization

ix

10. 1D NMR spectrum of p(DTEC-c-35%DTC) with peptide (ARKKAAKA)4 attached via a lysine bridge..........................................................................................75

11. Attachment data from Table 8 for functional polymers, p(DTEC-c-X%TDC) X=0, 13, 23, 35, and 50% and positive (FN) and negative (MeCl2) controls. Attachment increases with additional functional groups.................................................................76

12. Attachment data for functional polymers and positive and negative controls. P(DTEC-c-X%TDC) X=0, 13, 23, 35, and 50% are shown at both the low (300 g) and medium (600 g) force. FN (positive control) and Blank (negative control) are drawn as straight lines as they are not subject to the same functionalization. The lines for the FN and the Blank indicate the cell attachment at the specified force......77

13. Cell attachment data for p(DTEC-c-13%DTC) alone (13%), with attached lysine (13wL), attached polylysine (PolyL), and attached heparin-binding 32mer peptide (32mer PEP).................................................................................................................78

14. Optical microscope images of cell attachment and growth on 13% poly(DTE-co-X%DT carbonate) at a) 1 hour, b) 1 day, c) 3 days and d) 7 days. Note that at 1 hour, the cells show flattening indicative of attachment. The total number of cells increases through day 7. On day 7, there are regions of cell confluence.....................79

15. Optical microscope image of cell attachment at day 5 on a) poly(DTE carbonate), b) 13% poly(DTE-co-X%DT carbonate), c) 35% poly(DTE-co-X%DT carbonate), d) PLA, e) PET, f) PVC and g) glass. .......................................................80

16. Optical microscope image of stained cells at the end of the growth assay, a) 12X shows areas of semi-confluence b) 23X shows spread cells indicating attachment c) 46X shows spread cells with visible nuclei and cells in the process of dividing. ...81

Page 10: Peptide-modified polymer for endothelialization

x

Abstract

Peptide-Modified Polymer For Endothelialization Jamie Lyn Ostroha

Thomas Twardowski

A functionalized polycarbonate material derived from tyrosine containing varying

levels of free acid functionality along the backbone, poly(DTE-co-X%DT carbonate), has

been functionalized with a peptide to promote endothelial cell attachment and evaluated

as a potential synthetic vascular graft material. Lysine derivatives have been produced to

promote the attachment of engineered peptides for promoting endothelialization.

NMR has been used to analyze the initial structure of the poly(DTEC-co-

X%DTC) and changes in the structure with the attachment of L-lysine ethyl ester and

polylysine, polyalanine, and the (ARKKAAKA)4 peptide. The polymer peak at 4.8 ppm

has been labeled the α-peak. This peak is the primary peak used to evaluate changes in

the spectra related to changing polymer functionality and the attachment of molecules to

the pendant acid group. Changes in the β-peak (3.05 ppm), the R2 peak (4.1 ppm) and the

R3 peak (1.3 ppm) are compared to the α-peak for quantification purposes. Key NMR

peaks have been identified and lysine attachment and quantification is discussed. A lysine

peak has been identified at 4.38 ppm. This peak is also compared to the polymer α-peak,

indicating 85-90% attachment efficiency. The use of polylysine, polyalanine, and

peptides establish that peptide binding is successful; qualitative and semi-quantitative

results are presented. Semi-quantitative results with polyalanine indicate a large degree of

attachment, although some interference with the spectral peaks has been noted.

Page 11: Peptide-modified polymer for endothelialization

xi

The effect of poly(DTEC-co-X%DTC) on human endothelial cell attachment and

growth is also studied and compared to current synthetic materials. Cell attachment has

been shown to increase significantly as the polymer becomes more hydrophilic. Non-

functionalized polymers show that less than five percent of cells remained attach at a

force of 300 g, while 50% functionalized polymers show 40% attachment at the same

force. Cell growth studies show an increase in cell affinity with increasing hydophilicity.

35% functionalized polymer shows over double the number of cells as the non-

functionalized material after seven days. The results for the p(DTEC) at increasing levels

of functionality bracket Dacron and PLA depending. Cell spreading, confluence, and

division have been confirmed by visual analysis.

Page 12: Peptide-modified polymer for endothelialization

xii

Page 13: Peptide-modified polymer for endothelialization

1

1 Introduction

Mundth et al predicted in 1978 [1] that there would be an increasing number of

cardiovascular problems due to the American lifestyle. In the late 1970s, deaths from

cardiovascular disease were around 750,000 a year. In 1998, cardiovascular diseases

claimed close to one million lives [2]. Though an increase in deaths from cardiovascular

disease has been observed, it has not increased as rapidly as predicted. This is mostly due

to rapid and continual advances in medicine and technology.

In 1949 the vascular graft was introduced [3]. The 1950s saw a great increase of

resources, both public and private, devoted to cardiovascular education and research [4].

The availability of synthetic grafts made it possible to prevent death from rupture of

arteriosclerotic aneurysms, to prevent loss of limbs in many patients with occlusive

diseases, and to improve the quality of life in others [5]. However, even though surgical

cardiovascular treatment has been in practice since the 1940s [4], there is an ongoing

need to develop improved vascular repair materials. According to 1998 estimates,

60,800,000 Americans have one or more forms of cardiovascular disease (CVD) [2], with

55,000 lower limb vascular reconstructions being necessary each year [6]. The current

materials and techniques used for cardiovascular repairs exhibit some problems, e.g.

inadequate long-term patency, especially in diseased blood vessels of the lower limbs [6-

9].

The focus of this research is the development of a material, modified to increase

patency and decrease the need for surgical revision. Native vessels have a structure that

resists thrombosis, intimal hyperplasia, and other causes of blockage. This is primarily

due to the inner lining of endothelial cells on the vessels. However, disease or surgery

Page 14: Peptide-modified polymer for endothelialization

2

often impairs these cells, increasing the risk of blockage. A functionalized, biodegradable

polymeric material has been developed with the potential to enhance endothelial cell

attachment.

Page 15: Peptide-modified polymer for endothelialization

3

2 Background

There are a variety of diseases that can disrupt blood flow and result in the need for

surgery. CVD, the biology of vasculature and some current treatments and synthetic

materials for vascular prostheses are reviewed.

2.1 Cardiovascular Diseases

Vascular diseases often result in a need to open, replace or bypass diseased or

damaged blood vessel segments. The primary CVD is atherosclerosis. Atherosclerotic

vascular disease, particularly coronary heart disease, is a leading cause of death

worldwide [10], and accounts for nearly three-quarters of all deaths from CVD.

2.1.1 Atherosclerosis

Atherosclerosis is a disease of the arterial wall. This disease often begins with

intima (inner lining) thickening, especially in locations corresponding to high shear

stress. Over time, lesions (scars) form and grow until the vessel becomes occluded [11].

It is caused primarily by environmental factors, although individuals may have an

increased or reduced susceptibility that is genetically predetermined. Such environmental

factors include a high-fat diet, cigarette smoking, and hypertension [2, 10, 12]. CVD

involves deposits of fatty substances, cholesterol, cellular waste products, calcium or

other substances on the inner lining of an artery. This build-up is called plaque, and

usually affects large and medium-sized arteries [2]. It has been shown that injured

endothelial cells lead to atherosclerosic lesions, which in turn lead to occluded vessels in

Page 16: Peptide-modified polymer for endothelialization

4

a three-step process: (1) smooth muscle cells accumulate, (2) the smooth muscle cells

form a connective tissue matrix, and (3) lipids accumulate [13]. This process occurs

because the damaged endothelial cells can no longer carry out the multiple, complex

interactions involving blood and vessel wall components that maintain vascular

homeostasis, in which healthy endothelial cells are a key component [14]. When these

cells are damaged or removed, homeostasis can longer be maintained and cardiovascular

disease results.

2.1.2 Stenosis

The simplest local disturbance is a restricted lumina, known as stenosis [12]. The

artery may become clogged by the buildup of fat, cholesterol or other substances over

time [2]. This produces highly increased shear stresses that are capable of removing

endothelial cells from the luminal lining [12].

2.1.3 Thrombosis

Plaques that rupture form blood clots (thrombus) that can block blood flow or

break off and travel to another part of the body (embolus). If either of these events occurs

and blocks a blood vessel that feeds the heart, it causes a “heart attack”. If it blocks a

blood vessel that feeds the brain, it causes a “stroke”. If the blood supply to the arms or

legs is reduced, it can cause difficulty in proper functioning and, in extreme cases,

gangrene [2].

Page 17: Peptide-modified polymer for endothelialization

5

2.2 Biology 2.2.1 Artery and Vein Biology

The vascular system is divided into two parts, the arteries and the veins. The

arteries consist of relatively thick-walled, viscoelastic tubes that undergo high pressure as

they carry blood away from the heart. On the other hand, veins are thinner, but have

larger diameter, elastic tubes that experience low-pressure conditions as they carry blood

toward the heart [15]. The arteries originate from the aorta and its branches, becoming

smaller and smaller as they branch out. As the cross-sectional area of the vessel

decreases, the blood velocity increases to maintain a constant total blood volume. Veins

collect blood from the capillaries and successively join together to form progressively

larger veins. They then return blood to the heart. Veins can accommodate much larger

volumes of blood with very slight changes in pressure [4]. Except for the differences in

thickness, the walls of the largest arteries and veins consist of the same three distinct,

well-defined layers (Figure 1). The innermost layer is the thinnest, tunica intima, a

continuous and well-developed lining consisting of a single layer of simple, flat

endothelial cells that are held together by VE cadherin [16]. The endothelial cells are

surrounded by a thin layer of subendothelial connective tissue interconnected with a

number of circularly arranged elastic fibers to form the subendothelium, and are

separated from the next adjacent wall layer by a thick elastic band called the internal

elastic lamina. The middle layer, tunica media, is the thickest and is composed of many

elastic fibers, a significant number of smooth muscle cells, some interlacing collagenous

tissue, and an intercellular mucopolysaccharide substance. These are separated from the

Page 18: Peptide-modified polymer for endothelialization

6

next adjacent layer by another thick elastic band called the external elastic lamina. The

medium-sized tunica adventitia is the outer vascular sheath consisting entirely of

connective tissue [11, 15, 17]. Veins and arteries are traditionally divided into three

categories based on size: large (>6mm), medium (4-6mm) and small diameter vessels

(<4mm) [3, 18, 19]. The difference in the vessel size alters the flow characteristics of the

blood, which can in turn alter the vessel-blood interface. There is an increased surface to

volume ratio and a reduced blood flow volume in smaller-diameter vessels. This causes

an increase in the blood-surface contact time, which can result in increased activation of

surrounding blood elements if the endothelium becomes damaged. This may explain why

patency rates are significantly lower in medium and especially smaller synthetic vessels.

2.2.2 Luminal Endothelial Cell Biology

The endothelium is generally a single layer of tightly packed cells that line the

vascular lumen [20] to create the intima. The cells of the lining are dynamic partners in

multiple, complex interactions involving macromolecules, cellular blood components,

and vessel wall components, such as smooth muscle cells and extracellular matrix [21].

The luminal endothelium senses, transmits, and participates in the adjustments that are

necessary as a result of deviations in homeostasis [22]. Deviations in homeostasis can

include blood loss, reduced or elevated blood pressure, and constricted or dilated vessels.

Arterial endothelial cells form a continous, nonthrombogenic, metabolically active lining

for the vascular system [23]. One key function performed by the endothelial cells is

heparan sulfate synthesis [24, 25], which is known to have antithrombogenic and

anticoagulant properties [26, 27]. This coagulation/anticoagulation system is very

Page 19: Peptide-modified polymer for endothelialization

7

important in the homeostasis of the circulatory system, promoting fluidity within and

rapid clot formation outside of the blood vessels. During homeostasis, anticoagulant

mechanisms dominate, while in a disturbed state coagulation mechanisms dominate [28].

So, any change in the normal structure and function of the endothelial lining will cause

coagulation. For example, following denudation injury, where endothelial cells are

stripped locally from a small path of vascular tissue, the cells in and around the wound

area are stimulated to spread, migrate, and proliferate to reconstitute a continuous lining

[23]. If the cells are unable to re-establish a continuous lining, platelet attachment, release

of platelet growth factor, and smooth muscle cell migration and proliferation are

stimulated instead [29]. The migration and adhesion of platelets at the site of vascular

injury depends on a combined interaction between the platelet cell surface receptors and

the adhesive matrix proteins on the exposed subendothelium [30, 31].

2.2.2.1 Endothelial Cell Surface

The complex interaction of endothelial cells and blood components is a result of

the complex surface composition of the endothelial cell. The endothelial cell surface is

made up of proteins, glycoproteins, glycolipids, and proteoglycans (PG). The prevalence

of anionic sites, predominately from the proteoglycans, gives the luminal surface a net

negative charge [22].

2.2.2.2 Endothelial Cell Injury

Endothelial cell injury can result from many causes, including hemodynamic

stress, mechanical trauma, hypercholesterolemia, infectious agents, oxygen, and chemical

Page 20: Peptide-modified polymer for endothelialization

8

agents like homocysteine [32]. These forces play an important role in the occurrence of

cardiovascular disease [33] by causing what is known as endothelial cell dysfunction.

The best method for prevention of CVD is primarily to prevent damage to the endothelial

layer or secondarily to repair a damaged endothelial layer [23]. This has been an area of

great interest since the various functions of the endothelial layer have become more

clearly understood.

2.3 Current Treatments of Cardiovascular Disease

Treatments of CVDs include medical therapy with specific medication; minimally

invasive surgical treatments such as balloon angioplasty, laser angioplasty, atherectomy,

and stents; and invasive surgical methods, which includes the implantation of vascular

grafts.

2.3.1 Non-Surgical Therapy

There are several risk factors for vascular disease, as mentioned earlier, the most

notorious being high cholesterol levels. More specifically, elevated serum levels of total

cholesterol (TC) and low-density lipopotein cholesterol (LDL-C) lead to CVD [34, 35].

Dietary intervention and drug therapy have been shown to lower LDL-C levels. Another

non-surgical procedure in current practice is the use of LDL adsorption compounds [36].

Synthetic and natural anti-oxidants [37], like vitamin E and gene therapy [38], have been

examined for the prevention of atherosclerosis and have been shown to reduce the

Page 21: Peptide-modified polymer for endothelialization

9

progression of this disease [39, 40]. Generally, however, these techniques are not

sufficient to treat acute or developing CVD.

2.3.2 Minimally Invasive Surgical Methods 2.3.2.1 Angioplasty

Angioplasty is a minimally invasive procedure and is now the customary

procedure for treating coronary artery disease: the common technique being percutaneous

transluminal coronary angioplasty (PTCA) [41]. During angioplasty, a balloon-tipped or

laser-tipped catheter is introduced into a diseased blood vessel, usually from a remote site

like the femoral artery. In balloon angioplasty, a balloon is inflated at the site of closure

and the vessel opens further, allowing for improved flow of blood [16].

Because of the complex nature of vascular tissue, angioplasty can result in several

complications. First, balloon dilation can cause a fracture of the atherosclerotic plaque.

Second, endothelial damage can occur that in turn can stimulate platelet adhesion.

Finally, the traumatic dilation can cause damage to the media, ranging from stretching to

tearing [41]. Angioplasty often results in early restenosis, due to shrinkage of the

muscular components of the arterial wall [41-43]. However, stenting after angioplasty has

been shown to reduce the likelihood of restenosis [41, 44]. The stent prevents shrinkage

of the vessel and maintains vessel patency mechanically. In 1998, 926,000 angioplasties

were performed in the United States. Of these, 539,000 were PTCAs [2].

Similar to angioplasty, atherectomy is a procedure for opening coronary arteries

blocked by plaque. Coronary atherectomy uses a laser catheter to vaporize the plaque, or

Page 22: Peptide-modified polymer for endothelialization

10

a small rotating “shaver” to shave off the plaque. The catheter is inserted into the body

the same way as in angioplasty. Balloon angioplasty or stenting may then be used after

the atherectomy [2].

A primary disadvantage of these techniques is a limited capability in small vessel

applications and a high rate of restenosis.

2.3.2.2 Stents

Stents date back to 1912 when Carrel used a glass tube to open occluded canine

arteries [45]. Vascular stents were introduced into human vascular surgery in the 1960s as

a way of maintaining the patency of angioplasties [46]. During stenting a wire mesh stent

is used to support an artery that has recently been cleared by angioplasty. The stent is

collapsed to a small diameter, placed over an angioplasty balloon catheter and moved into

the blocked area. When the balloon is inflated, the stent expands, locks in place and

forms a scaffold to hold the artery open. The stent remains in the artery permanently to

hold it open, which improves blood flow to the heart muscle [2]. Stents can also be used

for opening occluded blood vessels and spanning small dissections [44].

The use of vascular stents to combat cardiovascular disease has, in many cases,

reduced the need for large incisions and long hospital stays that generally occur with

major vascular surgery. Today vascular stents are still being used to maintain patency

following angioplasty of occluded arteries and veins. The stent procedure is fairly

common, now representing 70-90% of CVD treatment procedures [47]. A stent may be

used as an alternative to, or in combination with, angioplasty.

Page 23: Peptide-modified polymer for endothelialization

11

In recent years doctors have used new stents, some of which are covered with

drugs that help prevent the blood vessel from re-closing [48]. These new stents have

shown some promise for improving the long-term success of this procedure [2]. Also,

stent-grafts, in which synthetic material covers the interstices of the stent [49], have

become a possible solution to increase long-term stent patency. Again, difficulties arise

when using the stent in small-diameter vessels. The stent may not be flexible enough to

travel through the necessary vasculature to reach the occluded vessel. Significant damage

can also occur to the endothelial lining as the stent moves to the desired location. The

disturbances in flow caused by the expanded stent often bring about restenosis. This is

because of the different flow characteristics of the small-diameter vessel and can result in

increased activation of surrounding blood elements.

2.3.3 Invasive Surgical Methods 2.3.3.1 Grafting

When a blood vessel becomes damaged or occluded, it often must be replaced.

Venous autografts have been the ideal graft since the vascular graft was introduced in

1949 [3]. Autografting, generally using the autologous saphenous vein (ASV), is

preferred because of increased patency and fewer post-surgical complications. The ASV

is removed from one of a limited number of harvest sites and sutured above and below

the damaged site to redirect blood flow. After surgery the graft begins to remodel: a 60-

70% loss of endothelial cells within 48 hours, then re-endothelialization after 6-8 weeks

Page 24: Peptide-modified polymer for endothelialization

12

[7]. Each year, 300,000 patients in the United States receive a venous autograft to bypass

an occluded or narrowed coronary or peripheral artery.

However, 10-30% of patients who need vascular bypass surgery do not have

suitable saphenous veins due to prior harvesting or disease damage [3, 50]. Alternative

biological grafts can occasionally be used. The internal mammary artery (IMA) is used

extensively for coronary artery bypass grafting and has better long-term patency rates

than ASV [51] (Table 1). Less desirable, but used as needed, are allografts and

xenografts. The allograft of choice is the gluteraldehyde-treated umbilical cord vein graft

[50]. Allografts have been able to reach patency rates similar to ASV in large diameter

applications. Xenografts can result in a high occurrence of thrombus formation, infection,

and rupture [3], but modified bovine heterografts are nevertheless infrequently used [50,

51].

ASV grafts are not completely successful. They typically have five-year patency

rates of only 30-75% for femoropopliteal bypasses [7]. One of the causes of saphenous

graft failure is graft disease [52], which follows the same progression of events discussed

earlier to reach atherosclerosis. One possible cause for graft disease is the damage done

to the endothelial cell lining during harvesting of the vein. The loss of this layer results in

accumulation of fibrin at the luminal surface and the activation of coagulation elements.

Graft occlusion occurs following a series of events: first trauma to the natural vessels

brings about the implantation procedure; then the implanted graft produces fluid dynamic

disturbances in the blood flow that cause endothelial cell damage. Because the graft is

also thrombogenic, blood derived mitogens act on exposed smooth muscle cells leading

to hyperplasia [53].

Page 25: Peptide-modified polymer for endothelialization

13

2.3.3.2 Synthetic Grafts

Synthetic alternatives to autografts such as ePTFE and Dacron have the advantage

of being readily available. However, synthetic prosthetics suffer from decreased patency

due to factors including smooth muscle cell hyperplasia and thrombus formation. The

characteristics of an ideal prosthetic graft include biocompatibility with tissue and blood

elements, nonthrombogenicity, durability, proper interaction with surrounding connective

tissue, resistance to the formation of aneurysms, and acceptable mechanical properties.

Synthetic vascular grafts used in large diameter vessel applications have achieved

a relatively high degree of patency (Table 2). However, in medium and small-diameter

grafts patency remains greatly reduced [51]. Failure within 30 days of implantation is

primarily caused by thrombosis; failure within six months is generally attributed to

intimal hyperplasia [50]. The size difference has been observed to greatly affect the

performance of grafts. The first clinically used synthetic vascular implants, developed in

the 1950s, were made from woven or knitted Dacron, (polyethylene terephthalate, PET),

and Teflon. Despite possessing thrombogenic surfaces and demonstrating a lack of elastic

compliance, they functioned sufficiently well in large-vessel applications. The grafts

were incorporated in fibrous tissue and infection rarely occurred [7]. However, these

vascular grafts were inadequate when used in medium or small diameter applications.

There are several reasons why synthetic grafts function well in large but not in

small diameter applications. The increased surface to volume ratio in smaller-diameter

vessels results in an increased activation of surrounding blood elements, which leads to

thrombosis. A reduced blood flow volume also causes an increase in the contact time

with the luminal graft surfaces. For these reasons, synthetic grafts are rarely used for

Page 26: Peptide-modified polymer for endothelialization

14

bypass or reconstructive procedures in small diameter applications, including the

coronary artery or arteries below the knee. Additionally, compliance affects the function

of small-diameter grafts. Compliance, in this application, is generally a structural and not

a material property so it depends on the geometry (diameter and wall thickness) of the

blood vessel [54]. It is desirable that the compliance match the compliance of the vessel it

is replacing in order to avoid a discontinuity of blood-flow velocity and potential stagnant

regions [18]. The smaller grafts are unable to meet the compliance requirements for

small-diameter vessels.

Currently, the material with the best patency is expanded poly

(tetrafluoroethylene), ePTFE [7]. It exhibits satisfactory tensile strength,

thromboresistance, and resistance to neointimal hyperplasia in large-diameter

applications [55] and shows partial healing with endothelial cell coverage [56]. However,

both ePTFE and Dacron have unacceptable long-term patency rates in small-diameter

applications [8, 51, 57].

2.3.3.3 Biodegradable Synthetic Grafts

A new trend in biomaterials has been the use of biodegradable polymers [58].

Biodegradation generally implies a material that can be broken down by hydrolysis or

enzymatic mechanisms. This process has also been referred to as absorption, erosion, and

resorption in the literature [59]. In principle, cells attach and grow on a tissue scaffold;

the cells proliferate and are eventually able to adsorb and replace the scaffold.

Biodegradable materials have several advantages over non-biodegradable materials: they

do not elicit a permanent, chronic foreign-body response; they can be used for tissue

Page 27: Peptide-modified polymer for endothelialization

15

engineering; and they can be used to deliver drugs to the implantation site. However,

using degradable polymers in bio-application requires consideration of two inter-related

processes: (1) the degradation of the polymer due to blood contact, and (2) the interaction

of degradation products and the body. Poly (L-lactic acid), PLA, is especially useful due

to its slow biodegradation, useful mechanical properties, and especially, prior FDA

approval in other applications [60]. This family of polymers was first used for

biodegradable wound closures and now is being investigated for vascular applications.

However, the acidic degradation products have been shown to cause inflammation and

swelling [61] in vascular applications.

2.3.3.4 Alternative Materials

In very recent years a large variety of polymers have been and are being evaluated

as possible vascular materials. At the top of the list are many polyurethane based

materials like polyurethane with plasminogen [62], elastomeric polyurethanes [63], and

peptide modified gold-coated polyurethanes [64]. Others possible vascular materials

include poly(ether-amidoamines) and polyamidoamines [65], plasma surface modified

poly(ethylene terephthalate) [66], hydrogels [67], and many more. However, no polymer

or polymer combination has yet been able to meet all the requirements for a vascular

material. For instance, the adverse effects of polyurethane degradation products have

been observed to reduce patency [7] in vascular applications. In spite of this, several

common trends have been observed in potential vascular materials; surface or bulk

modification and the use of endothelial cell seeding to improve biological interactions.

Page 28: Peptide-modified polymer for endothelialization

16

2.4 Polymer Modification

In 1955 Edwards and Tapp first attempted to chemically modify a commercially

available polymer to develop a negatively charged coating promoting heparin attachment

[68]. There are many different ways to chemically modify a polymer to change the

properties (Table 3). Since Edwards and Tapp’s first attempt, many others have used

surface chemistry in an attempt to increase the patency of vascular grafts [62, 65, 66, 69-

75]. A great deal of work has focused on altering the surface chemistry of a substrate to

specifically affect cell functions like motility, proliferation, and adhesion [76].

Attachment of physiologically active substances such as fibronectin [71] and heparin

[77], to substrate surfaces has been investigated. A number of natural and synthetic

growth factors have been identified that are capable of stimulating endothelial growth

[78] or regulating the initial adsorption of proteins and deposition of platelets and blood

cells [79, 80]. For example, the synthetic surface can be treated with fibronectin (FN).

The FN has a high affinity for binding heparin, with two classes of affinities (with

binding sites located towards opposite ends of the molecule) Kd=100 nm to 1.0 nm [80].

FN-coated PTFE grafts retained more than six times the number of seeded endothelial

cells compared to untreated controls at 24 hours following the restoration of blood flow

[71]. Despite advances with the Teflon graft, intimal hyperplasia [53, 81], restenosis, and

occlusion remain severe in most small-diameter applications. To date, no clinically

applicable small-diameter vascular prostheses have been developed [82].

Page 29: Peptide-modified polymer for endothelialization

17

2.4.1 RGD Containing Peptides

A major objective in the design of functionalized polymer surfaces for tissue

engineering applications has been the covalent attachment of peptides that regulate cell

adhesion: specifically those that promote integrin-mediated cell attachment [83, 84].

Since the vast majority of mammalian cells are anchorage-dependent, they must attach

and spread on a substrate to proliferate and function normally [23, 76]. By controlling the

surface properties of the substrate, cell attachment and growth in vitro can be altered [85].

A great deal of work has focused on RGD containing peptides to regulate cell adhesion

[83, 86-88].

Peptides containing the RGD (R=arginine, G=glycine, D=aspartic acid) sequence

are considered important small cell-adhesive ligands [83, 86]. This cell-binding sequence

is present in adhesive proteins like fibrinogen, vitronectin, collagens, and fibronectin [89,

90]. Membrane proteins of blood platelets, endothelial cells, and several other cell types

can bind RGD-containing peptides whether the peptide is in solution or immobilized onto

a solid surface [87, 91]. For example, a 25-residue RGD-containing peptide binds to the

α4β1 integrin receptor with 40% of the activity of fibronectin [92].

In an effort to promote cell adhesion onto biodegradable implants, RGD peptides

have been covalently grafted onto poly(amino acid-lactic acid) copolymers [93] and

polystyrene [94]. Modification of both PET and PTFE arterial patches with an RGD-

containing cell adhesion peptide was reported to improve healing characteristics after

implantation in both dogs and sheep [88]. However, the mere presence of the RGD

sequence does not mean that the protein will possess adhesive properties [80].

Page 30: Peptide-modified polymer for endothelialization

18

Other potentially useful, non-RGD containing peptides have been identified for

vascular applications, which include consensus amino acid sequences such as

(XBBXBX) (X= hydrophobic residue, B=positively charged basic residue). These are

shown to be involved in heparin binding [84, 95, 96]. These peptides are discussed in

more detail later.

2.4.2 Endothelial Seeding

Endothelial seeding is a process where endothelial cells are placed on the luminal

surface of a vascular graft and allowed to proliferate and migrate to cover the surface

before implantation. The concept of endothelial cell seeding was initially conceived to

address the low patency rates resulting from using small-diameter vascular prosthesis [8].

The first successful attempt to seed a synthetic vascular graft with endothelial cells was

achieved in 1978 by Herring et al [97]. A great deal of research has been conducted on

the topic since this initial discovery. Results have shown that endothelial seeding

promotes endothelialization in some cases, which increases patency [98-100]. Animal

studies have demonstrated the resistance of seeded grafts to thrombosis and

pseudointimal hyperplasia [3, 8, 101]. One of the disadvantages of graft seeding is the

increased time required for the cells to multiply and cover the graft surface. In addition,

many cells are lost shortly after implantation and the cell functions are not easy to control

[55]. Despite successful endothelialization in animal studies, clinical evaluation of

endothelial cell seeded grafts has not been as successful; significant neoendothelialization

of vascular prostheses, beyond 1-2 mm, has not yet been observed in humans [8, 19].

Page 31: Peptide-modified polymer for endothelialization

19

3 Design Concept

The overall object of the project, of which this research is a part, is to engineer a

vascular substitute that can (1) be directly implanted without requiring secondary surgery,

and (2) increase patency. The specific goal of this part of the project was to modify a

polymer with a peptide to produce a surface that actively promotes endothelial cell

coating. Since surface chemistry affects protein adsorption, cell adhesion, and plays an

important role in blood compatibility, a polymer has been chosen that allows surface

modification. The functional groups on the surface are used for derivatization with

peptides having an affinity for endothelial cell PGs.

3.1 Material Selection

Poly(DTE-co-X%DT carbonate) is the polymer of choice for the development of

synthetic vascular graft material. Poly(DTE-co-X%DT carbonate) is a copolymer based

on the natural amino acid tyrosine, with a functional pendant group (Figure 2a) [102].

This polymer is termed a pseudo-poly(amino acid), because the amino acids are linked

together by non-amide bonds [103]. The use of these backbone-modified pseudo-

poly(amino acid)s was first investigated in 1984 by Kohn et al [104]. The polymerization

process controls the amount of free acid available. The amount of free acid available is

indicated by the addition or removal of the R protecting group (Figure 2b). When R =

CH2CH3 the repeat unit is desaminotyrosil-tyrosine-ethyl ester (DTE); when R = H the

free unit is desaminotyrosil-tyrosine (DT). X is the molar fraction of repeat units with

pendant acid groups. When the R group is absent, the polymer has 100% free acid

Page 32: Peptide-modified polymer for endothelialization

20

groups. The amount of free acid in turn controls the functionality and the degradation rate

[105]. In this project, the functional group is used to bind a peptide via a lysine bridge.

The attachment of the lysine bridge is illustrated in Figure 2c. The degradation rate can

vary from a few days to several years, depending on the free acid and environmental

conditions. With a DT content of X=0%, the polymer requires 366 days to reach 50% of

the starting molecular weight when implanted subcutaneously in rats [106]. The polymer

degrades by random hydrolytic chain cleavage that results in an increase in polydispersity

[106]. As the DT content increases, the polymer becomes more hydrophilic and anionic

in nature, and more susceptible to hydrolytic degradation [107]. In vitro cytotoxicity and

short term in vivo evaluations in rats have shown tyrosine-derived polycarbonates to be

generally biocompatible [58, 105, 108] and to support attachment and growth of various

cell types [109, 110]. The degradation products are non-acidic [111], especially when

compared with PLA [105]. The pendant functional groups provide a location to

covalently attach functional side groups, including peptides. Poly(DTEC-co-X%DTC) is

completely amorphous, with a Tg of approximately 80°C (176°F) [112], and is readily

formed [109] using common melt and solution techniques. The tensile strength (Table 4)

is also greater than that of thoracic tissue.

3.2 Heparin-Binding Peptides

Similar to the RGD containing peptides, heparin-binding peptides have been

designed to specifically bind PGs and can therefore be inferred to be useful in promoting

endothelialization [96]. These peptides are based on the heparin binding (XBBXBX) or

(XBBBXXBX) (X= hydropathic residue, B=positively charged basic residue) sequence

Page 33: Peptide-modified polymer for endothelialization

21

discovered by Cardin and Weintraub [95]. Instead of binding to an integrin, these

peptides target PGs. These peptides are less specific than RGD containing peptides and

are capable of binding to all of the PGs on the cell surface, including those carrying

chondroitin, dermatan, keratan, and heparan sulfates. RGD containing peptides, on the

other hand, are only capable of binding to specific integrin molecules. This is because the

RGD containing peptides are integrin receptor mediated [90], whereas the heparin

binding peptides attract PGs electrostatically [84]. In addition, RGD containing peptides

must use a spacer arm to achieve a suitable distance between the substrate and the RGD

end of the peptide for cells to attach. The RGD sequence must be available at the surface

of the substrate and its conformation must fit integrin receptors for attachment to occur.

In this particular aspect of the project the (ARKKAAKA) sequence, with alanine,

arginine and lysine (ARK), is used. The important aspects of peptides with regard to

binding are generally surface charge, structure, and attachment location [113]. The

predominant molecules on endothelial cells are PGs, which contain a core protein with

one or more covalently bound anionic glycoaminoglycan (GAG) chains [27]. The heparin

binding peptide provides positively charged sites to bind these PGs [95, 96]. GAGs

participate in the adhesion of endothelial cells to the extracellular matrix or another

substrate. Heparan Sulfates (HS) comprise 85% of the GAG on endothelial cells [114]. If

a peptide with a high affinity for heparin is designed, the same peptide should also have a

high affinity for endothelial cells.

Modeling these peptides predicts arrangement of the amino acids into an α–helix

[96], which has been confirmed by circular dichroism [115]. The helix structure gives a

three-dimensional arrangement of multiple heparin-binding consensus sites; it allows for

Page 34: Peptide-modified polymer for endothelialization

22

clustering of noncontiguous basic amino acids on one side of the helix, thus forming a

charged domain to which GAGs can bind. The formation of this structure has been found

to be critical to the binding efficiency [96]. To achieve this, the peptides contain multiple

copies of the eight-mer ARK consensus. In addition, the amino acids that make up the

peptide should be conducive to helix formation. Alanine is used because of its stabilizing

activity on α-helices [116] and the basic amino acids are chosen to represent those with

the highest probability of occurrence in each basic position in the heparin-binding

consensus sequences of native heparin-binding proteins [95]. For the peptide to take an

active role in binding, it must bind specifically at the end of the chain and posses

sufficient chain length to assume the helical structure. The total length of the peptide also

plays an important role in the binding efficiency; maximum binding of the peptide to low

molecular weight heparin reaches a plateau around 30 amino acid residues [96].

Page 35: Peptide-modified polymer for endothelialization

23

4 Experimental

Specific experiments designed to functionalize and characterize the polymer are

discussed. Preliminary tests for the effect of the polymer and functionalized polymer on

cell functions are also discussed.

4.1 Polymer Derivatization

The aim of the project was to produce a polymer surface with peptides covalently

attached. P(DTEC-c-X%DTC) polymers with 0, 5, 13, 23, 35, 50% free acid were

provided by Integra LifeSciences Corporation (San Diego, CA). The functional groups on

the surface are used for derivatization, via a lysine bridge, with peptides having an

affinity for endothelial cell PGs. By creating this functionalized polymer surface with

attached peptides, protein adsorption and cellular interactions may be controlled.

L-lysine ethyl ester (lysine) (Sigma, Lot #17H1152) was attached to the free acid

group on p(DTEC-c-X%DTC) (X>0%) as a linking agent to connect peptide chains to the

polymer. By using lysine as a bridge, connection to the polymer at the unique carboxylic

acid terminus of the peptide is assured. This leads to a basic surface to promote cell

interaction. Without the lysine bridge, binding could occur at any lysine residue

containing point along the peptide chain. The reaction was conducted in dilute solution

with excess lysine to inhibit multiply attached lysines and to drive the reaction to

completion.

The polymer was reacted in solution with excess lysine and 1,3-

dicyclohexylcarbodiimide (DCC) to catalyze the reaction. DCC has been shown to couple

Page 36: Peptide-modified polymer for endothelialization

24

amines and carboxylic acid groups with a high efficiency [60, 117, 118]. While pDTEC

is soluble in organic solvents including chloroform and methylene dichloride, the desired

attaching groups, including lysine, polyamino acids, and the peptide, are not. This

allowed for easy removal of unreacted molecules, which were filtered from the polymer

solution with a 0.45µm nylon filter. To ensure removal of all unreacted species, the

solvent was evaporated from the solution to make thin films. These films were then

washed with water and dried to further remove water-soluble lysine and peptides. This

process was repeated several times to be sure all unattached molecules were removed.

The process for attaching lysine was repeated with the other polyamino acids and the

synthesized peptide to attach them to the lysine functionalized polymer.

Polylysine (Sigma, Lot# 30K5906, Mr=9,000 g/mol) and polyalanine (Sigma,

Lot# 42H5546, Mr=5,000 g/mol) were used to simulate the attachment of the peptide.

Polylysine is a hydrophilic model for the charged heparin binding peptide. Polylysine

will not form an α-helix alone in physiological solution because of the density of charged

groups along the chain, but when attached to a substrate has a tendency to form the α-

helix. This is similarly true for the peptide [96]. Polyalanine, on the other hand, is a

hydrophobic model for the peptide. In addition, since the amino acid alanine is the first

and last amino acid found in the peptide, polyalanine simulates the specific attachments

possible by the ARK peptide. The peptide is also composed of 50% alanine, which has a

stabilizing effect on the α-helix. Polyalanine with this molecular weight has a chain

length similar to that of the peptide.

The specific peptide attached to the poly(DTEC-co-X%DTC) via the lysine

bridge is (ARKKAAKA)4. This peptide has sufficient length to form an α-helix and also

Page 37: Peptide-modified polymer for endothelialization

25

has been demonstrated to bind heparin (Kd≈50 nm) and endothelial cell PGs with a

Kd≈300 nm [96]. The general advantage to covalently binding a peptide to a substrate is

the strong chemical bond present to prevent desorption [117]. Peptides were

commercially obtained from the University of Virginia Biomolecular Research Facility

(Charlottesville, VA) or Genosys Biotechnologies (The Woodlands, TX). Peptides were

synthesized by standard FMOC solid phase synthesis. Peptide Mr was certified by mass

spectroscopy, and purity certified by HPLC.

4.2 NMR Spectroscopy

NMR was performed to confirm the attachment of the lysine and peptides and to

determine the efficiency of the attachment. The structure of the p(DTEC-c-X%DTC), and

the attachment of various molecules, was studied using both 1D and 2D NMR

spectroscopy. NMR measures the energy produced as a hydrogen proton’s spin returns to

the equilibrium state in an applied, pulsed magnetic field, providing information about

the chemical structure of the polymers.

NMR spectra were recorded on a Bruker AMX 600 NMR spectrometer (3 channel

MCI) operating at 600 MHz (1H) and a temperature of 300 K. Assignments of the spin

systems were based on double quantum filtered correlated spectroscopy, COSY, and total

correlated spectroscopy, TOCSY, experiments. These two-dimensional techniques

measure correlation in related spin systems, which were used to determine connectivity

between subunits. This allowed spectral peaks to be mapped to specific molecules in the

1-D spectrum. Peak integration from the 1-D spectrum indicated the number of hydrogen

protons producing a given peak. Peak integration values have a deviation of ±5.0% [119].

Page 38: Peptide-modified polymer for endothelialization

26

At least two samples were run at least twice (minimum of four points) to determine

integration values.

4.3 Cell Culture

Cell attachment and growth studies were conducted with human umbilical vein

endothelial cells (HUVEC). Cells were isolated from human umbilical cords [120] and

cultured [121] in medium 199 (Gibco BRL) with 10% FBS, 80 µg/mL endothelial growth

supplement (ECGS), 50-60 µg/mL heparin from pig intestinal mucosa (Sigma: Grade I-

A), penicillin, streptomycin and fungizone, on gelatinized tissue culture flasks (Falcon).

Cells were incubated at 37°C and 5% CO2. HUVEC from sub-confluent cultures of less

than passage seven are removed from the culture flask using trypsin/EDTA for five

minutes. When using passage greater than seven, the cells begin to de-differentiate and

may not function in the same manner as young cells [122].

4.4 EC Attachment

A modified cell detachment assay [123] was used to determine the degree or

strength of cell attachment to a given material. In this assay, cells are added to coated

wells, forced to adhere to the substrate at 40 g, then removed from the material by

varying centrifugal forces. The degree of force required to remove the cells corresponds

to the strength of attachment between the cells and the substrate.

Plates were prepared by coating the well bottoms of 96 well plates with p(DTEC-

c-X%DTC) X=0, 13, 23, 35, and 50%, and X=13% with lysine, lysine plus polylysine,

and lysine plus peptide attached. The materials were dissolved in dichloromethane at

Page 39: Peptide-modified polymer for endothelialization

27

≈0.05g/mL and 200 µL was applied to the well bottoms. The solvent was allowed to

evaporate in air. The plates were placed in a vacuum oven at 30 in Hg and room

temperature to draw off any residual solvent.

During assay preparation, the 96-well plates were kept on ice to reduce the

contribution of cytoskeletal and other intracellular components to cellular attachment.

15,000 HUVEC were added to each well with enough media 199 to give a positive

meniscus. The plates were sealed with tape and centrifuged using a microplate carrier at

40 g for three minutes at 4°C. This brings the HUVEC in contact with the test material.

The plates were then inverted and spun at an equal or greater force for three minutes,

again at 4°C. After centrifuging, the plates were kept inverted and incubated for two

hours to allow the attached cells to strengthen their attachment and spread on the well

bottom; unattached cells will not make contact and adhere to the test material. After

incubation, the media was removed from the inverted plates.

Because the various polymers were not transparent, spectrophotometry could not

be performed directly on these plates. The attached cells were removed to a transparent

surface for spectrophotometry. The plates were washed with 100 µL of PBS+ then 75 µL

of trypsin was added to each well and the plates were placed in the incubator for five

minutes. The trypsin was pipetted to a new tissue culture treated (TCT) plate. Cells are

able to attach and grow directly on TCT plates; they do not attach to non-TCT plates. The

original wells were washed twice with complete media and the media was added to the

wells containing trypsin in the new TCT plates. These plates were then inclubated for at

least two hours to allow cells to recover and attach. The media was gently removed and

the attached cells were fixed with 200 µL 1% glutaraldehyde and stained with crystal

Page 40: Peptide-modified polymer for endothelialization

28

violet [124]. Spectrophotometry was performed directly on both the wells coated with

fibronectin and the wells where the attached cells were transferred from the coated wells

to a new TCT plate using the above procedure. No significant variations in cell count

were seen between the two sets of plates.

To measure the amount of bound dye after the cells were fixed and stained, plate

wells were treated with 1% Triton X-100 for at least four hours to solubilize the cell

membranes and release the bound stain. The concentration of crystal violet is then

measured as A60015 using a spectrophotometer (Molecular Devices). A standard curve was

generated using cell quantities from 500 to 25,000/well on tissue culture treated wells.

Fibronectin was used as a positive control; cells show a linear decrease in attachment

with increasing force similar to that discribed in the literature [96, 115]. The maximum

force, 900 g, was chosen because it is large enough to remove most cells from

fibronectin. Forces were chosen at fairly even increments: 100, 300, 600, and 900 g.

MeCl2 washed wells were used as a negative control and show no cell attachment at any

force. The experiments were repeated at least three times for accuracy; the standard

deviation is ±4%.

4.5 EC Growth

Cell functions such as adhesion and growth depend on the nature of the substrate,

of the hydrophobic or hydrophilic character of its surface [74]. For this reason, polymers

with varying functional levels are studied to determine the effect of the surface on three

cell functions: attachment, viability, and division. The growth behavior of HUVEC on

P(DTEC-c-X%DTC) and several control materials was evaluated over seven days. Glass

Page 41: Peptide-modified polymer for endothelialization

29

slide cover slips (18mm dia., 0.15mm thick, Fisher Scientific) were solution coated with

P(DTEC-c-X%DTC) at X=0, 13, and 35%; PET (Eastpak 12270), PVC (Cognis), and

PLA (Sigma 33135-50-1), at a dry weight of roughly 0.005-0.010g. Samples were UV

sterilized for 20 minutes. The cover slips were then placed in the bottom of a six-well

plate (Falcon) with 50,000 cells suspended in media 199, and 2 mL of complete media.

This allows enough surface area for cells to grow and spread over the seven-day period.

Cells were incubated at 37°C and 5% CO2 and observed at 1, 3, 5, and 7 days under an

optical microscope at 100X.

After the seventh day, the cells attached to the coated slides were removed and

counted. The cells remaining in the well, attached to the well bottom adjacent to the glass

slide, were fixed and stained using a modified Wright-Giemsa staining protocol (Hema 3

stain set, Fisher Scientific). These cells were examined on an inverted Olympus phase

contrast microscope at various magnifications to evaluate signs of abnormal cell

functions. These experiments were conducted three times using three identical samples

per experiment. The cell counting accuracy is 95%.

The cell density was calculated by counting cells in four representative 1cm2

regions in the optical image at 23x. The accuracy of this method is ± 1.0x104 cells/cm2.

Page 42: Peptide-modified polymer for endothelialization

30

5 Results and Discussion 5.1 NMR

With NMR, the attachment points and the attachment efficiency of different

molecules can be determined. In 1-D 1H NMR, resulting peaks correspond to the number

of hydrogen protons. Cross peaks in 2-D correlated spectroscopy (COSY) and total

correlated spectroscopy (TOCSY) NMR show spin-spin coupling between hydrogens.

Both 1-D and 2-D techniques were used to determine binding efficiency. Once peaks

have been mapped to certain molecular subunits via 2-D NMR, ratios of 1-D peak heights

give the relative amounts of subunits present.

Peak assignments for p(DTEC), obtained using TOCSY (Figure 3), were shown

to be similar to those of Hooper et al [105, 106]. In the TOCSY spectrum peak

correlation systems are identified as indicated via the solid connector lines. This enables

the assignment of the spectral peaks. A key peak at 4.7 ppm is taken as the primary

hydrogen (α). This hydrogen is in a system with the γ (6.0 ppm) and β (3.0 ppm)

hydrogens, as seen in the TOCSY spectrum. To determine exactly which peak is from

which hydrogen, peak integration must be performed. Integration values indicate the

number of hydrogens responsible for an individual peak. Upon integration, the α-peak is

set to 100.0 and values of 198.0 (β), and 99.3 (γ) are obtained, giving values of 200 and

100 within experimental error. This indicates that the β-peak has double the hydrogens

and the γ-peak has equal the hydrogens as the α-peak, which leads to the identifications

shown in Figure 2. The ε (2.4 ppm) and δ (2.7 ppm) hydrogens are likewise correlated,

but separate from the α-β-γ system. These peaks give integration values of 198.7 (ε), and

Page 43: Peptide-modified polymer for endothelialization

31

201.5 (δ), which are both close to 200, indicating two hydrogens per peak. By the same

analysis, the peaks at 4.1 and 1.2 ppm correspond to the CH2-CH3, R2 and R3 hydrogens,

respectively, of the pendant group. The total intensity in the CH2-CH3 peaks should

integrate to two and three times the α-peak, as there are two and three protons in the R2

and R3 positions, respectively.

P(DTEC-co-X%DTC) has a more complex structure because both the free and

protected forms of the acid are present. Changes in several peaks are expected to occur

because of the heterogeneous nature of the structure. There is no longer a pendant group

for every α hydrogen. In addition, peak shifting is expected for the hydrogens in close

proximity to the pendant group. The 1D NMR spectra of both p(DTEC) and p(DTEC-co-

13%DTC) are shown in Figure 4. The integration values are shown below each peak.

When the α hydrogen integration is set to one hundred in the situation where the material

is p(DTEC-co-X%DTC), integration values of less then two and three hundred are

expected for the R2 and R3. Integration values of 172.1 and 269.2 are obtained for the

CH2 and CH3 peaks, respectively, corresponding to values of 14 and 11% DT. In

addition, there is a small shoulder at 3.1 ppm in the 13% DT spectrum. This corresponds

to a shift in the β-peak caused by the OH pendant group, and is 12.6% when integrated.

This confirms the samples have approximately 13% DT content. There is also a shoulder

on the γ-peak arising from the same interactions that cause the β-peak shift. This was not

used due to the difficulty in separating the two peaks. The α-, β-, and γ-peaks also should

be affected by the attachment of additional groups, including lysine. Analysis of

p(DTEC-co-35%DTC) confirms the same change in integration values (Table 5), giving

Page 44: Peptide-modified polymer for endothelialization

32

values for the CH2 and CH3 peaks corresponding to values of 37 and 33% DT,

respectively.

With the attachment of lysine to the polymer free acid, new peaks will be present

in the NMR spectrum. To understand the relation of these new peaks to the polymer it is

first useful to determine how the lysine peaks relate to one another in the absence of the

polymer. Figure 5 shows the 1-D and 2-D NMR spectrum for lysine and how these peaks

correspond to the lysine molecule. By comparing these coupled peaks to the spectrum of

the polymer with the lysine attached, several unique lysine peaks can be identified.

A 2D spectrum of p(DTEC-co-13%DTC) with lysine attached is shown in Figure

6. There is a large concentration of additional peaks between 0.8 and 4.5 ppm. These

have been determined to be lysine based on the previously discussed coupled systems.

The spectrum becomes more complex with the addition of lysine, both by the possibility

of α- and ε- terminus attachment of the lysine to the polymer backbone and by the

interactions between the lysine and the polymer. Multiple attachment locations produce

groups of similar peaks that are slightly shifted from one another. The two circled regions

on the graph are consistent with multiple attachment locations. There are a large number

of as yet unidentified peaks, as well. A corresponding 1D spectrum between 4.0 and 5.0

ppm (Figure 7) has integration values for peaks assigned to the α at 4.7 ppm and a yet

unassigned lysine peak at 4.38 ppm. The α-peak is a function of the polymer backbone

and can be considered independent of reaction and set to unity. The ratio of the α-peak to

the lysine peak gives approximately 11.5 ± 0.65% attachment of lysine. Attachment of

the lysine to p(DTEC-co-35%DTC) gives a ratio of the α-peak to the lysine peak of 30.2

Page 45: Peptide-modified polymer for endothelialization

33

± 1.5%. This corresponds to attachment efficiency between 85-90%. The NMR spectra

give clear evidence of lysine attachment.

Once the lysine attachment was verified, polylysine, a homo-polyamino acid, was

attached to the lysine amine terminus. Polylysine was chosen to simulate the attachment

of the peptide. Polylysine is a charged poly(amino acid) with similar chemical and

physical properties to the target peptide. TOCSY analysis (Figure 8) shows an increase in

intensity of the correlated peaks assigned to lysine. This is consistent with attachment to

the polymer, and further supports the original assignment of these correlation systems.

However, the polylysine peaks occurred at the same locations as the lysine peaks and

could not be quantified to determine attachment efficiency.

To further investigate attachment of peptides to the polymer via the lysine bridge,

polyalanine was attached. This molecule was also chosen to simulate the attachment of

the peptide; it is another polyamino acid chemically similar to the peptide, conserving the

hydrophobic nature and the tendancy to form an α helix. Unlike polylysine, polyalanine

shows unique peaks in the NMR spectrum. Figure 9 shows the 1D spectrum of the

polymer with the polyalanine attached. Peak integration values have been added for the α

polymer peak at 4.7 ppm and several peaks determined to be polyalanine peaks from a

TOCSY spectrum. The TOCSY spectrum is difficult to understand without more detailed

NMR analysis so it is not shown in this paper. The α polymer peak has been set to 1.0 so

the peak ratios can be calculated. The peaks at 4.0 and 3.5 ppm, uniquely, from

polyalanine do not overlap peaks from the polymer or lysine and can be used to

determine percent polyalanine attachment. However, the peaks at 3.2, 1.9, and 1.7 ppm

do not give clear attachment data. The integration values (Table 6) for these peaks are

Page 46: Peptide-modified polymer for endothelialization

34

more than would be expected from polyalanine alone because they contain areas from

other peaks occurring in the same location. The peaks at 1.9, and 1.7 ppm also

correspond to the lysine β- and δ-peaks shown in Figure 5. The peak at 3.2 ppm occurs

when the lysine has been attached to the polymer.

Table 6 shows integration values for the peaks at 4.0, 3.5, 3.2, 1.9, and 1.7 ppm.

The particular values of interest are the 4.0 and 3.5 peaks. These peaks are approximately

zero for X=0% polymer. As the functionality increases, the amount of polyalanine

increases significantly. The data in Table 6 has been normalized by subtracting the X=0%

value from the values for X=23 and 50%, and is shown in Table 7. The attachment is

30% for p(DTEC-c-23%DTC)and between 60-70% for p(DTEC-c-50%DTC). This

implies that 30% of the pendant groups on the polymer backbone of X=23% and between

60-70% of the pendant groups on the polymer backbone of X=50% have polyalanine

molecules attached. However, it would be expected that no greater amount of polyalanine

could be attached than the degree of functionalization. It is unclear what is causing the

increase in measured attachment of the polyalanine. One possibility is an increase in acid

functionality during the various treatments, caused by degradation of the polymer

backbone. However, an increase in attachment is not seen with the attachment of lysine,

as would be expected if this were the case because it uses the same chemical process for

attachment. Another possibility arises from potential subunit exchange when the

polyalanine chains come into close proximity with each other. If a subunit exchange

occurred it would likely change the molecular weight of the attached molecules. This

change in molecular weight may alter the spectral peaks. To assure proper quantification,

Page 47: Peptide-modified polymer for endothelialization

35

peaks must be identified that (1) do not overlap other polymer or lysine peaks and (2) are

not sensitive to molecular weight and side chain interactions.

While there appears to be some interference with the peaks used to measure

polyalanine attachment, Table 7 does show roughly a doubling of attached polyalanine

with a doubling of the functionality. In addition, even though the peaks at 3.2, 1.9, and

1.7 ppm do not give unambiguous attachment data, they still show increasing attachment

with increasing functionality. Even though quantification is possibly subject to some

interference, polyalanine is clearly being attached to the polymer in significant quantities.

Accurate quantification of the polyalanine or peptide attachment will require careful

selection of reference peaks.

The final NMR spectrum produced was p(DTEC-c-35%DTC) with the peptide

(ARKKAAKA)4 attached to the lysine bridge (Figure 10). From this spectrum the

addition of many new NH peaks (7-6.5 ppm) and methyl group peaks (2.0-1.2 ppm) can

be seen, as compared to Figure 6. This is expected given that multiple alanines and

arginines, in addition to lysine, make up the peptide. There also appears to be a great deal

of overlapping in peaks. This overlapping, the shifting in peaks due to multiple

attachment mechanisms, and the influence of conformational changes in the peptide

creates a spectrum that is very difficult to assign. From this spectrum, there is clear

evidence of peptide attachment. Detailed assignment will require more NMR work,

potentially using a spin labeled peptide.

Page 48: Peptide-modified polymer for endothelialization

36

5.2 Cell Attachment

This experiment is used to determine the strength of attachment between the

HUVECs and the polymer substrate. The attachment results in Table 8 show that the

HUVECs do not readily attach to p(DTEC-co-0%DTC). However, as the amount of free

acid groups on the polymer increases and the polymer becomes more hydrophilic the cell

attachment increases (Figure 11). In fact, from Figure 12, the most hydrophilic polymer

tested, p(DTEC-co-50%DTC), has attachment values close to the positive control (FN).

Since the cells do not generally attach to p(DTEC-co-0%DTC), any cellular attachment

seen with functionalization and with the addition of L-lysine ethyl ester and peptides

should be related only to the modification of the polymer.

A cell attachment assay was also performed on a variety of modified materials:

p(DTEC-c-13%DTC), with attached lysine, with polylysine attached, and with the

heparin-binding peptide attached (Figure 13). It is important to note that since the overall

attachment, for 15,000 initial cells, for this experiment is lower than expected based on

positive controls, and since the experimental error is high, the results are largely

inconclusive. There is some indication that the addition of lysine to the polymer may

increase attachment slightly while the addition of polylysine or peptide may decrease the

ability of the cells to attach to the polymer substrate. The lack of observed binding in

these systems may be caused by increased polymer solubility, unstable polylysine

attachment, or cell incompatibility.

The overall polymer-peptide-cell system may be producing a more soluble system

that the applied force can more easily remove from the well bottoms. This is supported by

other observations. In another experiment conducted in the San Antonio laboratory [122],

Page 49: Peptide-modified polymer for endothelialization

37

cells were tested for attachment to FN, blank and 8mer and 36mer peptides coated onto

well bottoms. The majority of the cells were removed from blank, 8mer and 36mer wells

but showed excellent binding to the FN. While there was no evidence of attachment in

the blank and 8mer wells, there were islands of superior cell attachment in the 36mer

wells. In separate observations, particularly rapid degradation or dissolution of the higher

functionalized polymers was observed, which is consistent with increased degradation of

the more hydrophilic p(DTEC-c-X%TDC) polymers [107]. The attachment of lysine and

peptides is expected to make the polymer system even more hydrophilic in nature. It is

reasonable to assume that this will also increase both the solubility and the degradation

rate. More work must be conducted in this area to understand any peptide changes that

occur upon binding to the polymer. If the cells are attaching to the peptide and then the

entire system is being pulled away from the well bottoms, then collecting and analyzing

the cells removed during the adhesion assay should show polymer still attached. Reacting

the polymer with less than stoichiometric quantities of lysine in order to allow the lysine

to cross-link the polymer slightly might solve the degradation problem.

Alternatively, the peptide may be changing shape when it is attached to the

polymer and this may be affecting cellular attachment functions. If the structure is not an

α-helix, the attachment locations may not be oriented for enhanced PG attachment. Steric

effects may also inhibit helix formation [117] or the peptide may not be attaching to the

polymer at the desired terminus, which would also cause reduced attachment. This type

of effect would most likely be observed as a modest instead of strong increase in binding.

This is not consistent with the observation that no significant change occurs at all.

Page 50: Peptide-modified polymer for endothelialization

38

5.3 Cell Growth

The effect of the surface on three cell functions have been studied on p(DTEC-c-

X%DTC): attachment, viability, and division. Figure 14 shows the results of the cell

growth study on p(DTEC-c-13%DTC). Table 9 shows the quantitative results of this

study. After 1 hour (Figure 14a), a number of cells have attached to the polymer. After 1

day, the cells change conformation, indicative of surface attachment and cell growth. The

cells multiply and continue to spread throughout the duration of the experiment (e.g. 3

days, Figure 14c). By day 7 (Figure 14d) there are regions of confluence.

Figure 15 shows the cell growth on a) p(DTEC), b) p(DTEC-co-13%DTC), c)

p(DTEC-co-35%DTC), d) PLA, e) PET, f) PVC (negative control), and g) blank glass

(positive control) at day 5. The cells grow more readily on the p(DTEC-co-X%DTC) as

the amount of functionalization increases. As the amount of copolymerized, non-

protected acid groups increases, both hydrophilicy and charge sites increase, either of

which may result in the observed increase. For example, cell proliferation and spreading

have been shown to increase with increasing surface hydrophilicity by Kottke-Marchant

et al [125]. The PLA and PET both exhibit moderate affinity for HUVEC growth. The

PVC shows no HUVEC affinity, as expected, because the PVC is expected to cause cell

death [3]. The heparin-coated glass shows strong affinity for cell growth, again as

expected. The final comparative, quantitative results are summarized in Table 9. The

results for the p(DTEC) bracket Dacron and PLA depending on the degree of

functionalization. All polymers are less supportive of cell growth than the heparin coated

blank glass.

Page 51: Peptide-modified polymer for endothelialization

39

Cells remaining in the six-well plate were fixed and stained and then evaluated for

signs of abnormal cell functions. Figure 16 shows the stained cells at increasing

magnifications. Figure 16a shows cells with semi-confluence, Figure 16b shows cells

spread on the plate surface, and Figure 16c shows cells in the process of dividing. Cells

spreading, confluence, and division are indicative of healthy, growing cells.

The endothelial cell lining in a healthy vessel is 105 cells/cm2 [114, 126]. Cell

density is calculated from Figure 14d. The HUVEC dimensions are roughly 50 x 10 µm,

consistent with the literature [122]. Calculations for four different regions give an

average cell density of 5.0x104 ± 1.0x104 cells/cm2. This is about half the density of cells

lining a healthy vessel. However, the relationship between the cell density in this in vitro

experiment and any in vivo applications is a matter of speculation at this point.

Page 52: Peptide-modified polymer for endothelialization

40

6 Conclusions The functionalization of a poly DTE carbonate polymer with a heparin binding

peptide has been accomplished. DTE Carbonate is a commercially available polymer that

has side groups available at varying levels for functionalization. Previous studies have

uncovered a peptide based on Cardin and Weintraub sequences that have high affinity for

EC PGs. A technique for attaching these peptides to the polymer has been developed. Use

of dilute solution insured no cross-linking between polymer chains. Excess lysine insured

maximum conversion during the attachment reaction. Since the reactants were insoluble

in organic solvents, excess reactant was efficiently removed using filtration and water

washing.

NMR has been used successfully to quantify available polymer functionality and

lysine bridge attachment and qualitatively or semi-quantitatively for additional

functionalization with long chain molecules. The peaks of the poly(DTE-co-%DT

carbonate) have been identified by NMR and attachment of the L-lysine ethyl ester to the

p(DTEC-co-X%DTC) has been verified. Lysine is shown to attach with an efficiency of

85-90%. Binding of polylysine, polyalanine and peptide establish that peptide binding

can be accomplished. Changes in the spectra confirm the attachment of polylysine,

polyalanine, and the (ARKKAAKA)4 peptide. Semi-quantitative results with polyalanine

indicate a large degree of attachment, although some interference among the spectral

peaks has been noted. As alanine provides the carboxylic acid terminus with which the

peptide will attach to the polymer lysine bridge, it is expected that the peptide will bind

similarly. Integration of polyalanine peaks shows increasing attachment that is directly

proportional to the increase in polymer functionality.

Page 53: Peptide-modified polymer for endothelialization

41

Cellular studies have confirmed that the material is non-cytotoxic and increasing

hydrophilicity increases the affinity of HUVECs. Preliminary results from experiments

with the peptide attached to the polymer were inconclusive. Additional work must be

completed to study the cell affinity and adherence to the attached peptide.

The NMR results demonstrate the feasibility of chemical functionalization of

DTE Carbonate, resulting in a peptide-modified polymer surface. The cell work

demonstrates that the materials are biocompatible and support cell growth. The premise

of making a vascular graft seems reasonable based on this data. A graft from this material

could result in a synthetic vascular graft that is much closer to the biology of native

vessels than currently available grafts. A graft from this material could have a significant

impact on the treatment of CVDs.

6.1 Recommendations: Characterization 6.1.1 Peptide Attachment

Quantifying the amount of peptide attached to the polymer via the lysine bridge

has yet to be determined. Quantification of attachment will be critical to determining the

efficiency of the cell attachment and regulating cell density. To quantify the attachment, a

spin- or radio-label containing peptide can be attached to the polymer. This should

provide a unique and easily identifiable peak to compare with known polymer peaks.

Page 54: Peptide-modified polymer for endothelialization

42

6.1.2 Peptide Structure

An α-helix secondary structure, or the ability to form it, may be important in

binding affinity of endothelial cells to the peptide [96]. Circular dichroism cannot be used

on materials that are not in solution, so a helix-sensitive peak in NMR or IR must be

located.

6.1.3 Cell Affinity

Tests of the affinity of HUVECs for peptide-modified polymer surfaces have been

inconclusive. One possible cause is degradation and subsequent removal of the entire

polymer-peptide-cell system with an applied force. One method to reduce degradation is

to cross-link the polymer. This can be accomplished by reacting the polymer with less

than stoichiometric amounts of lysine, in a more concentrated solution. This will allow

cross-linking between two polymer chains via the lysine bridge. Cross-linking the

polymer should increase the overall strength of the polymer-peptide-cell system.

In addition, variations in the cell attachment assay may be required. For example,

collection and analysis of the material removed from the well bottoms during

centrifugation may reveal polymer-peptide-cell systems.

6.1.4 Polymer Selection

Once the peptide-modified polymer has been completely characterized it will be

necessary to determine the cell density produced by seeding this surface with HUVECs.

The density can then be tailored by altering either the chemistry involved to attach the

Page 55: Peptide-modified polymer for endothelialization

43

lysine and peptide, or by using a polymer with more or fewer acid pendant groups. The

resulting cell density must then be optimized with the mechanical properties and

degradation rate produced from the polymer-peptide system. Cross-linking the polymer

via the lysine bridge can alter both the mechanical properties and degradation rate.

6.2 Recommendations: Fabrication 6.2.1 Processing

Once the peptide-polymer combination has been proven as a vascular material, it

must be made into vascular graft constructs. This can be accomplished by multi-layer,

two-dimensional braiding. This technique can make three-dimensional structures where

the braid parameters allow the tailoring of properties, like fiber volume fraction, porosity,

strength, and compliance.

Another possible technique to form the vascular graft construct is by extrusion.

This polymer has been extruded at a temperature of 80°C to 90°C above the glass

transition temperature [106]. However, the extrusion process would not allow for creating

and adjusting porosity.

Other possible techniques include compression mesh [127-129], solvent

casting/salt leaching [130, 131], emulsion freeze-drying [132], expansion from

pressurized carbon dioxide [133, 134], and phase separation [135-137]. These techniques

have all been shown to produce highly porous biodegradable polymer scaffolds [138].

Any technique chosen to fabricate the vascular graft construct would require

careful examination as to the effect of the processing on the attached peptides. Attaching

Page 56: Peptide-modified polymer for endothelialization

44

the peptide after the construct has been formed is a possible means of avoiding the

aforementioned problems.

6.2.2 Sterilization

Any surgical material used in the body must first be sterilized to prevent infection.

Sterilizability of biomedical polymers is important because polymers have lower thermal

and chemical stability than metals and ceramics [139]. Sterilizing techniques include dry

heat, autoclave, radiation, and ethylene oxide gas [140]. In dry heat sterilization the

temperature ranges between 160-190°C, which is above the melting and softening

temperatures and can cause deformation of the pDTEC construct or destruction of the

peptide. Autoclaving is performed under high pressure and higher temperatures ranging

between 125-130°C [75]. This technique would also cause significant degradation and

deformation of the pDTEC construct because of water vapor attack. Either ethylene oxide

gas [106] or radiation is the suggested method for sterilization. Both have been shown to

be effective with pDTEC. Both are used at low temperatures. Radiation can sometimes

cause degradation, but has not been found to do so significantly in pDTEC. However, no

experiments have been conducted as to the affects of ethylene oxide gas and radiation on

the L-lysine bridge and the peptide attached to the polymer surface. These techniques

must be evaluated for racemization[141] and degradation of the pendant molecules.

6.2.3 Storage

One final concern for the polymer/peptide construct is the result of short- and

long-term storage. The polymer itself showed no loss in molecular weight when stored at

Page 57: Peptide-modified polymer for endothelialization

45

0°C with dessication for more than one year. Sterilized polymer devices can be safely

stored at room temperature for one year, but storage over longer periods of time requires

dessication and cooling [106].

Page 58: Peptide-modified polymer for endothelialization

46

List of References 1. Mundth, E.D., Surgical aspects of coronary artery bypass, in Cardiovascular system dynamics, J. Baan, A. Noordergraaf,J. Raines, Editors. 1978, The MIT Press: Boston. p. 55-62.

2. Heart and stroke statistical update. 2001, American Heart Association.

3. Ratner, B.D., ed. Biomaterials science: An introduction to materials in medicine. 1996, Academic Press: New York. 287.

4. Hurst, J.W.,R.B. Logue, The heart, arteries, and veins: Introduction, in The heart, arteries, and veins, J. Willis Hurst, M.D,. R. Bruce Logue, M.D., Editors. 1970, McGraw-Hill: New York. p. 3-32.

5. Glover, J.L., Historical perspective on endothelial seeding, in Endothelial seeding in vascular surgery, M. Herring,J.L. Glover, Editors. 1987, Grune and Stratton: New York. p. 1-6.

6. Seifalian, A.M., H.J. Salacinski, G. Punshon, B. Krijgsman,G. Hamilton, A new technique for measuring the cell growth and metabolism of endothelial cells seeded on vascular prostheses. J Biomed Mater Res, 2001. 55: p. 637-644.

7. Back, M.A.R.W., Biomaterials in vascular surgery., in Human biomaterials applications, D.L. Wise, Editor. 1996, Humana Press: Totowa, New Jersey. p. 257-299.

8. Bowlin, G., S.P. Schmidt, S.E. Rittgers,K.J. Pawlowski., Endothelialization of vascular prostheses., in Biomaterials engineering and devices: Human applications, D.L. Wise, Editor. 2000, Humana Press: Totowa, New Jersey. p. 85-118.

9. Sonoda, H., K. Takamizawa, Y. Nakayama, H. Yasui,T. Matsuda, Small-diameter compliant arterial graft prosthesis: Design concept of coaxial double tubular graft and its fabrication. J Biomed Mater Res, 2001. 55: p. 266-276.

10. Graham, I.,K. Robinson, Homocysteine and vascular disease. 2000, the Netherlands: Kluwer Academic Publishers. 1-4.

11. Mates, R.E., Arterial macrocirculatory hemodynamics, in Biomechanics, D.J. Schneck, Editor. 1995, CRC Press: Hartford. p. 454-473.

12. Majno, G., T. Zand, J.J. Nunnari, M.C. Kowala,I. Joris, Intimal responses to shear stress, hypercholesterolemia, and hypertension: Studies in rat aorta, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 349-367.

Page 59: Peptide-modified polymer for endothelialization

47

13. Ross, R., Endothelial injury and artherosclerosis, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 371-384.

14. Curriden, S.A., T.J. Podor,D.J. Loskutoff, The fibrinolytic system of cultured endothelial cells, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 229-252.

15. Schneck, D.J., An outline of cardiovascular structure and function, in Physiological systems, R. Plonsey, Editor. 1995, CRC Press: Hartford. p. 3-14.

16. CancerWEB, On-line medical dictionary. 2001, CancerWEB.

17. Canfield, T.R.,P.B. Dobrin, Mechanics of blood vessels, in Biomechanics, D.J. Schneck, Editor. 1995, CRC Press: Hartford. p. 291-303.

18. Roeder, R., J. Wolfe, N. Lianakis, T. Hinson, L.A. Gedded,J. Obermiller, Compliance, elastic modulus, and burst pressure of small-intestine submucosa (sis), small-diameter vascular grafts. J Biomed Mater Res, 1999. 47: p. 65-70.

19. Dixit, P., D. Hern-Anderson, J. Ranieri,C.E. Schmidt, Vascular graft endothelialization: Comparative analysis of canine and human endothelial cell migration on natural biomaterials. J Biomed Mater Res, 2001. 56: p. 545-555.

20. Franke, W.W., P. Cowin, C. Grund, C. Kuhn,H.-P. Kapprell, The endothelial junction: The plaque and its components, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 147-166.

21. Gimbrone, M.A.,M.P. Bevilacqua, Vascular endothelium: Function modulation at the blood interface, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 255-273.

22. Simionescu, M., Receptor-mediated transcytosis of plasma molecules by vascular endothelium, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 69-104.

23. Madri, J.A., B.M. Pratt,J. Yannariello-Brown, Endothelial cell-extracellular matrix interactions: Matrix as a modulator of cell function, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 167-188.

24. Marcum, J.A.,R.D. Rosenberg, The biochemistry and physiology of anticoagulantly active heparin-like molecules, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 207-228.

Page 60: Peptide-modified polymer for endothelialization

48

25. Colman, R.W., V.J. Marder, E.W. Salzman,J. Hirsh, in Hemostasis and thrombosis: Basic principles and clinical practice, R.W. Colman, J. Hirsh, V.J. Marder,E.W. Salzman, Editors. 1994, J. B. Lippincott Co: Philadelphia. p. 3-18.

26. Barbucci, R., F. Venturi,A. Albanese, Different sulfation degree of hyaluronic acid as heparin-like molecule: Physico-chemical characterization and biological performance, in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 157-170.

27. Wright, T.N., D.K. Heinegard,V.C. Hascall, Proteoglycans: Structure and function, in Cell biology of extracellular matrix, E.D. Hay, Editor. 1991, Plenum Press: New York. p. 45-78.

28. Stern, D.M., D.A. Handley,P.P. Nawroth, Endothelium and the regulation of coagulation, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 275-306.

29. Heimark, R.L.,S.M. Schwartz, Endothelial morphogenesis, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 123-143.

30. Plow, E.F.,S.E. D'Souza, Interaction of adhesive proteins with the platelet receptor gpiib-iiia, in Atherosclerosis: Cellular interactions, growth factors, and lipids, P.C. Weber,A. Leaf, Editors. 1993, Raven Press: New York. p. 29-38.

31. Bourin, M.-C.,U. Lindahl, Glycosaminoglycans and the regulation of blood coagulation. Biochem J, 1993. 289: p. 313-330.

32. Kefalides, N.A., Response of blood vessel cells to viral infection, in Endothelial cell biology in health and disease, N. Simionescu,M. Simionescu, Editors. 1988, Plenum Press: New York. p. 431-449.

33. Gimbrone, M.A., N. Kume,M.I. Cybulsky, Vascular endothelial dysfunction and pathogenesis of atherosclerosis, in Atherosclerosis: Cellular interactions, growth factors, and lipids, P.C. Weber,A. Leaf, Editors. 1993, Raven Press: New York. p. 1-9.

34. Kannel, W.B., W.P. Castelli, T. Gordon,P.M. McNamara, Serum cholesterol, lipoproteins and the risk of coronary heart disease: The framingham study. Ann Intern Med, 1971. 74: p. 1-12.

35. Castelli, W.P., P.J. Gerrison,P.W.F. Wilson, Incidence of coronary heart disease and lipoprotein cholesterol levels: The framingham study. JAMA, 1986. 256: p. 2835-2838.

36. Agishi, T.,B. Gordon, Clinical imporvement in patients with coronary and peripheral vascular diseases treated by ldl-adsorption using the liposorber la-15 system, in Biomaterials engineering and devices: Human applications, D.L. Wise, Editor. 2000, Humana Press: Totowa, New Jersey. p. 189-200.

Page 61: Peptide-modified polymer for endothelialization

49

37. Edelman, E.R., Vessel size, antioxidant, and restenosis: Never too small, not too little, but often too late. Circulation, 1998. 97: p. 416-420.

38. Nabel, E.G., Gene therapy for cardiovascular disease. Circulation, 1995. 91: p. 541-548.

39. Puhl, H., G. Waeg,H. Esterbauer, Inhibition of ldl oxidation by vitamin e and other antioxidants, in Atherosclerosis: Cellular interactions, growth factors, and lipids, P.C. Weber,A. Leaf, Editors. 1993, Raven Press: New York. p. 277-285.

40. Riemersma, R.A., Antioxidant vitamins, atherosclerosis, and coronary heart disease, in Atherosclerosis: Cellular interactions, growth factors, and lipids, P.C. Weber,A. Leaf, Editors. 1993, Raven Press: New York. p. 287-292.

41. Bolz, A., C. Harder, M. Unverdorben,M. Schaldach, Development of a new hybrid coronary stent design with optimized biocompatible properties, in Biomaterials engineering and devices: Human applications, D.L. Wise, Editor. 2000, Humana Press: Totowa, New Jersey. p. 201-222.

42. Mintz, G.S.,R. Hoffman, In-stent restenosis: The washington hospital center experience. Am J Cardiol, 1997. 81: p. 7E-13E.

43. Berglund, H.,H. Luo, Highly localized arterial remodeling in patients with coronary atherosclerosis: An intravascular ultrasound study. Circulation, 1997. 96: p. 1470-1476.

44. Snowhill, P.B., J.L. Nosher,F.H. Silver, The importance of radial forces in vascular stent design, in Biomaterials engineering and devices: Human applications, D.L. Wise, Editor. 2000, Humana Press: Totowa, New Jersey. p. 145-160.

45. Bankhead, C., Coronary stents. AAMI, 1996. 3: p. 8-10.

46. Schatz, R.A., View of vascular stents. Circulation, 1989. 79: p. 445-457.

47. Stent procedure. 2001, American Heart Association.

48. Topol, E.J.,P.W. Serruys, Frontiers in interventional cardiology. Circulation, 1998. 98: p. 1802-1820.

49. Rundback, J.H., R. Leonardo,G.N. Rozenblit, Peripheral vascular stents, in Stent-grafts: Current clinical practices, B.L. Dolmatch,U. Blum, Editors. 2000, Thieme: New York. p. 1-21.

50. Chandran, K.B.,S.W. Shalaby, Soft tissue replacements, in Biomaterials, J.B. Park, Editor. 1995, CRC Press: Hartford. p. 648-665.

51. Ku, D.N.,R.C. Allen, Vascular grafts, in Prostheses and artificial organs, P.M. Galletti, Editor. 1995, CRC Press: Hartford. p. 1871-1878.

Page 62: Peptide-modified polymer for endothelialization

50

52. Motwani, J.G.,E.J. Topol, Aortocoronary saphenous vein graft disease: Pathogenesis, predisposition, and prevention. Circulation, 1998. 97: p. 916-931.

53. Eskin, S.G., L.T. Navarro, J.L. Zamora, C.L. Ives, J.M. Anderson, D.G. Weilbaecher, Z.R. Gao,G.P. Noon, Preliminary studies on autologous versus homologous endothelial cell seeding of artiovenous grafts, in Endothelial seeding in vascular surgery, M. Herring,J.L. Glover, Editors. 1987, Grune and Stratton: New York. p. 155-163.

54. Hellener, G., D. Cohn,G. Marom, Dynamic compliance measurements of synthetic biodegradable grafts and natural blood vessels, in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 175-187.

55. Noh, I., S.L. Goodman,J.A. Hubbell, Chemical modification and photograft polymerization upon expanded poly(tetrafluoroethylene), in Polymers for tissue engineering, M.S. Shoichet,J.A. Hubbell, Editors. 1998, VSP: Utrecht. p. 145-164.

56. Lyman, D.J., J. Murray-Wijelath, E. Ambrad-Chalela,E.S. Wijelath, Vascular graft healing. II. FTIR analysis of polyester graft samples from implanted bi-grafts. J Biomed Mater Res (Appl Biomater), 2001. 58: p. 221-237.

57. Lynch, T.G.,R.W.I. Hobson, An overview of femoropopliteal and fermoral infrapopliteal reconstruction, in Vascular grafting: Clinical applications and techniques, C.B. Wright, R.W.I. Hobson, L.F. Hiratzka,T.G. Lynch, Editors. 1983, John Wright: Boston. p. 89-116.

58. Pulpapura, S.,J. Kohn, Trends in the development of bioresorbable polymers for medical applications. J Biomed Mater Res, 1992. 6: p. 216-250.

59. Chu, C.-C., Biodegradable polymeric biomaterials: An overview, in Biomaterials, J.B. Park, Editor. 1995, CRC Press: Hartford. p. 611-626.

60. Lavik, E.B., J.S. Hrkach, N. Lotan, R. Nazarov,R. Langer, A simple synthetic route to the formation of a block copolymer of poly(lactic-co-glycolic acid) and polylysine for the fabrication of functionalized, degradable structures for biomedical applications. J Biomed Mater Res (Appl Biomater), 2001. 58: p. 291-294.

61. Daniels, A.U., M.S. Taylor, K.P. Andriano,J. Heller, Toxicity of absorbed polymers proposed for fracture fixation devices. Proc Orthop Res Soc, 1992. 17: p. 88.

62. Marconi, W., A. Piozzi,D. Romoli, Immobilization of plasminogen onto polyurethane surfaces, in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 231-237.

63. Skarja, G.A.,K.A. Woodhouse, Synthesis and characterization of degradable polyurethane elastomers containing an amino acid-based chain extender, in Polymers for tissue engineering, M.S. Shoichet,J.A. Hubbell, Editors. 1998, VSP: Utrecht. p. 73-97.

Page 63: Peptide-modified polymer for endothelialization

51

64. Sun, X., H. Sheardown, P. Tengvall,J.L. Brash, Peptide modified gold-coated polyurethanes as thrombin scavenging surfaces. J Biomed Mater Res, 2000. 49: p. 66-78.

65. Tanzi, M.C., P. Petrini, M. Moscatelli,G. Garlaschelli, Poly(ether-amidoamines) verses polyamidoamines: Chemico-physical characterization and cytotoxicity, in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 143-156.

66. Cohn, D.,T. Stern, Multi-step surface tailoring of poly(ethylene terephthalate), in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 307-314.

67. Suggs, L.J., E.Y. Kao, L.L. Palombo, R.S. Krishnan, M.S. Widmer,A.G. Mikos, Preparation and characterization of poly(propylene fumarate-co-ethylene glycol) hydrogels, in Polymers for tissue engineering, M.S. Shoichet,J.A. Hubbell, Editors. 1998, VSP: Utrecht. p. 99-112.

68. Edwards, W.S.,J.S. Tapp, Surgery, 1955. 38: p. 61.

69. Hersh, L.S., V.L. Gott,F. Najjar. Thermal and ionic methods of heparinizing small-diameter dacron grafts. in Biomedical Materials Symposium: Biomaterials for Skeletal and Cardiovascular Applications. 1971. Atlantic City, NJ: John Wiley & Sons.

70. Grode, G.A., R.D. Falb,J.P. Crowley. Biocompatible materials for use in the vascular system. in Biomedical Materials Symposium: Biomaterials for Skeletal and Cardiovascular Applications. 1971. Atlantic City, NJ: John Wiley & Sons.

71. Kempczinski, R., E.C. Douville, G. Ramalanjaona, J.D. Ogle,E.B. Silberstein, Endothelial cell seeding on a fibronectin-coated substrate, in Endothelial seeding in vascular surgery, M. Herring,J.L. Glover, Editors. 1987, Grune and Stratton: New York. p. 57-77.

72. Thelen, H., H. Bienert, R. Engisch, H. Thissen, D. Klee, H. Hocker, K. Hoffmeister, J.V. Dahl,P. Hanrath, Surface modified bioresorbable polymers for blood contact: Study of haemocompatibility, in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 145-158.

73. Baumann, H., R. Keller,U. Baumann, Development of microvascular artificial platelet inert blood vessels, basic principles and animal experiments, in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 159-174.

74. Barra, S.L.,C. Boisson-Vidal, Blood-compatible bioactive polymers, in Biomaterials engineering and devices: Human applications, D.L. Wise, Editor. 2000, Humana Press: Totowa, New Jersey. p. 25-47.

75. Lee, H.B., S.S. Kim,G. Khang, Polymeric biomaterials, in Biomaterials, J.B. Park, Editor. 1995, CRC Press: Hartford. p. 581-597.

Page 64: Peptide-modified polymer for endothelialization

52

76. Baldwin, S.P.,W.M. Saltzman, Materials for protein delivery in tissue engineering. Advanced Drug Delivery Reviews, 1998. 33: p. 71-86.

77. Laemmel, E., J. Penhoat, R. Warocquier-Clerout,M. Sigot-Luizard, Heparin immobilized on proteins usable for arterial prosthesis coating: Growth inhibition of smooth-muscle cells. J Biomed Mater Res, 1998. 39: p. 446-452.

78. Baumann, H.,B. Huppertz, The influence of functional groups on semisynthetic heparin-like molecules on binding to vascular endothelial growth factor (vegf 165), in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 131-144.

79. Greisler, H.P., Regulation of vascular graft healing by induction of tissue incorporation, in Human biomaterials applications, D.L. Wise, Editor. 1996, Humana Press: Totowa, New Jersey. p. 393-413.

80. Yamada, K.M., Fibronectin and other cell interactive glycoproteins, in Cell biology of extracellular matrix, E.D. Hay, Editor. 1991, Plenum Press: New York. p. 111-146.

81. Palder, S., R. Kirkman,A. Whittemore, Vascular access for hemodialysis. Ann Surg, 1983. 202: p. 235-239.

82. Yoneyama, T., K. Sugihara, K. Ishihara, Y. Iwasaki,N. Nakabayashi, The vascular prosthesis without pseudointima prepared by antothrombogenic phospholipid polymer. Biomaterials, 2001. Article in Press.

83. Shakesheff, K.M., S.M. Cannizzaro,R. Langer, Creating biomimetic micro-environments with synthetic polymer-peptide hybrid molecules, in Polymers for tissue engineering, M.S. Shoichet,J.A. Hubbell, Editors. 1998, VSP: Utrecht. p. 113-124.

84. Rezania, A.,K.E. Healy, Biomimetic peptide surfaces that regulate adhesion, spreading, cytoskeletal organization, and mineralization of the matrix deposited by osteoblast-like cells. Biotechnol Prog, 1999. 15: p. 19-32.

85. Kim, B.S., J. Nikolovski, J. Bonadio, E. Smiley,D.J. Mooney, Engineered smooth muscle tissues: Regulating cell phenotype with the scaffold. Experimental Cell Research, 1999. 251: p. 318-328.

86. Besselink, G.A.J., T. Beugeling,A. Banties, Affinity-binding of cells to sephadex derivatized with peptides and proteins, in Frontiers in biomedical polymer applications, R.M. Ottenbrite, Editor. 1998, Technomic: Lancaster. p. 201-213.

87. Hynes, R.O., Integrins: Versatility, modulation, and signaling in cell adhesion. Cell, 1992. 69: p. 11-25.

Page 65: Peptide-modified polymer for endothelialization

53

88. Chinn, J.A., J.A. Sauter, R.E. Phillips, W.J. Kao, J.M. Anderson, S.R. Hanson,T.R. Ashton, Blood and tissue compatibility of modified polyester: Thrombosis, inflammation, and healing. J Biomed Mater Res, 1998. 39: p. 130-140.

89. Ruoslahti, E.,M.D. Pierschbacher, New perspectives in cell adhesion: RGD and integrins. Science, 1987. 238: p. 491-497.

90. Ruoslahti, E., Integrins as receptors for extracellular matrix, in Cell biology of extracellular matrix, E.D. Hay, Editor. 1991, Plenum Press: New York. p. 343-363.

91. Plow, E.F., M.H. Ginsberg,G.A. Marguerie, in Biochemistry of platelets, D.R. Phillips,M.A. Shuman, Editors. 1986, Academic: Orlando. p. 225-226.

92. Humphries, M.J., A. Komoriya, S.K. Akiyama, K. Olden,K.M. Yamada, Identification of two distinct regions of the type III connecting segment of human plasma fibronectin that promotes cell type-specific adhesion. J Biol Chem, 1987. 262: p. 6886-6892.

93. Barrera, D.A., E. Zylstra,P.T. Lansbury, Synthesis and RGD peptide modification of a new biodegradable copolymer: Poly(lactic acid-co-lysine). J Am Chem Soc, 1993. 115: p. 11-20.

94. Holland, J., L.S. Hersh, M. Bryhan, E. Onyiriuka,L. Ziegler, Culture of human vascular endothelial cells on an RGD-containing synthetic peptide attached to a starch-coated polystyrene surface: Comparison with fibronectin-coated tissue grade polystyrene. Biomaterials, 1996. 17(22): p. 2147-2156.

95. Cardin, A.D.,H.J.R. Weintraub, Molecule modeling of protein-glycosaminoglycan interactions. Arteriosclerosis, 1989. 9: p. 21-32.

96. Verrecchio, A., M.W. Germann, B.P. Schick, B. Kung, T. Twardowski,J.D.S. Antonio, Design of peptides with high affinity for heparin and endothelial cell proteoglycans. J Biol Chem, 2000. 275(11): p. 7701-7707.

97. Herring, M., A. Gardner,J.L. Glover, A single-staged technique for seeding vascular grafts with autogenous endothelium. Surgery, 1978. 84: p. 498-504.

98. Graham, L., W. Burkel, J. Ford,e. al, Expanded polytetrafluoroethylene vascular prostheses seeded with enzymatically derived and cultured canine endothelial cells. Surgery, 1982. 91: p. 550-559.

99. Kempczinski, R., J. Rosenman,W. Pearce, Endothelial cell seeding of a new PTFE vascular prosthesis. J of Vasc Surg, 1985. 2: p. 424-429.

100. Schmidt, S.P., T. Hunter,L. Falkow, Effects of antiplatelet agents in combination with endothelial cell seeding on small-diameter dacron vascular graft performance in the canine carotid artery model. J Vasc Surg, 1985. 2: p. 898-906.

Page 66: Peptide-modified polymer for endothelialization

54

101. Graham, L.M., W. Burkel, J. Ford, D. Vinter, R.H. Kahn,J. Stanley, Expaned polytetrafluoroethylene vascular prostheses seeded with enzymatically derived and cultured canine endothelial cells. Surgery, 1982. 91: p. 550-559.

102. Kohn, J., Desaminotyrosyl-tyrosine alkyl esters: New diphenolic monomers for the design of tyrosine-derived pseudopoly(amino acids), in Polymeric drugs and drug delivery systems, R.L. Dunn,R.M. Ottenbrite, Editors. 1991. p. 157-167.

103. Engelberg, I.,J. Kohn, Physico-mechanical properties of degradable polymers used in medical applications: A comparative study. Biomaterials, 1991. 12: p. 292-304.

104. Kohn, J.,R. Langer. A new approach to the development of bioerodible polymers for controlled release applications employing naturally occurring amino acids. in ACS Division of Polymeric Materials, Science and Engineering. 1984. Washington D. C.: American Chemical Society.

105. Brode, G.L., J. Kohn,J.E. Kemnitzer, Biosmart tyrosine polycarbonates. Polym Preprints, Am Chem Soc, Div Polym Chem, 1998. 39(2): p. 230-231.

106. Hooper, K., N.D. Macon,J. Kohn, Comparative histological evaluation of new tyrosine-derived polymers and poly(l-lactic acid) as a function of polymer degradation. J Biomed Mater Res, 1998. 41: p. 443-454.

107. Tschopp, J.F., J.O. Tolley, T. Malaney, C. Mazur, R. Bellantoni, S. Abramson, J. Kohn, B. Wong,J.E. Kemnitzer, In vivo resorption profile of matrices from tyrosine-derived polycarbonates with variation in free carboxylate content. Society for Biomaterials: Sixth World Biomaterials Congress Transactions, 2000.

108. Silver, F.H., M. Marks, Y.P. Kato, C. Li, S. Pulpapura,J. Kohn, Tissue compatibility of tyrosine derived polycarbonates and polyiminocarbonates: An initial evaluation. J Long-Term Effects Med Implants, 1992. 1: p. 329-346.

109. Ertel, S.,J. Kohn, Evaluation of a series of tyrosine-derived polycarbonates as degradable biomaterials. J Biomed Mat Res, 1994. 28: p. 919-930.

110. Brocchini, S., K. James, V. Tangpasuthadol,J. Kohn, A combinatorial approach for polymer design. J Am Chem Soc, 1997. 119: p. 4553-4554.

111. Choueka, J., J. Charvet, K. Koval, H. Alexander, K. James, K. Hooper,J. Kohn, Canine bone response to tyrosine-derived polycarbonates and poly(l-lactic acid). J Biomed Mat Res, 1996. 31: p. 35-41.

112. Tangpasuthadol, V., A. Shefer, K. Hooper,J. Kohn, Thermal properties and physical ageing behaviour of tyrosine-derived polycarbonates. Biomaterials, 1996. 17(4): p. 463-468.

113. Berthiaume, F.,M.L. Yarmush, Tissue engineering, in Biotechnology, M.L. Yarmush, Editor. 1995, CRC Press: Hartford. p. 1556-1566.

Page 67: Peptide-modified polymer for endothelialization

55

114. Herring, M., Endothelial seeding of blood flow surfaces, in Vascular grafting: Clinical applications and techniques, C.B. Wright, R.W.I. Hobson, L.F. Hiratzka,T.G. Lynch, Editors. 1983, John Wright: Boston. p. 275-314.

115. Verrecchio, A., Discovery of peptides with high affinities for heparin and endothelial cell proteoglycans, in Biomedical Chemistry. MS Thesis 1999, Thomas Jefferson University: Philadelphia. p. 41.

116. Ferran, D.S., M. Sobel,R.B. Harris, Biochem J, 1992. 31: p. 5010-5016.

117. Drumheller, P.D.,J.A. Hubbell, Surface immobilization of adhesion ligands for investigations of cell-substrate interactions, in Tissue engineering, J.A. Hubbell,B.O. Palsson, Editors. 1995, CRC Press: Hartford. p. 1583-1596.

118. Aguado, M.T.,P.H. Lambert, Controlled-release vaccines-biodegradable polylactide/polyglycolide PL/PG) microspheres as antigen vehicles. Immunology, 1992. 184: p. 113.

119. Germann, M. 2001, Personal Communication "1-D NMR Integration".

120. Chalupowicz, D.G., Z.A. Chowdhury, T.L. Bach, C. Barsigian,J. Martinez., Fibrin II induces endothelial cell capillary tube formation. J Cell Biol, 1995. 130(1): p. 207-215.

121. Maciag, T., J. Cerundolo, S. Ilsley,R. Forand, An endothelial cell growth factor from bovine hypothalamus: Identification and partial characterization. Pro Nat Acad Sci USA, 1979. 76(11): p. 5674-5678.

122. San Antonio, J. 2000, Personal Communication.

123. Lotz, M.M., C.A. Burdsal, H.P. Erickson,D.R. McClay, Cell adhesion to fibronectin and tenascin: Quantitative measurements of initial binding and subsequent strengthening response. J Cell Biol, 1989. 109: p. 1795-1805.

124. Westergren-Thorsson, G., P.O. Onnervik, L.A. Fransson,A. Malmstrom, Proliferation of cultured fibroblasts is inhibited by L-iduronate-containing glycosaminoglycans. J Cell Physiol, 1991. 147: p. 523-530.

125. Kottke-Marchant, K., A.A. Veenstra,R.E. Marchant, Human endothelial cell growth and coagulation function varies with respect to interfacial properties of polymer substrates. J Biomed Mater Res, 1996. 30: p. 209-220.

126. Jarrell, B.E., S.K. Williams, R.A. Carabasi,F.A. Hubbard, Immediate vascular graft monolayers using microvessel endothelial cells, in Endothelial seeding in vascular surgery, M. Herring,J.L. Glover, Editors. 1987, Grune and Stratton, Inc: Orlando. p. 37-55.

Page 68: Peptide-modified polymer for endothelialization

56

127. Freed, L., J.C. Marquis, A. Nohria, J. Emmanual,A.G. Mikos, Neocartilage formation in vitro and in vivo using cells cultured on synthetic biodegradable polymers. J Biomed Mater Res, 1993. 27: p. 11-23.

128. Mikos, A.G., Y. Bao, L.G. Cima, D.E. Ingber, J.P. Vacanti,R. Langer, Preparation of poly(glycolic acid) bonded fiber structures for cell attachment and transplantation. J Biomed Mater Res, 1993. 27: p. 183-189.

129. Mooney, D.J., C.L. Mazzoni, C. Breuer, K. McNamara, D. Hern, J.P. Vacanti,R. Langer, Stabilized polyglycolic acid fiber-based tubes for tissue engineering. Biomaterials, 1996. 17: p. 115-124.

130. Mikos, A.G., G. Sarakinos, S.M. Leite, J.P. Vacanti,R. Langer, Laminated three-dimensional foams for use in tissue engineering. Biomaterials, 1993. 14: p. 323-330.

131. Mikos, A.G., A.J. Thorsen, L.A. Czerwonka, Y. Bao, R. Langer, D.N. Winslow,J.P. Vacanti, Preparation of poly(l-lactic acid) foams. Polymer, 1994. 35: p. 1068-1077.

132. Whang, K., C.H. Thomas, K.E. Healy,G.A. Nuber, Novel methods to fabricate bioabsorbable scaffolds. Polymer, 1995. 36: p. 837-842.

133. Mooney, D.J., D.F. Baldwin, N.P. Suh, J.P. Vacanti,R. Langer, Novel approach to fabricating porous sponges of poly(D,L-lactic-co-glycolic acid) without the use of organic solvents. Biomaterials, 1996. 17: p. 1417-1422.

134. Harris, L.D., B.S. Kim,D.J. Mooney, Open pore biodegradable matricies formed with gas foaming. J Biomed Mater Res, 1998. 42(396-402).

135. Lo, H., S. Kadiyala, S.E. Guggino,K.W. Leong, Poly(l-lactic acid) foams with cell seeding and controlled-release capacity. J Biomed Mater Res, 1996. 30: p. 475-484.

136. Schugens, C.H., V. Maguet, C.H. Grandfils, R. Jerome,P.H. Teyssie, Polylactic acid macroporous biodegradable implants for cell transplatation 2. Preparation of polylactide foams by liquid-liquid phase separation. J Biomed Mater Res, 1996. 30: p. 449-461.

137. Nam, Y.S.,T.G. Park, Porous biodegradable polymeric scaffolds prepared by thermally induced phase separation. J Biomed Mater Res, 1999. 47: p. 8-17.

138. Nam, Y.S., J.J. Yoon,T.G. Park, A novel fabrication method of macroporous bioderadable polymer scaffolds using foaming salt as a progen additive. J Biomed Mater Res (Appl Biomater), 2000. 53: p. 1-7.

139. Callister, W.D., Materials science and engineering: An introduction. 5th ed. 2000, New York: John Wiley and Sons, Inc.

Page 69: Peptide-modified polymer for endothelialization

57

140. Block, S.S., ed. Disinfection, sterilization and preservation. 2 ed. 1977, Rea and Febiger: Philadelphia.

141. Morrison, R.T.,R.N. Boyd, Organic chemisrty. 6th ed. 1992, Englewood Cliffs, NJ: Prentice Hall.

142. Kleinman, L.H., D.C. Mullen,R.J. Flemma, Aortocoronary grafting techniques, in Vascular grafting: Clinical applications and techniques, C.B. Wright, R.W.I. Hobson, L.F. Hiratzka,T.G. Lynch, Editors. 1983, John Wright: Boston. p. 198-206.

143. Fox, S.I., Human physiology. 1990.

Page 70: Peptide-modified polymer for endothelialization

58

Appendix A- Tables

Table 1- Patency rates for ASV and IMA grafting of the coronary artery of varying diameters, from [142]

Patency rates >1 year (%) Coronary Artery Size

(mm) ASV IMA 3.0 100 100 2.5 97 94 2.0 91 96 1.5 87 93 1.0 56 91

Page 71: Peptide-modified polymer for endothelialization

59

Table 2- Patency data for representative vascular grafts, adapted from Ku [51].

Graft Diameter (mm) Conduit 5-year patency (%)

Dacron 90 Aortobifemoral >6 ePTFE 90 Dacron 80 Femorofemoral <6 ePTFE 80 ASV 70

Dacron 50 Femoropopliteal <4 ePTFE 40

Page 72: Peptide-modified polymer for endothelialization

60

Table 3-Improving hemocompatibility of artificial biomaterials [73].

Modifications to Achieve Hemocompatibility

Bulk Material Surface Modification • Crystalline, amorphous, conformation

• Functional Groups

• Hydrophilic/hydrophobic balance

• Charge

• Block copolymers

• Hydrogel

• Charge

• Hydrogel

• Topology

• Cell seeding endothelial cells

• Immobilization of biologically active or passive biopolymers or low molecular weight substances

Page 73: Peptide-modified polymer for endothelialization

61

Table 4- Properties of native tissue and several synthetic polymers.

Material Tensile Modulus MPa (ksi)

Tensile Strength MPa (ksi)

Thoracic Tissue - 3.8 (0.6) DTE Carbonate 1600-2000 (232-290) 67 (9.7)

Reinforced PTFE - 330 (47.9) Dacron 2800-4100 (406-595) 48-72 (7.0-10.4)

PLA 2700 (392) 50 (7.3)

Page 74: Peptide-modified polymer for endothelialization

62

Table 5- NMR peak integration values for p(DTEC-c-X%DTC), X=0, 13, and 35, peaks.

Degree of Functionality Peak (ppm)

0% 13% 35%

α (4.8) 100.0 100.0 100.0

β (3.05) 197.7 185.4 163.2

γ (6.0) 96.4 99.7 98.2

δ (2.9) 202.5 198.0 201.4

ε (2.4) 198.7 205.6 204.3

Shoulder (3.1) - 12.6 35.3

R2 (4.1) 202.0 172.1 126.2

R3 (1.3) 305.0 269.2 201.2

Page 75: Peptide-modified polymer for endothelialization

63

Table 6- Integration values corresponding to polyalanine peaks for attachment to p(DTEC-c-X%DTC), X=0, 23, 50.

Integration values Peak location (ppm) 0% 23% 50%

4.0 0.03 0.32 0.64

3.5 0.04 0.34 0.72

3.2 0.35 0.59 0.94

1.9 0.96 2.0 3.08

1.7 1.07 2.14 3.51

Table 7- The corresponding attachment percentages have been calculated.

Attachment of polyalanine Peak location

(ppm) 23% 50%

4.0 29 % 61 %

3.5 30 % 69 %

3.2 24 % 59 %

1.9 104 % 212 %

1.7 107 % 244 %

Page 76: Peptide-modified polymer for endothelialization

64

Table 8- Cell attachment strength as assessed using a centrifugation assay. Note that the poly(DTE carbonate) shows no significant improvement in attachment over the negative control, but attachment increases with additional functional groups.

Cells attached (%) at g force Material 100 g 300 g 600 g 900 g

MECL2 0 0 0 0

0% DT < 5 0 0 0

13% DT 10 5 0 0

23% DT 15 12 10 2

35% DT 35 30 20 3

50% DT 40 35 25 6

FIBRONECTIN 80 60 30 5

Page 77: Peptide-modified polymer for endothelialization

65

Table 9- Comparative cell growth data at 7 days for various experimental and common vascular implant materials.

Material Average cells per slide

X=0 4950 X=13 5500

poly(DTEC-co-

X%DTC) X=35 12100 PLA 9350 PET 6050 PVC 0 Blank 18700

Page 78: Peptide-modified polymer for endothelialization

66

Appendix B- Figures

Figure 1- Cross-Section of a typical artery and vein [143].

Page 79: Peptide-modified polymer for endothelialization

67

Figure 2- a) Molecular model of poly(DTEC) repeat unit showing bond angles and bond lengths. b) Repeat unit of poly(DTE-co-X%DT carbonate). When R = CH2CH3 the repeat unit is desaminotyrosil-tyrosine-ethyl ester (DTE); when R = H the free unit is desaminotyrosil-tyrosine (DT). X is the molar fraction of repeat units with pendant acid groups. (Note: H’s are labeled α, β, γ, δ, ε, for identification and quantification by NMR spectra. The hydrogens in the pendant R groups are labeled by proximity to the ester group. Thus, the hydrogen on the free acid is R1, while the hydrogens in the ethyl ester are labeled R2 and R3.) c) Repeat unit of poly(DTEC-co-X%DTC) and L-lysine ethyl ester. Arrows indicate both possible attachment locations.

O

Carbon Oxygen Hydrogen Nitrogen

c)

b)

a)

α-terminus ε-terminus

Page 80: Peptide-modified polymer for endothelialization

68

Figure 3- 2D TOCSY of poly(DTEC) in CDCl3. Peaks are labeled as detailed in Figure 2. Solid lines indicate three different correlation systems with interactions between neighboring hydrogen.

Page 81: Peptide-modified polymer for endothelialization

69

Figure 4- 1D NMR spectra of a) P(DTEC-c-0%DTC) and b)P(DTEC-c-13%DTC) with peak integration values. The peak labels correspond to hydrogen as marked in Figure 2.

Page 82: Peptide-modified polymer for endothelialization

70

Figure 5- 1-D and 2-D COSY NMR spectrum of L-lysine ethyl ester. Solid lines indicate correlation systems. Below is the corresponding molecule.

Page 83: Peptide-modified polymer for endothelialization

71

Figure 6- TOCSY spectrum of P(DTEC-c-13%DTC) with attached L-lysine ethyl ester. The original correlation systems from the poly(DTE-co-X%DT carbonate) system are indicated by lines. The solid lines indicate correlation systems from the L-lysine. Note that the system has ( ) ( ) two related correlation systems, corresponding to α– and ε–amine attachment of the lysine.

Page 84: Peptide-modified polymer for endothelialization

72

Figure 7- NMR spectrum of P(DTEC-c-13%DTC) with attached L-lysine ethyl ester. Integration values are below peaks. The ratio of the α poly(DTE carbonate)peak, which should be constant at a concentration of 1 across chemistries, to the lysine spectral group at ca. 4.38 shows an approximately 11.5±0.5% lysine attachment.

polymer

Page 85: Peptide-modified polymer for endothelialization

73

Figure 8- TOCSY spectrum of p(DTEC-c-13%DTC) with attached L-lysine ethyl ester and polylysine. Note that there are no additional peaks, only intensification of existing peaks.

Page 86: Peptide-modified polymer for endothelialization

74

Figure 9- a) 1D NMR spectrum of p(DTEC-c-50%DTC). Peak integration values are shown below select peaks. The integration value for the peak at 4.7 ppm corresponds to the α peak in the polymer. Characteristic polyalanine peaks are also indicated. b) Structure of polyalanine.

α polymer Polyalanine Polyalanine

a)

b)

Page 87: Peptide-modified polymer for endothelialization

75

Figure 10- 1D NMR spectrum of p(DTEC-c-35%DTC) with peptide (ARKKAAKA)4 attached via a lysine bridge.

NH peaks

Methyl group peaks

Page 88: Peptide-modified polymer for endothelialization

76

0%

20%

40%

60%

80%

100%

0 300 600 900 1200

Force (g)

Cel

ls A

ttach

edFibrin50%35%23%13%0%MeCl2

Figure 11- Attachment data from Table 8 for functional polymers, p(DTEC-c-X%TDC) X=0, 13, 23, 35, and 50% and positive (FN) and negative (MeCl2) controls. Attachment increases with additional functional groups.

Page 89: Peptide-modified polymer for endothelialization

77

0%

10%

20%

30%

40%

50%

60%

70%

0% 10% 20% 30% 40% 50%

Polymer Functionalization (%)

Cel

ls A

ttach

ed (%

)300 g

600 g

FN 300

FN 600

Bl 300

Bl 600

Figure 12- Attachment data for functional polymers and positive and negative controls. P(DTEC-c-X%TDC) X=0, 13, 23, 35, and 50% are shown at both the low (300 g) and medium (600 g) force. FN (positive control) and Blank (negative control) are drawn as straight lines as they are not subject to the same functionalization. The lines for the FN and the Blank indicate the cell attachment at the specified force.

Page 90: Peptide-modified polymer for endothelialization

78

0%

3%

5%

8%

10%

13%

15%

0 300 600 900 1200

Force (g)

Cel

ls a

ttach

ed13%

13wL

PolyL

32mer PEP

Figure 13- Cell attachment data for p(DTEC-c-13%DTC) alone (13%), with attached lysine (13wL), attached polylysine (PolyL), and attached heparin-binding 32mer peptide (32mer PEP).

Page 91: Peptide-modified polymer for endothelialization

79

a) b)

c) d)

Figure 14- Optical microscope images of cell attachment and growth on 13% poly(DTE-co-X%DT carbonate) at a) 1 hour, b) 1 day, c) 3 days and d) 7 days. Note that at 1 hour, the cells show flattening indicative of attachment. The total number of cells increases through day 7. On day 7, there are regions of cell confluence.

200 µm

Page 92: Peptide-modified polymer for endothelialization

80

a) b)

c) d)

e) f)

g)

Figure 15- Optical microscope image of cell attachment at day 5 on a) poly(DTE carbonate), b) 13% poly(DTE-co-X%DT carbonate), c) 35% poly(DTE-co-X%DT carbonate), d) PLA, e) PET, f) PVC and g) glass.

200 µm

Page 93: Peptide-modified polymer for endothelialization

81

Figure 16- Optical microscope image of stained cells at the end of the growth assay, a) 12X shows areas of semi-confluence b) 23X shows spread cells indicating attachment c) 46X shows spread cells with visible nuclei and cells in the process of dividing.

100 µm 50 µm

200 µm a)

b) c)

Page 94: Peptide-modified polymer for endothelialization

82

Appendix C- Nomenclature Allografts- a graft from one species to another Angioplasty- a balloon angioplasty is a noninvasive procedure where a balloon-tipped catheter is introduced into a diseased blood vessel. As the balloon is inflated, the vessel opens further allowing for improved flow of blood. Atherosclerosis- a disease of the blood vessels characterized by thickening of the vessel wall and eventually occlusion of the vessel Autograft- a graft from one individual to another Compliance- structural property of a tube that expresses dimensional change in response to a change in intraluminal pressure Denudation- the act of stripping off, or removing the surface, in this case removal of ECs Endothelium/Endothelial cell lining- flat cells that line the innermost surfaces of blood and lymphatic vessels and the heart Homeostasis- a tendency to uniformity or stability in an organism by maintaining within narrow limits certain variables that are critical to life Hydrogel- polymer that can absorb water to 30% or more of its weight Integrin- cellular transmembrane proteins that act as receptors for adhesive extracellular matrix proteins such as fibronectin. The tripeptide RGD is the sequence recognized by many integrins. Neointima- newly formed intimal surface Neointimal hyperplasia- growth of a new intimal surface formed by fibroblasts or smooth muscle cells Occlusion- closed vein or artery Patency- the time a repair remains functional Platelet- one of the formed elements of blood responsible for blood coagulation

Page 95: Peptide-modified polymer for endothelialization

83

Stenosis- tissue ingrowth into vessel causing a narrow lumen and reduction of blood flow Thrombosis- the formation of an aggregation of blood factors, primarily platelets and fibrin with entrapment of cellular elements, frequently causing vascular obstruction. VE Cadherin- integral membrane proteins involved in calcium dependent cell adhesion. Formed of a 600 amino acid extracellular domain, containing 4 repeats believed to contain the Ca binding sites, a transmembrane domain and a 150 amino acid intracellular domain.

Page 96: Peptide-modified polymer for endothelialization

84

Appendix D- Abbreviations ASV autologous saphenous vein

COSY correlated spectroscopy

CD circular diochroism

CVD cardiovascular disease

DCC 1,3-Dicyclohexylcarbodiimide

EC endothelial cell

ePTFE expanded poly(tetrafluoroethylene)

FDA Federal Drug Administration

FMOC 9-fluorenylmethoxycarbonyl

FN fibronectin

GAG glycosaminoglycan

HUVEC human umbilical vein endothelial cell

IMA internal mammary artery

MeCl2 methylene dichloride

NMR nuclear magnetic resonance

PBS phosphate buffered saline

PET poly(terephthalate)

PG proteoglycan

PLA poly(L-lactic acid)

PTCA percutaneous transluminal coronary angioplasty

PVC poly(vinyl chloride)

TOCSY total correlated spectroscopy