198
Duquesne University Duquesne Scholarship Collection Electronic eses and Dissertations Summer 8-11-2018 Self-Assembling Bioaffinity Hydrogels for Localized Antibody Delivery Wen Liu Follow this and additional works at: hps://dsc.duq.edu/etd Part of the Pharmaceutics and Drug Design Commons is One-year Embargo is brought to you for free and open access by Duquesne Scholarship Collection. It has been accepted for inclusion in Electronic eses and Dissertations by an authorized administrator of Duquesne Scholarship Collection. For more information, please contact [email protected]. Recommended Citation Liu, W. (2018). Self-Assembling Bioaffinity Hydrogels for Localized Antibody Delivery (Doctoral dissertation, Duquesne University). Retrieved from hps://dsc.duq.edu/etd/1469

Self-Assembling Bioaffinity Hydrogels for Localized

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Self-Assembling Bioaffinity Hydrogels for Localized

Duquesne UniversityDuquesne Scholarship Collection

Electronic Theses and Dissertations

Summer 8-11-2018

Self-Assembling Bioaffinity Hydrogels forLocalized Antibody DeliveryWen Liu

Follow this and additional works at: https://dsc.duq.edu/etd

Part of the Pharmaceutics and Drug Design Commons

This One-year Embargo is brought to you for free and open access by Duquesne Scholarship Collection. It has been accepted for inclusion in ElectronicTheses and Dissertations by an authorized administrator of Duquesne Scholarship Collection. For more information, please [email protected].

Recommended CitationLiu, W. (2018). Self-Assembling Bioaffinity Hydrogels for Localized Antibody Delivery (Doctoral dissertation, Duquesne University).Retrieved from https://dsc.duq.edu/etd/1469

Page 2: Self-Assembling Bioaffinity Hydrogels for Localized

SELF-ASSEMBLING BIOAFFINITY HYDROGELS FOR LOCALIZED ANTIBODY

DELIVERY

A Dissertation

Submitted to the Graduate School of Pharmaceutical Sciences

Duquesne University

In partial fulfillment of the requirements for

the degree of Doctor of Philosophy

By

Wen Liu

August 2018

Page 3: Self-Assembling Bioaffinity Hydrogels for Localized

Copyright by

Wen Liu

2018

Page 4: Self-Assembling Bioaffinity Hydrogels for Localized

iii

SELF-ASSEMBLING BIOAFFINITY HYDROGELS FOR LOCALIZED ANTIBODY

DELIVERY

By

Wen Liu

Approved June 6, 2018

________________________________

Wilson S. Meng, Ph.D.

Associate Professor of Pharmaceutical

Sciences

Duquesne University, Pittsburgh, PA

(Committee Chair)

________________________________

Ellen S. Gawalt, Ph.D

Department Chair and Professor of

Chemistry and

Biochemistry

Duquesne University, Pittsburgh, PA

(Committee Member)

________________________________

James K. Drennen, III, Ph.D.

Associate Dean for Graduate Programs

and Research

Associate Professor of Pharmaceutical

Sciences

Duquesne University, Pittsburgh, PA

(Committee Member)

________________________________

Patrick Flaherty

Associate Professor of Medicinal

Chemistry

Duquesne University, Pittsburgh, PA

(Committee Member)

________________________________

Jelena M. Janjic, Ph.D

Associate Professor of Pharmaceutical

Sciences

Duquesne University, Pittsburgh, PA

(Committee Member)

________________________________

Peter Wildfong, Ph.D

Associate Professor of Pharmaceutical

Sciences

Duquesne University, Pittsburgh, PA

Page 5: Self-Assembling Bioaffinity Hydrogels for Localized

iv

ABSTRACT

SELF-ASSEMBLING BIOAFFINITY HYDROGELS FOR LOCALIZED ANTIBODY

DELIVERY

By

Wen Liu

August 2018

Dissertation supervised by Wilson S. Meng, Ph.D.

The scope of illnesses that are being treated using monoclonal antibodies (mAbs)

has been expanding from cancers to autoimmune diseases due to their high specificity to

their molecular target. Because many of the new mAbs are aimed to modulate the

immune system, an emerging concern is that patients with chronic illnesses are subjected

to immune-related Adverse Events (irAEs). Loss of immunological tolerance to self-

antigens or increased susceptibility to infections have been documented in some mAb

treatments. Localized sustained release of mAb at disease sites might mitigate the risk of

such toxicities. Antibody drugs are formulated typically for parenteral delivery because

of degradative mechanisms in gastrointestinal tract. The assumption has been that

systemic administration via the vascular compartment is the best or only option in

achieving therapeutic drug concentrations at diseased tissues. This paradigm is being

Page 6: Self-Assembling Bioaffinity Hydrogels for Localized

v

challenged as the benefit to risk ratio of exposing non-targeted tissues or organs is lower

in many of the new mAb drugs. Direct local mAb administration to target tissues is

enabled by biomaterials scaffolds in which drug retention at the intended sites can be

modulated. This dissertation is focused on two bioaffinity hydrogels designed for

administering mAbs locally. One was built on a fluorescent-based module: fluorogen

activating protein dL5 and its fluorogenic partner malachite green (MG) for immobilizing

IgG Fc binding molecule protein A/G into the hydrogel matrix. This design was shown

to render sustained release of mAbs and detectable binding sites in vivo through a built-in

imaging feature. The second design was built upon pG_EAK, a bifunctional polypeptide

in which the IgG-binding domain was genetically fused to the self-assembling sequence

(AEAEAKAK)3 (single amino acid code; hereafter “EAK”). The rationale was that

reducing the number of interactions would enhance the translational prospect of the

bioaffinity hydrogel. Sustained release and prolonged retention of IgG delivered in the

pG_EAK gel were observed in vitro and in vivo in skin allografts. In conclusion,

bioaffinity hydrogels built on dL5_EAK and pG_EAK are two injectable platforms on

which IgG mAbs of different specificities can be localized and sustained in diseased

tissues.

Page 7: Self-Assembling Bioaffinity Hydrogels for Localized

vi

DEDICATION

This dissertation is dedicated to my loving and supportive parents Yunxia Jiang

and Zhiling Liu.

Page 8: Self-Assembling Bioaffinity Hydrogels for Localized

vii

ACKNOWLEDGEMENT

I would like to give my sincere thanks to my advisor, Dr. Wilson Meng, for his

continuous support and guidance through my Ph.D. study and research. His diligent

devotion to teaching and research sets a very good role model for me as I continue with

my future career and life. I also would like to express my deepest gratitude to my

dissertation committee members, Dr. James Drennen, Dr. Ellen Gawalt, Dr. Jelena Janjic

and Dr. Patrick Flaherty, for their insightful advice and valuable feedback to my research

projects as well as generous provision of access to the instruments in their labs. I also

would like to extend my special thanks to Dr. Eric Freeman, Dr. Lauren O’Donnell, and

Mrs. Denise Butler-Buccilli for their help on release modeling, flow cytometry and

animal studies. I want to thank pharmaceutics faculty Dr. Carl Anderson, Dr. Peter

Wildfong, Dr. Ira Buckner, Dr. Dr. Robert E. Stratford and Dr. Devika Manickam for the

excellent pharmaceutics training and education I received at Duquesne.

I appreciate the companion and help I received from our current and previous

group members: Yi Wen, Shana L Roundebush, Yiwei Wang, Ngoc Pham, Amy

Spigelmyer, Justin Nedzesky. I want to thank Nina Reger for her patient help on FTIR

and SEM, Priya Ganesan for her help on flow cytometry, Yi Han for his help in drawing

dissertation figures. Besides, I am grateful to have the friendship with all the graduate

students in school of pharmacy. They added color to my life here and help me grow to a

better person.

Last but not least, I would like to thank my parents for their endless love and

support through the journey. Without their encouragement, I would not get to this point.

Page 9: Self-Assembling Bioaffinity Hydrogels for Localized

viii

TABLE OF CONTENTS

Page

ABSTRACT ............................................................................................................................... iv

DEDICATION ........................................................................................................................... vi

ACKNOWLEDGEMENT ......................................................................................................... vii

TABLE OF CONTENTS ......................................................................................................... viii

LIST OF TABLES ..................................................................................................................... xi

LIST OF FIGURES ................................................................................................................... xii

LIST OF ABBREVIATIONS .................................................................................................. xiv

Chapter 1 ......................................................................................................................................... 1

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

Monoclonal antibody therapeutics .......................................................................................... 1

Unmet needs in therapeutic antibody delivery ........................................................................ 9

Sustained-release strategies for therapeutic antibodies ......................................................... 10

Drug delivery hydrogel formed by β-sheet fibrillizing peptide ............................................ 12

Bioaffinity hydrogel for sustained release of antibodies ....................................................... 13

Projects overview .................................................................................................................. 17

Significance ........................................................................................................................... 19

Summary ............................................................................................................................... 22

Chapter 2 ....................................................................................................................................... 24

Literature Review: Self-assembling Peptides as Hydrogel Forming Materials ......................... 24

Introduction ........................................................................................................................... 25

Analytical techniques in characterizing the assemblies......................................................... 28

Fluorescent spectroscopy ...................................................................................... 28

Atomic force microscopy (AFM) ......................................................................... 29

Circular dichroism (CD) ....................................................................................... 30

Fourier transform infrared (FTIR) spectroscopy .................................................. 32

Oscillatory rheology.............................................................................................. 33

Surface tension measurement ............................................................................... 34

Other analytical techniques ................................................................................... 35

Self-assembling peptide amphiphiles and their corresponding properties ............................ 35

I. Ionic-complementary self-assembling peptide .................................................. 35

Page 10: Self-Assembling Bioaffinity Hydrogels for Localized

ix

II. Self-assembling β-hairpins ............................................................................... 52

III. Alkylated peptide amphiphiles ....................................................................... 59

Common features of these self-assembling peptide amphiphiles .......................................... 69

Conclusions and future prospects .......................................................................................... 72

Comparison and property predictions ................................................................... 72

Standardization of analytical techniques .............................................................. 73

Chapter 3 ....................................................................................................................................... 76

Genetic Engineering of a Fluorogenic Module in Bioaffinity Hydrogels for Antibody Delivery

................................................................................................................................................... 76

Background and Rationale .................................................................................................... 77

Materials and Methods .......................................................................................................... 80

EAK16-II and expression of dL5_EAK ............................................................... 80

Conjugation of protein A/G .................................................................................. 81

Gel electrophoresis................................................................................................ 82

Viscosity measurement ......................................................................................... 82

Titration of pAGMG to dL5_EAK gel ................................................................... 83

Laser-scanning confocal microscopy .................................................................... 83

In vitro IgG release ............................................................................................... 84

Modeling Methodology ........................................................................................ 85

Mice and in vivo studies ....................................................................................... 88

Results ................................................................................................................................... 90

pAGMG conjugate .................................................................................................. 94

Localization of IgG in pAGMG-embedded dL5 hydrogel ..................................... 95

In vitro IgG release from pAGMG-embedded dL5 hydrogel ................................. 98

Modeling and simulation of release .................................................................... 100

Deposition of dL5 gel formulated IgG in vivo ................................................... 105

Discussion ........................................................................................................................... 109

Conclusions ......................................................................................................................... 113

Chapter 4 ..................................................................................................................................... 114

A Facile Fc-binding Composite in Bioaffinity Hydrogel for Antibody Delivery ................... 114

Background and rationale .................................................................................................... 115

Materials and Methods ........................................................................................................ 118

Peptides and reagents .......................................................................................... 118

Expression and purification of pG_EAK ............................................................ 118

Page 11: Self-Assembling Bioaffinity Hydrogels for Localized

x

MALDI-TOF mass spectrometry ........................................................................ 120

Coassembly experiments .................................................................................... 120

IgG binding experiments..................................................................................... 123

In vitro IgG release ............................................................................................. 126

Mice and in vivo studies ..................................................................................... 126

Results ................................................................................................................................. 130

Confirmation of pG_EAK molecular mass......................................................... 130

Co-assembly ........................................................................................................ 131

Injectable composite ........................................................................................... 132

Capture of IgG in pG gels in vitro ...................................................................... 133

In vitro IgG release ............................................................................................. 137

Mice and in vivo studies ..................................................................................... 139

Discussion ........................................................................................................................... 146

Conclusion ........................................................................................................................... 149

Appendix: .................................................................................................................................... 150

I. Miniaturization of IgG binding domain in the bioaffinity hydrogel design ..................... 150

Z34c functionalized EAK gel ............................................................................. 150

Inhibition of neutrophil trafficking in skin transplant model using anti-neutrophil

surface marker IgGs displayed by Z34c gel ....................................................... 154

II. Bioaffinity Hydrogels for Cell Therapeutics .................................................................. 160

3D cell culture facilitated by hydrogels .............................................................. 160

Construction of mini thymus unit through bioaffinity hydrogel assisted 3D

aggregates of thymus epithelium cells (TECs) ................................................... 161

References ................................................................................................................................... 166

Page 12: Self-Assembling Bioaffinity Hydrogels for Localized

xi

LIST OF TABLES

Page

Table 1-1 FDA approved mAb. .......................................................................................... 4 Table 2-1 The correlations between common protein structures and amide I frequencies

........................................................................................................................................... 33

Table 2-2 amino acid sequences studies by Caplan et al .................................................. 42 Table 2-3 Comparison for properties of representative peptide amphiphiles from each

category ............................................................................................................................. 70 Table 3-1 Formulations used in in vivo experiments........................................................ 89 Table 3-2 Apparent diffusivities (Dapp) determined using approximation of early releasea

......................................................................................................................................... 102

Table 4-1 Equilibrium constants comparison between protein G and protein A with IgG

from various species ....................................................................................................... 116

Table 4-2 Binding and kinetic constants of pG_EAK .................................................... 134

Table 4-3 Calculated apparent diffusion coefficients for both IgG displayed by EAK gel

and pG gel and the IgG remaining in the gel at the end of release study. ...................... 139 Table 5-1 immunogenicity prediction for different IgG binding domains based on MHC-

II ligand scanning using immune epitope database (IEDB) ........................................... 152

Page 13: Self-Assembling Bioaffinity Hydrogels for Localized

xii

LIST OF FIGURES

Page

Figure 1-1 the Y shaped structure of antibody.................................................................... 2 Figure 1-2 schematic representation of the FAP dL5 embedded bioaffinity hydrogel

design ................................................................................................................................ 16

Figure 1-3 schematic representation of pG embedded bioaffinity hydrogel .................... 17 Figure 2-1 CD spectra for polypeptides and proteins with representative secondary

structures ........................................................................................................................... 32 Figure 2-2 Schematic three-dimensional molecular model of EAK16-II ......................... 37 Figure 2-3 Initially proposed model of the membrane by Zhang et al ............................. 38

Figure 2-4 SEM images of membranes formed by EAK16-II with serial magnification . 39

Figure 2-5 proposed β-hairpin structure and amino acid sequence of MAX 1 ................. 53 Figure 2-6 (a) proposed self-assembly pathway for MAX1, (b) the structure of self-

assembled MAX1 β-hairpin molecules in a fibril and (c) TEM micrographs of MAX1 . 54

Figure 2-7 Schematic representations for two different self-assembled cross-links ....... 56 Figure 2-8 (a) Chemical structure of one alkylated peptide amphiphile with 5 key

structure regions. (b) Molecular model of the PA showing the overall conical shape of the

molecule going from the narrow hydrophobic tail to the bulkier peptide region; (c)

Schematic showing the self-assembly of peptide amphiphile molecules into a cylindrical

micelle; (d) Vitreous ice cryo-TEM image of the fibers reveals the diameter of the fibers

........................................................................................................................................... 61 Figure 2-9 The simulated morphologies that could exist during PA molecule self-

assembly process ............................................................................................................... 63

Figure 2-10 Schematic representation of a single β-sheet ................................................ 65 Figure 3-1 Diagrams showing (a): the construct of dL5_EAK and (b) pAGMG embedded

dL5 gel displaying IgG molecules. ................................................................................... 80

Figure 3-2 Morphological features of EAK16-II and dL5_EAK/EAK16-II hydrogels as

revealed in (a) Optical images and (b) SEM images ........................................................ 91

Figure 3-3 Co-assembly and association of dL5 gel system components ........................ 92 Figure 3-4 Viscosity of EAK16-II at different salt concentrations................................... 94 Figure 3-5 Electrophoretic analysis of native pAG and pAGMG using SDS-PAGE. ........ 95

Figure 3-6 Binding and localization of pAGMG in dL5 hydrogel ..................................... 96 Figure 3-7 Loading and co-localization of IgGFITC in dL5 hydrogels .............................. 97 Figure 3-8 Monitoring in vitro IgG release and pAGMG – dL5 interaction ...................... 99

Figure 3-9 Biological activity of released anti-human IgGFITC tested by binding to

immobilized human IgG ................................................................................................. 100

Figure 3-10 Modeling and simulation of IgG release in vitro ........................................ 101 Figure 3-11 Retention of IgG and binding sites in footpad ............................................ 106 Figure 3-12 Distribution of IgG800 injected in footpads 72 h after injection .................. 108 Figure 4-1 Diagrams showing (a): the construct of pG_EAK and (b) coassembled pG gel

displaying IgG molecules. .............................................................................................. 117

Figure 4-2 Diagram showing the procedure flow for pG_EAK plasmid construction,

transformation, expression and purification. .................................................................. 120 Figure 4-3 MALDI-TOF mass spectroscopy of pG_EAK protein. ................................ 130

Page 14: Self-Assembling Bioaffinity Hydrogels for Localized

xiii

Figure 4-4 Coassembly between EAK16-II with pG_EAK ........................................... 132

Figure 4-5 Viscosity profile of pG gel and its physical appearance visualized by Congo

red ................................................................................................................................... 133 Figure 4-6 Capture of IgG in pG gels in vitro ................................................................ 135 Figure 4-7 Emission scan of three different dye labeled (FITC, PE and PerCP-Cy5.5)

IgGs on the pG hydrogel. ................................................................................................ 136 Figure 4-8 Self quenching phenomenon of FITC-IgG loaded on pG gel ...................... 137

Figure 4-9 In vitro release of IgG from the gel ............................................................... 138 Figure 4-10 IgG retention in subcutaneous space after injection into the back of the mice.

......................................................................................................................................... 140 Figure 4-11 Interleukin 6 production .............................................................................. 141 Figure 4-12 Retention of subcutaneously injected IgG800 under inflammatory condition

(skin transplantation) ...................................................................................................... 143

Figure 4-13 representative flow cytometry plots of in vitro allogeneic re-stimulated

splenocytes from mice treated with pG gel displayed Fc-CTLA and Fc-TGF-β1 (exp) or

only pG gel (control) after allo skin transplantation surgery .......................................... 145

Figure 5-1 Diagrams showing (a): the construct of Z34c_EAK and (b) coassembled Z34c

gel displaying IgG molecules.......................................................................................... 151 Figure 5-2 IgG loading on Z34c bioaffinity hydrogel and its colocalization with gel fibrils

......................................................................................................................................... 153 Figure 5-3 The fluorescence decay over 2 days after injection of IgG800 with (test) or

without (control) Z34c gel at the interface of host-skin graft. ........................................ 154 Figure 5-4 Schematic diagram showing the design of anti-neutrophil Z34c gel delivered

at the interface of host-graft in skin transplant model .................................................... 155

Figure 5-5 verification of neutrophil infiltration after skin transplantation surgery at the

graft site .......................................................................................................................... 156 Figure 5-6 Graft homing-neutrophil impediment at host-graft interface. ....................... 157 Figure 5-7 Odyssey scan (800nm channel) for particle distribution before and different

time points after the addition of release medium to the system. ..................................... 158 Figure 5-8 Schematic drawing showing the strategies to generate the TECs mini thymus

unit .................................................................................................................................. 161 Figure 5-9 Generation of TECs mini thymus complex in vitro ...................................... 164

Page 15: Self-Assembling Bioaffinity Hydrogels for Localized

xiv

LIST OF ABBREVIATIONS

Ab: Antibody

Ag: Antigen

Ig: Immunoglobulin

pAbs: polyclonal antibodies

mAb: monoclonal antibody

irAEs: immune-related adverse events

API: active pharmaceutical ingredient

βFP: β-sheet fibrillizing peptide

SAPs: self-assembling peptides

EAK16-II: AEAEAKAKAEAEAKAK or (AEAEAKAK)2

FAP: fluorogen-activating protein

IgG: Immunoglobin G

pA/G: protein A/G

MG: Malachite Green

pAGMG: protein A/G Malachite Green conjugate

dL5_EAK: dL5_(AEAEAKAK)3

pG_EAK: protein G_(AEAEAKAK)3

E. Coli.: Escherichia Coli.

CTLA-4: cytotoxic T-lymphocyte antigen 4

PD-1: programmed cell death 1

Page 16: Self-Assembling Bioaffinity Hydrogels for Localized

xv

PD-L1: programmed cell death ligand 1

SC: subcutaneous

IV: intravenous

ECM: extracellular matrix

TECs: thymus epithelial cells

PA: peptide amphiphile

scFv: single-chain variable fragment

ADCC: antibody-dependent cell-mediated cytotoxicity

CDC: complement-dependent cytotoxicity

Page 17: Self-Assembling Bioaffinity Hydrogels for Localized

1

Chapter 1

Introduction

Monoclonal antibody therapeutics

Antibodies (Abs), also known as immunoglobulins (Igs), are glycoproteins

produced by activated B lymphocytes in response to antigens (Ags). The Ags often

originate from bacteria, viruses and other microorganisms, but can also arise from

endogenous proteins. While the body generates Ab with diverse specificities, each Ab

binds to its cognate Ag in a specific way, with the paratope from Ab interacting with the

epitope from the Ag by spatial complementarity like lock and key. Electrostatic forces,

hydrogen bonding, hydrophobic interactions and van der Waal forces constitute the

reversible interaction between Ab and Ag. The basic structural unit of Ab is typically a

Y shaped molecule composed of two large heavy chains and two small light chains

(Figure 1-1). Light chains, comprising λ or κ peptide subunits, are connected to heavy

chains by disulfide bonds, giving rise to a symmetrical molecule with two identical

antigen-binding sites. The two heavy chains are also connected together through

disulfide bonds which constitute the hinge region providing various flexibility to each

antigen binding site in different subclasses of antibodies. The antigen-binding fragment

(Fab) is located on the tip of the “Y”. It is composed of heavy and light chains and is

made up of six hypervariable loops known as hypervariable region with each variable

region binding to the specific Ag. The hypervariable regions are also called

complementarity-determining regions (CDRs). The enormous diversity of Fab allows the

immune system to recognize a wide variety of Ags. On the other hand, crystallizable

Page 18: Self-Assembling Bioaffinity Hydrogels for Localized

2

fragment (Fc) of the Ab is located at the base of "Y” composed of only heavy chains. Fc

is a constant region that can be used to distinguish between the five different subtypes of

Igs because the constant heavy chain can bind to its specific Fc receptor to communicate

with other components of the immune system, helping generate appropriate immune

response for different foreign objects the body may encounter. Thus, the general

structure of Ab is embedded with multiple functionalities [1].

Figure 1-1 the Y shaped structure of antibody.

There are five Ab classes in mammals: IgA, IgD, IgE, IgG and IgM, named

according to the type of their heavy chain types in α, γ, δ, ε, and μ respectively. They

differ in their biological properties, distribution and functions. IgA, usually present in

dimer, is mostly found in mucosal area, preventing colonization by pathogens [2]. IgE is

mainly involved in allergy by binding to allergens and triggers histamine release from

other immune cells. Both IgM (in the form of monomer) and IgD are expressed on the

surface of naive B cells, rendering the B cells ready to respond to Ag. IgM also exists as

a pentamer with very high avidity when in a secreted form and eliminates pathogens in

Page 19: Self-Assembling Bioaffinity Hydrogels for Localized

3

the early stages of humoral immunity before sufficient IgG is produced [3]. After

activation through the binding to Ag, B cells can divide and differentiate into Ab

producing cell (plasma cells). As the most abundant isotype of Ab (~75% of serum Ab),

IgG provides the majority of Ab based immunity protecting the body from infection. It is

also the only isotype that has receptors to facilitate passage through human placenta to

provide protection to the fetus in utero before its own immune system develops [1].

As the major component of humoral immunity, IgG plays its role by several

mechanisms including neutralizing binding to pathogens, immobilization and

agglutination of pathogens, coating of pathogen surfaces for the recognition and ingestion

by phagocytic immune cells (opsonization, phagocytosis and clearance), intracellular

antibody-mediated proteolysis, activation of classical pathway of complement system

(complement-dependent cytotoxicity, CDC) and antibody-dependent cell-mediated

cytotoxicity (ADCC) [4]. There are four subclasses of IgG in human: IgG1, IgG2, IgG3

and IgG4, named in order of their abundance in serum. IgG1 is the most abundant

subclass, comprising about two thirds of all IgG molecules. The four subclasses of IgG

differ mainly in the structure of hinge regions, leading to unique biological properties of

each class. They emerge at different stages of immune response and vary in segmental

flexibility, binding affinity to Fc receptor on phagocytic cells as well as the ability to

activate complement system [5].

Because of the high specificity of Ab to their cognate Ag, Ab has risen as the

leading molecular platform in the development of therapeutics for autoimmune diseases,

cancer, and inflammatory disease. While polyclonal antibodies (pAbs) are a collection of

Page 20: Self-Assembling Bioaffinity Hydrogels for Localized

4

Ig molecules that react against different epitopes on a specific antigen, monoclonal

antibodies (mAbs) are produced by a unique parent cell, binding to the same epitope of

an antigen. Serum-pooled human pAb treatment provides wide population diversity of

Ab families against many possible targets, but suffers from the disadvantages including

nonspecific Ab components with possible side effects and no rational efficacy against

diseases, diluting effector functions specific to desired targets and effectors. Cost,

inconsistent efficacy and questionable reliability, toxicity and safety issues for a pooled

human blood-derived protein product hinders its clinical acceptance as an Ab therapy [6].

After mAb technique was first created in 1975 by using mouse x mouse hybridoma

(fusion of myeloma cell lines with B cells), the production of mAb at bulk quantity have

become possible [7]. Since FDA approved the first therapeutic mAb in 1986 to reduce

acute rejection after organ transplants [8], there has been over 70 therapeutic mAb

approved in the US as of March 29, 2018 (Table 1-1). The approval rate of mAb drugs is

historically high recently, with more than 10 therapeutic mAb products approved in 2017

alone. The global sales revenue for all mAb products constitutes approximately half of

the total sales of all biopharmaceutical products. The continuous growth in sales of the

approved mAb products, along with hundreds of mAb product candidates currently in

clinical trials, will lead to sustained growth in sales of mAb products and an expanding

market share of mAb in biopharmaceutical industry.

Table 1-1 FDA approved mAb. Information is up to date as of March 29, 2018.

Drug

Name

Active Ingredients Initial U.S

Approval

Company Target* Therapeutic

indications*

Ilumya Tildrakizumab-

Asmn

3/20/2018 Merck Sharp

Dohme

IL-23 Ps

Trogarzo Ibalizumab-

Uiyk

3/6/2018 Taimed

Biologics USA

CD4 HIV-1

Page 21: Self-Assembling Bioaffinity Hydrogels for Localized

5

Ixifi Infliximab-Qbtx 12/13/2017 Pfizer Inc TNFα CD; UC; RA; AS

etc

Ogivri Trastuzumab-

Dkst

12/1/2017 Mylan Gmbh HER2 breast cancer

Hemlibra Emicizumab 11/16/2017 Genentech Inc Factor Ixa-

& Factor X

bleeding episodes

Fasenra Benralizumab 11/14/2017 Astrazeneca Ab IL5Rα asthma

Mvasi Bevacizumab-

Awwb

9/14/2017 Amgen Inc VEGF Metastatic

Colorectal Cancer

etc.

Cyltezo Adalimumab-

Adbm

8/25/2017 Boehringer

Ingelheim

TNFα RA; JIA; PsA; AS;

CD; UC; Ps

Besponsa Inotuzumab

Ozogamicin

8/17/2017 Wyeth Pharms

Inc

CD22 ALL

Tremfya Guselkumab 7/13/2017 Janssen Biotech IL-23 Ps

Rituxan

Hycela

Hyaluronidase;

Rituximab

6/22/2017 Genentech Inc CD20 FL; DLBCL; CLL

Kevzara Sarilumab 5/22/2017 Sanofi

Synthelabo

IL-6R RA

Imfinzi Durvalumab 5/1/2017 Astrazeneca Uk

Ltd

PD-L1 urothelial

carcinoma

Renflexis Infliximab-

Abda

4/21/2017 Samsung

Bioepsis Co Ltd

TNFα CD; UC; RA etc

Dupixent Dupilumab 3/28/2017 Regeneron

Pharmaceuticals

Il-4Rα atopic dermatitis

Ocrevus Ocrelizumab 3/28/2017 Genentech Inc CD20 MS

Bavencio Avelumab 3/23/2017 Emd Serono Inc Pd-L1 MCC

Siliq Brodalumab 2/15/2017 Valeant

Luxembourg

Il-17Ra Ps

Zinplava Bezlotoxumab 10/21/2016 Merck Sharp

Dohme

Clostridium

Difficile

Toxin B

CDI

Lartruvo Olaratumab 10/19/2016 Eli Lilly And Co PDGFR-α STS

Amjevita Adalimumab-

Atto

9/23/2016 Amgen Inc TNFα RA etc.

Zinbryta Daclizumab 5/27/2016 Biogen IL-2Rα

(CD25)

MS

Tecentriq Atezolizumab 5/18/2016 Genentech Inc PD-L1 urothelial

carcinoma

Inflectra Infliximab-

Dyyb

4/5/2016 Celltrion Inc TNFα CD; UC; RA etc

Cinqair Reslizumab 3/23/2016 Teva Respiratory

LLC

IL-5 asthma

Taltz Ixekizumab 3/22/2016 Eli Lilly And Co IL-17A Ps

Anthim Obiltoxaximab 3/18/2016 Elusys

Therapeutics Inc

PA

component

of B.

anthracis

toxin

Inhalational

Anthrax

Empliciti Elotuzumab 11/30/2015 Bristol Myers

Squibb

SLAMF7 multiple myeloma

Portrazza Necitumumab 11/24/2015 Eli Lilly Co EGFR lung cancer

Darzalex Daratumumab 11/16/2015 Janssen Biotech CD38 multiple myeloma

Page 22: Self-Assembling Bioaffinity Hydrogels for Localized

6

Nucala Mepolizumab 11/4/2015 Glaxosmithkline

LLC

IL-5 asthma

Praxbind Idarucizumab 10/16/2015 Boehringer

Ingelheim

direct

thrombin

inhibitor

dabigatran

bleeding

Repatha Evolocumab 8/27/2015 Amgen Inc PCSK9 HeFH; HoFH

Praluent Alirocumab 7/24/2015 Sanofi Aventis PCSK9 HeFH etc

Unituxin Dinutuximab 3/10/2015 United Therap glycolipid

GD2

neuroblastoma

Cosentyx Secukinumab 1/21/2015 Novartis Pharms

Corp

IL-17A Ps

Opdivo Nivolumab 12/22/2014 Bristol Myers

Squibb

PD-1 melanoma; lung

cancer

Blincyto Blinatumomab 12/3/2014 Amgen CD19 ALL

Keytruda Pembrolizumab 9/4/2014 Merck Sharp

Dohme

PD-1 melanoma

Entyvio Vedolizumab 5/20/2014 Takeda Pharms

USA

integrin

receptor

UC; CD

Sylvant Siltuximab 4/23/2014 Janssen Biotech IL-6 MCD

Cyramza Ramucirumab 4/21/2014 Eli Lilly And Co VEGFR2 Gastric Cancer

Gazyva Obinutuzumab 11/1/2013 Genentech CD20 chronic

lymphocytic

leukemia

Kadcyla Ado-

Trastuzumab

Emtansine

2/22/2013 Genentech HER2 breast cancer

Raxibacumab Raxibacumab 12/14/2012 Human Genome

Sciences Inc

PA

component

of B.

anthracis

toxin

inhalational

anthrax

Perjeta Pertuzumab 6/8/2012 Genentech HER2 breast cancer

Adcetris Brentuximab

Vedotin

8/19/2011 Seattle Genetics CD30 Hodgkin

Lymphoma

Yervoy Ipilimumab 3/25/2011 Bristol Myers

Squibb

CTLA-4 melanoma

Benlysta Belimumab 3/9/2011 Human Genome

Sciences Inc.

BLyS or

BAFF

SLE

Prolia Denosumab 6/1/2010 Amgen RANK

ligand

osteoporosis

Xgeva Denosumab 6/1/2010 Amgen

Actemra Tocilizumab 1/8/2010 Genentech IL-6R RA

Arzerra Ofatumumab 10/26/2009 Glaxo Grp Ltd CD20 CLL

Stelara Ustekinumab 9/25/2009 Centocor Ortho

Biotech Inc

IL-12 &

IL-23

Ps

Stelara Ustekinumab 9/25/2009 Janssen Biotech IL-12 &

IL-23

Ps; PsA; CD

Ilaris Canakinumab 6/17/2009 Novartis Pharms IL-1β CAPS; FCAS;

MWS

Simponi Golimumab 4/24/2009 Centocor Ortho

Biotech Inc

TNFα RA; PsA; AS

Page 23: Self-Assembling Bioaffinity Hydrogels for Localized

7

Simponi Aria Golimumab 4/24/2009 Janssen Biotech TNFα RA

Cimzia Certolizumab

Pegol

4/22/2008 Ucb Inc TNFα CD

Soliris Eculizumab 3/16/2007 Alexion Pharm C5 PNH

Vectibix Panitumumab 9/27/2006 Amgen EGFR colorectal carcinom

Lucentis Ranibizumab 6/30/2006 Genentech VEGF-A macular

degeneration

Tysabri Natalizumab 11/23/2004 Biogen Idec α4-integrin MS

Avastin Bevacizumab 2/26/2004 Genentech VEGF carcinoma of the

colon & rectum

Erbitux Cetuximab 2/12/2004 Imclone EGFR colorectal

carcinoma

Xolair Omalizumab 6/20/2003 Genentech human IgE asthma

Humira Adalimumab 12/31/2002 Abbvie Inc TNF RA

Zevalin Ibritumomab

Tiuxetan

2/19/2002 Spectrum

Pharms

CD20 non-Hodgkin's

lymphoma

Campath Alemtuzumab 5/7/2001 Genzyme CD52

B-CLL

Lemtrada Alemtuzumab 5/7/2001 Genzyme CD52 B-CLL

Mylotarg Gemtuzumab

Ozogamicin

5/17/2000 Wyeth Pharms

Inc

CD33 myeloid leukemia

Herceptin Trastuzumab 9/25/1998 Genentech HER2 metastatic breast

cancer

Remicade Infliximab 8/24/1998 Centocor Inc TNFα CD

Synagis Palivizumab 6/19/1998 Medimmune RSV F

protein

lower respiratory

tract disease

Simulect Basiliximab 5/12/1998 Novartis IL-2Rα

(CD25)

acute organ

rejection

Rituxan Rituximab 11/26/1997 Genentech CD20 lymphoma.

Prostascint Capromab

Pendetide

10/28/1996 Cytogen PSMA diagnostic imaging

agent.

Myoscint Imciromab

Pentetate

7/3/1996 Centocor Inc human

myosin

cardiac imaging

agent

Reopro Abciximab 12/22/1994 Centocor Inc GP IIb/IIIa

receptor

of human 9

platelets

cardiac ischemic

complications

Summarized from FDA website [9].

* Abbreviations: Interleukin: IL; Cluster of differentiation: CD; tumor necrosis factor

alpha: TNFα; human epidermal growth factor receptor 2: HER2; receptor: R; vascular

endothelial growth factor: VEGF; human programmed death receptor-1: PD-L1; platelet-

derived growth factor receptor alpha: PDGFR-α; Signaling Lymphocytic Activation

Molecule Family member 7: SLAMF7; epidermal growth factor receptor: EGFR;

proprotein convertase subtilisin kexin type 9: PCSK9; protective antigen: PA; human

vascular endothelial growth factor receptor 2: VEGFR2; soluble human B lymphocyte

stimulator protein: BLyS or BAFF; cytotoxic T lymphocyte antigen-4: CTLA-4;

complement protein C5: C5; respiratory syncytial virus: RSV; prostate specific

membrane antigen: PSMA; glycoprotein: GP; acute lymphoblastic leukemia: ALL;

Crohn’s Disease: CD; Ulcerative Colitis: UC; Rheumatoid Arthritis: RA; Merkel cell

Page 24: Self-Assembling Bioaffinity Hydrogels for Localized

8

carcinoma: MCC; Follicular Lymphoma: FL; Clostridium difficile infection: CDI; soft

tissue sarcoma: STS; multiple sclerosis: MS; Juvenile Idiopathic Arthritis: JIA; Psoriatic

Arthritis: PsA; Ankylosing Spondylitis: AS; Plaque Psoriasis: Ps; Diffuse Large B-cell

Lymphoma: DLBCL; Chronic Lymphocytic Leukemia: CLL; multicentric Castleman’s

disease: MCD; systemic lupus erythematosus: SLE; heterozygous familial

hypercholesterolemia: HeFH; hypercholesterolemia: HoFH; active psoriatic arthritis:

PsA; CryopyrinAssociated Periodic Syndromes: CAPS; Familial Cold Autoinflammatory

Syndrome: FCAS; Muckle-Wells Syndrome: MWS; paroxysmal nocturnal

hemoglobinuria: PNH; B-cell chronic lymphocytic leukemia: B-CLL.

The first mAb drug OKT3 (Ortho Biotech) is murine-derived which brought acute

immunogenicity problems from patients responsive to murine-sourced Ab. Chimeric or

humanized mAb product emerged nearly a decade later to reduce the side effects

associated with the potential immunogenicity [10]. Most mAb products on market now

are humanized and their therapeutic effect is usually achieved through various

mechanisms including neutralization (bind to the ligand or receptor expressed on the cell

surface and block the target signaling pathway), ADCC, CDC or even antibody-mediated

toxin delivery [11]. When the body is in a diseased or dysfunctional condition, there is

typically an upregulation or downregulation of immune cell activation level and cytokine

secretion level. Exogenous therapeutic mAb could thus be administered to counteract

this irregular condition through immune response augmentation or suppression. In cancer

therapy, a need for cytotoxic T cell stimulation is usually favorable. For example, mAb

product Ipilimumab, which binds to the negative regulator of T cell activation CTLA-4,

could be used to block its interaction with its ligands CD80/CD86. The blockade of

CTLA-4 has been shown to augment T-cell activation and proliferation, thus enhancing

the treatment to melanoma. In a similar way, PD-L1, which blocks T-cell function and

activation through interaction with PD-1 and CD80, could serve as another anti-cancer

target. Delivering mAb binding to PD-L1 (e.g. Durvalumab) could release the inhibition

Page 25: Self-Assembling Bioaffinity Hydrogels for Localized

9

of immune responses, increase T-cell activation in vitro and decrease tumor size in vivo.

In contrast, immune system sometimes needs to be suppressed through mAb intervention

in transplantation management to reduce the incidence of rejection or prolong the graft

survival time. For example, Basiliximab, which binds to IL-2Rα expressed on the surface

of activated T cells, competitively inhibits IL-2 mediated activation of lymphocytes, a

critical pathway in the cellular immune response involved in allograft rejection.

Therefore, Basiliximab is used to prevent rejection in organ transplantation, especially in

kidney transplant. In summary, the immune regulation exerted by mAb therapeutics

could be two-ways: either in cancer therapy to boost effector T cell function against

tumor cells or as immunosuppressive agent in transplantation scenario or autoimmune

diseases to aid in induction of tolerance.

Unmet needs in therapeutic antibody delivery

Most mAb products on market are formulated for intravenous infusion or

injection, usually at high doses because of their stoichiometric rather than catalytic

pharmacodynamics. More recently there have been products approved for subcutaneous

administration. These are often packaged in pre-filled syringes designed for self-

administration by patients, which can improve compliance. Because of incomplete

absorption of subcutaneous administered mAbs into the systemic circulation, very high

doses are required (up to 100 mg/kg), making the costs of mAb therapies for patients who

need regular repeated dosing comparatively unaffordable [4]. Another important aspect

associated with systemic delivery of immune modulatory mAbs is the induction of

potential immune-related adverse events (irAEs). These effects could be significant

because the impacts could remain after cessation of the therapy, due to loss of T cell

Page 26: Self-Assembling Bioaffinity Hydrogels for Localized

10

tolerance. Such effects have been documented in checkpoint inhibitors in anti-CTLA-4

and anti-PD-1 mAbs in cancer patients, in some cases resulting in drug-induced

rheumatoid arthritis [12]. Based on these reasons, direct administration of therapeutic

mAbs into diseased tissues is justified. The technical challenge is that the mAb must be

retained at the site of intended action for extended durations. Therefore, localized

delivery vehicle for sustained release of mAbs at diseased site is essential to provide

opportunities to modulate the immune pathogenesis and at the same time reduce the risk

of irAEs.

Sustained-release strategies for therapeutic antibodies

Delivering active pharmaceutical ingredient (API) locally using matrix based

vehicles such as hydrogels or particles is a common strategy exploited to control the

release of the drug from delivery system. Drug loaded vehicle administered topically can

serve as a local drug depot to provide stable supply of the drug over a prolonged time

duration compared to drug delivered in aqueous formulation. The matrix geometry

including pore size, material architecture and interaction of the matrix with loaded drugs

would affect the release profile of the API from the vehicle [13].

Compared to conventional small molecular drugs, mAb drugs possess the

advantage of high intrinsic specificity and affinity to numerous targets and molecules of

various size and chemistry (binding constant at sub-nanomolar concentrations) [14].

However, due to its protein nature, it is easy to denature and aggregate under harsh

conditions. Their functions, which are linked to their 3D conformation, would be

compromised if they are subject to chemical and physical degradation including

Page 27: Self-Assembling Bioaffinity Hydrogels for Localized

11

proteolytic and chemical degradation, physical unfolding and aggregation. Therefore,

mAb products are usually limited to parenteral formulations to avoid proteolytic

degradation in GI tract. While particle-based system does improve the injectability of

these formulations and achieve prolonged release, most of them suffer from low

encapsulation efficiency or incomplete release [15]. Also, organic solvent and high

mechanical energy introduced during the preparation of the polymer-based particles

could be detrimental to mAb stability. In addition, environmental changes such as pH

within the particle matrix throughout the release process could destabilize the

encapsulated protein molecule [16].

Alternatively, hydrogels, which are made of crosslinked polymers, can serve as

vehicles for mAb loading and release. The gel formation process usually proceeds at

ambient temperature and organic solvents are rarely required in the process due to the

hydrophilic nature of gel forming polymers [17]. The mild preparation condition of

hydrogels can help to preserve protein stability. The resulting porous structure of the

hydrogels, along with their high water content, are suitable properties to accommodate

high loads of water soluble molecules like mAbs. The nanostructure of the hydrogel

network, especially the mesh size, determines the mobility of encapsulated molecules and

controls their release rate from the network. Upon injection, hydrogels can also adapt to

the tissue contour in vivo. In addition, ‘smart’ hydrogels, or stimuli-responsive hydrogels

can be constructed. With varying chemical compositions and ionic moieties designed to

monomer comprising the polymer, the gelation of polymer solution can be triggered upon

environment changes such as temperature, pH or ionic strength. When the solution-gel

transition occurs at physiological conditions, in-situ forming hydrogels could be obtained,

Page 28: Self-Assembling Bioaffinity Hydrogels for Localized

12

which is favorable in their biomedical applications including drug delivery. With the

exhibition of all the advantageous physicochemical properties, hydrogels can serve as an

excellent platform for local mAb delivery to preserve its stability while slow down its

release from the gel matrix [18].

Drug delivery hydrogel formed by β-sheet fibrillizing peptide

β-sheet fibrillizing peptide (βFP) is one type of self-assembling peptides (SAPs)

emerging as a novel material in the field of drug delivery and tissue engineering due to

their excellent biocompatibility and ease for manipulation [19-22]. The first discovered

βFP sequence EAK16-II (AEAEAKAKAEAEAKAK with A as alanine, E as glutamic

acid and K as lysine) self-assembles into cross-linking fibrils upon exposure to electrolyte

solutions [23]. EAK16-II can be used to formulate depots of small molecule drugs,

proteins, and even cells [24-27]. The in-situ gel forming feature of EAK16-II allows it to

be administered through conventional needled syringes. In gels made with the FP

sequence RADA, the molecular size of the encapsulated protein governs release kinetics;

IgG, the largest molecule tested, reaches maximum release by 24 hours [28].

Koutsopoulos et al. subsequently demonstrated that burst release can be suppressed by

encapsulating IgG in double-layer nanofibers using two amphiphilic building blocks,

RADA and KLDL [29]. These studies demonstrate the versatility of FP in controlling

the release of protein drugs.

The release rate of loaded protein from FP hydrogel is mainly controlled by the

relative size of the protein and the pore size of the hydrogel matrix because simple

diffusion is the only mechanism controlling the release and the diffusion coefficient is

Page 29: Self-Assembling Bioaffinity Hydrogels for Localized

13

affected by mesh size and protein size according to obstruction-based theories [17]. In

general, the smaller the drug molecule and the larger the pore size of the hydrogel matrix,

the faster the drug would release from the hydrogel. Therefore, large protein like IgG

would have the slowest release rate compared to other smaller molecules when the mesh

size of the gel, temperature and viscosity of the fluid are fixed and the protein gel non-

specific interaction is the key factor determining the release rate. Although Ab is

released slower in hydrogel than in simple aqueous solution, a burst release effect still

takes place within the first few hours [30]. Therefore, an improvement on the design of

the βFP based hydrogel is required to achieve the desired sustained release of Ab from

the gel.

Bioaffinity hydrogel for sustained release of antibodies

To further extend the release duration, an affinity mechanism controlling the

drug release from hydrogel could be introduced to the hydrogel design. This mechanism

is chemically-controlled release where the release of loaded protein drug is determined by

reactions that occurs within the hydrogel matrix. Reversible binding occurring between

the polymer network and protein falls into this category of reaction where the binding

equilibrium reflected by binding constant (Kd) and association rate (kon) or dissociation

rate (koff) would determine the drug release rate [17]. Biomolecules can provide affinity

interactions for tuning protein release from hydrogel matrices. The Sakiyama-Elbert

group has engineered heparin-embedded hydrogels as protein delivery systems. As

highly sulfated glycosaminoglycans, heparins bind growth factors through ionic

reversible interactions. These systems mimic reservoirs of endogenous growth factors in

extracellular matrix [31-36] in which secreted proteins are protected from denaturation

Page 30: Self-Assembling Bioaffinity Hydrogels for Localized

14

and aggregation. Hydrogels functionalized with aptamers have been used to control

protein release [37-40]. The Shoichet group has used Src homology domain 3 (SH3) to

tune the release of growth factors fused with SH3 ligands [41-44]. In these designs, rapid

diffusion is suppressed; instead, drug release is a function of the affinity of the binding

interactions (Kd), the relative ratio of binding sites (ligands) to proteins, and the

dissociation rate constant (koff) of the protein-binding site complexes [43]. Thus affinity-

based hydrogels afford the ability to tune release kinetics by adjusting classical

biochemical parameters. Based on this mechanism theory, through appropriate

functionalization of βFP, the retention or release of drug cargo could be extended by

introducing an affinity module onto the βFP sequence, allowing elongated therapeutic

effect. Hydrogels formed by the coassembly of EAK16-II and its functionalized analog

provide affinity mechanisms for tuning Ab release from hydrogel matrix.

The design of EAK16-II based bioaffinity hydrogel in our lab hinges on the

design of bifunctional constructs where IgG binding module is linked to gel forming

sequence EAK directly through recombinant technology or indirectly through a

functional domain attached to EAK to provide additional imaging advantage.

The IgG binding modules we employed in our design are essentially based on

protein A, protein G, or the recombinant protein A/G (pA/G). Protein A and protein G

are Ig-binding proteins found on the surface of cell wall of bacteria staphylococcus

aureus and streptococcal respectively [45, 46]. They are extensively used as antibody

purification tools because of their specific affinity to the Fc domain of Ab. Protein A

consists of five IgG binding domains and protein G consists of three IgG binding

Page 31: Self-Assembling Bioaffinity Hydrogels for Localized

15

domains [46, 47]. The interaction of protein A with Igs is pH dependent with the binding

capacity for protein A optimal at pH 8-9 whereas the binding capacity of protein G is

high over a broader pH range [48]. Protein G also has a wider reactivity profile than

protein A and it binds to all human IgG subclasses as well as rabbit, mouse and goat IgGs

but does not binds to IgM, IgA, and IgD [45]. Its binding to Igs is often stronger than

protein A [49]. pA/G is a recombinant protein comprising of four Ig binding sites from

protein A and two Ig binding sites from protein G and is ideal for binding the broadest

range of IgG subclasses from different species.

pA/G was used as the antibody binding moiety in the first generation of the

bioaffinity hydrogel design in my dissertation. In this design, protein A/G was

immobilized to the hydrogel matrix through a fluorescence-based module: a fluorogen-

activating protein (FAP) dL5 and its fluorogenic partner malachite green (MG). FAP

dL5 is a single chain variable antibody fragment light chain homodimer that has a

specific affinity for MG. The fluorescence of MG (λex = 653 nm, λem = 684 nm) is

enhanced 20,000-fold upon binding to dL5 due to the constraint of single bond rotation

within MG molecule. Functional dL5 was recombinantly attached to EAK to generate a

bifunctional construct dL5_EAK. dL5 can thus be embedded in the hydrogel by

coassembly of dL5_EAK with excess EAK16-II [50]. Both the immobilization and

detection of pA/G in the hydrogel could be achieved by conjugating MG to pA/G. With

the help of the conjugation, the resulting conjugate pAGMG helps to link the binding sites

pA/G to the hydrogel matrix through the binding between MG and embedded dL5. The

generated fluorescence upon binding provides signal for antibody binding sites detection

(Figure 1-2).

Page 32: Self-Assembling Bioaffinity Hydrogels for Localized

16

Figure 1-2 schematic representation of the FAP dL5 embedded bioaffinity hydrogel

design

The benefits of this design are three-fold: first, it provides affinity driven capture

and release; second, specific binding of IgG molecules to pA/G facilitates oriented IgG,

rendering reduced denaturation; third, the intrinsic fluorescence generated within the

fluorescence based module (dL5 and MG) allows easy monitoring of the IgG binding

sites.

The second generation of the bioaffinity hydrogel design (Figure 1-3) is based on

a novel bifunctional construct pG_EAK where a truncated Fc-binding domain pG was

directly attached to gel forming βFP in order to simplify the first generation design and

promote the translational prospect. With less components involved in the design,

therapeutic antibodies could be directly anchored to the bioaffinity hydrogel formed by

pG_EAK and EAK16-II coassemblies and the release rate of antibodies from the

hydrogel is directly controlled by their binding affinity with pG.

Page 33: Self-Assembling Bioaffinity Hydrogels for Localized

17

Figure 1-3 schematic representation of pG embedded bioaffinity hydrogel

Projects overview

My dissertation work centers on the design of biomaterials for therapeutic Ab

delivery. The biomaterial delivery platform is based on self-assembling βFP EAK16-II.

EAK 16-II and its functionalized analogs (dL5_EAK or pG_EAK) are formulated into

injectables to establish hydrogels in which antibodies could be stably localized. The

central hypothesis of this project is that the release of subcutaneously injected antibodies

could be extended using bioaffinity hydrogels made of βFP EAK16-II and its

functionalized analogs. To test this hypothesis, two designs of the bioaffinity hydrogel

were proposed and characterized.

In the first generation design (chapter 3), fluorogen activating protein dL5 was

attached to EAK sequence to form the functional analog dL5_EAK. The coassembly of

EAK16-II and dL5_EAK was optimized in molar ratio. The colocalization of dye labeled

IgG molecules with MG fluorescence after loading of the conjugate pAGMG and model

IgG within the bioaffinity hydrogels, have been experimentally verified. Prolonged

release and delivery of IgG were demonstrated in vitro and in vivo. A sustained release

profile was observed for our pAGMG-decorated dL5 hydrogels for at least two months.

Page 34: Self-Assembling Bioaffinity Hydrogels for Localized

18

For in vivo study, after subcutaneous foot pad injection of IgG loaded bioaffinity

hydrogels, local retention of the antibody showed higher signal intensity for 3 days

compared with hydrogels without pAGMG.

In the second generation design (chapter 4), Fc binding domain protein G (pG)

was directly attached to EAK sequence to form the functional construct pG_EAK. The

coassembly of pG_EAK with EAK16-II was verified using gel electrophoresis and

Congo Red absorbance change after binding. A superior IgG loading on pG hydrogel

and a sustained release profile of IgG from the gel for at least three months were

observed. The prolonged retention of IgG loaded on the bioaffinity pG hydrogels was

monitored in normal subcutaneous tissue for 2 days and murine skin allo-transplantation

model for 6 days after local injection compared to control system with free IgG or IgG

loaded in nonspecific EAK gel. Less systemic exposure of IgG displayed by pG

bioaffinity gel was reflected in the constrained organ distribution compared to the control

groups.

The continuous miniaturization of the Fc binding domain attached to EAK was

pursued to reduce the potential immunogenicity of the bioaffinity hydrogel. Z34c, a B

domain mimetic of native protein A, is a disulfide-linked cyclic peptide has a much

smaller molecular weight (4.18 kDa). It was fused to EAK to construct the bioaffinity

hydrogel with fewer potential MHC-II ligands in the Fc binding domain which indicates

a lower predicted immunogenicity compared to other Fc binding domains. Preliminary

studies have been performed to study the anti-neutrophil IgGs (anti-Ly6G and anti-

CD11b IgGs) loaded Z34c gel in inhibiting neutrophil trafficking in skin transplant model

Page 35: Self-Assembling Bioaffinity Hydrogels for Localized

19

(Appendix I). The infiltration of neutrophils at the graft site were verified both ex vivo

and in situ by detecting neutrophil elastase activity as the indicator. Reduced neutrophil

accumulation at transplantation site was observed in treated mouse by tracking the

neutrophils after tail vein injection of neutrophil specific, NIR fluorescent imaging agent.

The concept of combinational use of bioaffinity hydrogel and gel trapped particles

was proposed. Anti-neutrophil IgGs loaded bioaffinity hydrogel, coupled with anti-

neutrophil elastase particles (load with elastase inhibitor), could potentially work

synergistically to effectively impede continuous neutrophil infiltration as well as counter

tissue-destructive mechanism elicited by neutrophil elastase.

Another valuable application for bioaffinity hydrogel is cellular therapeutics

which is exemplified in the construction of mini thymic unit by 3D aggregation of TECs

with anti-EpCAM antibody displayed bioaffinity hydrogels (Appendix II). The formation

of 3D clusters of TECs facilitated by the bioaffinity hydrogel was observed in in vitro

culture and the cell viability was verified. The artificial mini thymus units were

transplanted into athymic nude mice and their support to T-lymphopoiesis in vivo was

confirmed [51].

Significance

Each year more than 1.1 million patients suffer from thermal and chemical burns

(Center for Disease Control and Prevention), with close to 30,000 hospitalizations in

specialized burn centers. Patients with extensive partial and full-thickness burns benefit

from intact epidermis and dermis which together serve as a protective barrier to minimize

desiccation of the underlying tissues, limit water evaporation, reduce bacterial

Page 36: Self-Assembling Bioaffinity Hydrogels for Localized

20

contamination, relieve pain, and promote wound healing by accelerating re-

epithelialization [52]. Despite the use of immunosuppressants, a significant fraction of

organ allografts undergoes acute rejection, resulting in substantial parenchyma loss

within weeks [53]. While multiple leukocytes are involved in the rejection mechanisms,

neutrophils are one of the early cell types infiltrating graft tissues and induce necrosis

[54]. Neutrophils rapidly infiltrate into wounds, constituting for more than 50% of all

cells on day one. These early arriving neutrophils can shape subsequent adaptive

immune responses [55, 56]. Currently there are no therapies approved for selective

suppression of neutrophil infiltration in transplant settings. Survival of skin allografts

can be prolonged by systemic depletion of neutrophils [57]. Wound healing is

accelerated in neutrophil-depleted mice [58].

One potential application of our bioaffinity hydrogel platform is to create a local

anti-neutrophil Ab depot at the transplant site to reduce graft damage by arresting local

neutrophil infiltration in the first 48 hours. Blunting the intensity of this early event is

expected to bend the overall trajectory of the rejection by limiting the production of

proteases, the flow of alloantigens to draining lymph nodes [59, 60], suppressing T cell

proliferation [61, 62], and subverting the migration of T cells to inflamed tissues [63].

Localized delivery of immunomodulatory mAb can increase drug concentration at the

inflamed graft while reducing non-specific toxicities in off-target tissues.

Another target for mAb therapeutics is tumor. mAbs could be locally

administered to the tumor site to reprogram the immune system more efficiently [64, 65].

The tumor site delivery can serve as a vaccination site to locally amplify the immune

Page 37: Self-Assembling Bioaffinity Hydrogels for Localized

21

system by redirecting the lymphocytes to the activated form instead of tolerant form,

thereby impeding the metastasis of the tumor. Delivery of Ab against the checkpoint

proteins such as PD-1 and CTLA-4 is an effective way to unleash T cell suppression in

cancer immunity [66]. Local delivery of Ab therapeutics by our bioaffinity hydrogels

could concentrate the IgGs to their antigen rich region, thus augmenting the immune

response toward tumor in a more efficient way.

The broader Impact of the research is that the bioaffinity hydrogels may lead to

development of new strategies in formulating mAb for subcutaneous (SC) administration

in attaining systemic absorption. While the majority of marketed mAbs are formulated

for intravenous (IV) administration, SC administration of mAb has the advantage of

allowing patients to self-administer in ambulatory settings. However, a key challenge in

SC administration of recombinant proteins is incomplete bioavailability, in particular at

high doses [67]. For macromolecules such as IgG (Mr >145 kDa), lymphatic uptake is

the main route of absorption from the SC space into the systemic circulation. Preclinical

studies indicate that movement of IgG through the extracellular matrix (ECM) is the rate-

limiting step. The slow diffusion of IgG through the ECM exposes the drug to pre-

systemic, intracellular catabolism [68]. The reduction of bioavailability via this

elimination pathway is compensated in part by binding of IgG to local tissue neonatal Fc

receptors (FcRn) and endosomal recycling, in which a fraction of the administered mAb

is protected from proteolytic degradation in cells. High doses of IgG (e.g. 40 mg/kg),

however, can saturate FcRn in local tissues, thereby increasing the fraction vulnerable to

proteolysis [69]. Reducing the rate of mAb release from SC depot constructed by our

Page 38: Self-Assembling Bioaffinity Hydrogels for Localized

22

bioaffinity hydrogels has the opportunity to increase bioavailability by keeping drug

concentrations in ECM below FcRn saturation.

Summary

In summary, embedding bioaffinity domain into fibrillar hydrogels can help

extend the retention of incorporated IgGs at injection sites and prolong the release of IgG

from local depot.

The first generation of EAK based functionalized gel described in this dissertation

incorporated a fluorescence based module to introduce an imaging feature in the design.

The fluorescence generated upon the binding of embedded functional domain dL5 to its

fluorogenic partner MG provide a convenient way to detect the IgG binding sites after

conjugating MG to protein A/G. The loading of IgG to the hydrogel was found to be

colocalized with MG fluorescence where protein A/G is located which indicates the

specific binding of IgG to protein A/G. The near to 700nm fluorescence peak wavelength

allows the imaging of binding sites in superficial tissue after subcutaneous injection of

the system locally. This facilitates the in vivo optimization of the antibody delivery

vehicle.

The second generation of EAK based bioaffinity hydrogel entails a simplified

design where the IgG binding domain protein G was directly attached to the self-

assembling domain EAK. With a wider and stronger reactivity profile towards Ab

compared to protein A, protein G embedded hydrogel provide a valid affinity mechanism

for efficient IgG loading and sustained release for IgGs of all human subtypes. The

improvement of IgG retention enabled by the bioaffinity hydrogels and restrained

Page 39: Self-Assembling Bioaffinity Hydrogels for Localized

23

systemic exposure was both verified in normal and inflamed tissues after subcutaneous

injection.

For both dL5 and pG gel designs, the ratio of the IgG dose and IgG binding sites

embedded in the hydrogel will determine the release profile for a specific formulation.

Higher dose of the IgG will render higher portion of IgG to be release through simple

diffusion with less percentage of IgG being sustained by the bioaffinity mechanism.

Therefore, by modulating the comparative amount of the two, locally administered IgG

aimed for systemic exposure and local effect could be manipulated to achieve the optimal

therapeutic outcome for a specific disease condition.

In addition, the bioaffinity hydrogel system could be exploited for cell

therapeutics in tissue engineering by enabling 3D organization and proper cell to cell

interaction in the cell culture. The formation of mini thymus unit by generation of 3D

aggregation of thymus epithelial cells (TECs) could be realized by mixing the cells with

anti-EpCAM (TECs surface marker) antibodies loaded bioaffnity hydrogels. The

successful support of T cell development by these mini thymus unit after transplanted to

athymic mouse suggests this bioaffinity gel platform could serve as a promising tool in

modulating immunity through cell therapeutics.

Page 40: Self-Assembling Bioaffinity Hydrogels for Localized

24

Chapter 2

Literature Review: Self-assembling Peptides as Hydrogel Forming

Materials

Self-assembly amphiphilic peptides have gained substantial interests among

biomaterials scientists in the past twenty years. The reasons being that peptides are

generally biodegradable and that novel peptides can be generated through amino acid

substitutions. They can form noninvasive and injectable supramolecular scaffolds upon

exposure to environmental stimuli such as temperature, pH, ionic strength. Although

initially present as soluble solutions, the self-assembled three-dimensional structures can

serve as a versatile platform for tissue regeneration and drug delivery. They are also

capable of displaying bioactive epitopes and/or proteins, as well as resembling

hierarchical structures of extracellular matrix (ECM) component. Extensive research has

been conducted on various self-assembly systems in recent years, from synthetic

molecules to naturally occurring biomaterials. This chapter will focus on the three most

widely studied self-assembling peptide amphiphiles: ionic-complementary self-

assembling peptides, self-assembling β-hairpins and alkylated peptide amphiphiles. The

first category is the design basis for our bioaffinity hydrogel. Common analytical

techniques used to characterize the micro- or macro-structure of the assemblies or to

investigate the self-assembling mechanisms will be summarized. The characteristics of

assemblies and their corresponding biomedical applications will be elaborated. The

limitations and challenges facing the development of peptide-based self-assembly

systems and the future directions will be discussed in this chapter as well.

Page 41: Self-Assembling Bioaffinity Hydrogels for Localized

25

Introduction

Molecular self-assembly is a very prevalent phenomenon in nature. For example,

phospholipid molecules can self-assemble into a bilayer and is the basis of cell membrane

composition. Detergent surfactant molecules can self-assemble into micelles when its

concentration exceeds critical micelle concentration (CMC) and help to clean poorly

soluble lipophilic species. Silk fibroin proteins can be fabricated into silk materials over

2km in length and have sufficient stiffness and toughness to be used in textile

manufacture [70]. By definition, molecular self-assembly is considered as the process by

which molecules spontaneously organize themselves into well-defined and stable

structures through non-covalent interactions under near thermodynamic equilibrium

conditions [71]. Essentially, it is a reversible formation of large molecular aggregates.

The effects of self-assembly can be both beneficial and deleterious. For instance, it can

serve as an essential mechanism for organism composition formation, such as cell

membranes. However, it can also be the cause of harmful diseases such as Alzheimer’s,

Parkinson’s, Huntington’s diseases as well as type II diabetes. For example, amyloid

fibril formation resulting from protein assembly is extensively considered as the cause for

the listed neurodegenerative diseases [72]. Amyloid β-protein (Aβ), a well-known self-

assembling biomolecule, has been considered as an important factor in the cause of

Alzheimer’s disease [73]. These protein assemblies are rich in cross β-sheet structures.

The macroscopic membranes or plaques formed by these assemblies are speculated to

interfere with and/or prevent cell-to-cell communications, thus leading to neurological

disorders [74].

Page 42: Self-Assembling Bioaffinity Hydrogels for Localized

26

There are numerous molecular systems that can self-assemble, such as collagen

mimetic peptides, peptide amphiphiles, protein polymer hybrids and DNA-based

materials. Among these different types of materials, self-assembling peptide amphiphiles

have been extensively designed to form well-ordered structures for biomedical

applications since 1990s. They were brought to the attention of scientists mainly because

of their noticeable advantages in biodegradability and low toxicity to the biological

entities. The self-assembly process can lead to the formation of different nanostructural

architectures such as nanotubes and ribbons. It can also further develop into fibers or

hydrogels [75]. The unique properties of the formed structure can provide different

applications in diverse fields especially in tissue engineering and drug delivery.

Meanwhile, due to their similarity in the resulting self-assembly structures to amyloid

proteins, some of them were studied as a simpler substitute for amyloid protein (40-42

amino acids) as a way to find inhibitors to prevent fibrillogenesis in neurological

disorders [76, 77].

These self-assembling peptide amphiphiles can serve as novel materials for tissue

engineering mostly due to the hierarchical supra-molecular structures they generate can

resemble biological ECM and interact with cells at the molecular level. They are

exploited to create scaffolds for drug delivery basically owing to the unique

characteristics of the matrix geometry formed after self-assembly and their non-covalent

interactions with the drug payload. In addition, the triggered self-assembly feature of

these peptide amphiphiles can provide a ‘switch’ function which means the formation of

insoluble structures (typically fibrous networks) can be induced by the changes in their

Page 43: Self-Assembling Bioaffinity Hydrogels for Localized

27

external environment. This makes it possible for self-assembly to take place in response

to physiologically relevant changes through rational design of the peptide composition.

Considerable studies have been done on three types of self-assembling peptide

amphiphiles, namely, ionic-complementary self-assembling peptides, self-assembling β-

hairpins and alkylated peptide amphiphiles. Each type possesses distinct peptide

sequence composition and/or structural organization. The ionic-complementary self-

assembling peptide consists of regular repeats of alternating hydrophobic and hydrophilic

amino acid residues. The hydrophilic residues contain both positively charged and

negatively charged amino acids. These charged residues exhibit a self-complementary

pattern of charge when the peptides are aligned in a β-sheet conformation. By

comparison, self-assembling β-hairpins are peptide amphiphiles comprising a central

tetrapeptide flanked by two extended strands. The central tetrapeptide has high β-turn

propensity, while the two strands are composed of alternating hydrophilic and

hydrophobic residues, primarily lysine and valine respectively. In contrast, alkylated

peptide amphiphiles are made of a long linear hydrophobic alkyl tail attached to a

polypeptide block that includes β-sheet-forming segments, charged residues, and

biological epitopes, etc. Although displaying distinct molecular organization and

adopting slightly different mechanisms during self-assembly, these peptide materials do

share common features in terms of self-assembled macrostructures and functional

applications. To facilitate a comprehensive understanding of peptide amphiphiles and

meanwhile identify their similarities and differences, the characteristics, self-assembling

mechanisms and applications for each peptide amphiphile will be summarized and

compared in this chapter.

Page 44: Self-Assembling Bioaffinity Hydrogels for Localized

28

Analytical techniques in characterizing the assemblies

In order to prepare a detailed discussion about the properties of each peptide

amphiphile, it is necessary to summarize the analytical methods commonly used to

characterize structures and properties of peptide-based materials. These techniques

usually offer insight into the secondary structures of the peptide in the self-assembled

state, the nanoscale morphology of the assembled structures and the bulk mechanical

properties of the resulting biomaterials.

Fluorescent spectroscopy

Fluorescence is the property of some substances to emit light at a longer

wavelength shortly after they absorb light at a certain wavelength [78]. Fluorescent

spectroscopy has been one of the most powerful methods to study protein folding,

dynamics, assembly and interactions due to its high signal-to-noise ratio and small

sample amount required for analysis. Extrinsic fluorescent dyes such as Thioflavin T,

ANS, Bis-ANS, and Nile Red are widely used as tools for protein characterization

because of their versatility, sensitivity and suitability for high-throughput screening [79].

The fluorescence emission of these dyes will change after interacting with protein

structures. This interaction will induce a change on the preferred relaxation pathway of

the dye molecule, therefore exhibiting a distinct fluorescent spectroscopy profile as

compared to the unbound dye molecule. This detected change in fluorescent

spectroscopy profile can be indicative of the presence of certain protein structures.

Among those dyes, Thioflavin T is commonly used in the field of fibrillation and

amyloid fibril characterization [80]. For Thioflavin T, its fluorescence properties can be

Page 45: Self-Assembling Bioaffinity Hydrogels for Localized

29

affected by the solvent viscosity and the rigidity of its microenvironment. In solvent of

low viscosity, Thioflavin T has low fluorescence. In the presence of amyloid aggregates,

Thioflavin T exhibits an additional absorption peak at 450 nm and a highly fluorescent

emission maximum at 480 nm [72]. The mechanism for the change in emission spectra

of Thioflavin T is still not very clear. Some propose that Thioflavin T interacts with β-

sheet structures [81] while others attribute the fluorescence increase to the formation of

Thioflavin T micelles which bind to the grooves of the twisted protofibrils or fibrils [82].

Researchers have been using the unique feature of thioflavin T to semi-quantify the

formed β-sheet fibril structures [83]. Given that peptide amphiphiles have been expected

to form similar fibrillar structures to amyloids, thioflavin T could be utilized to

characterize and quantify the self-assembled structures formed by these peptide

amphiphiles.

It is worth mentioning that Congo red is a dye frequently used in combination

with Thioflavin T for fibril analysis. When Congo red interacts with fibrils, the changes

lie in the shifts of UV absorbance (from 490 nm to 540 nm) and other optical effects such

as birefringence under plane polarized light, instead of the increase in fluorescent

intensity. Moreover, with the help of Congo Red, the invisible and insoluble fibrillar

structures could be visualized under bright field microscopy, which facilitates the

recognition of amyloid proteins and study of the macrostructure formed by self-

assembling peptides [84].

Atomic force microscopy (AFM)

Page 46: Self-Assembling Bioaffinity Hydrogels for Localized

30

AFM is a type of scanning probe microscopy with relatively high resolution. It

can serve as an important tool to study the local nanostructure or morphology of the self-

assembled materials because it can gather information by probing the surface of the

sample with a mechanical cantilever. In contrast to other popular microscopic

techniques, the outstanding signal-to-noise ratio of the AFM topograph and 3D AFM

images allows the observation of structural details of these assemblies [85-89].

Compared to scanning electron microscope (SEM), AFM provides superior topographic

contrast as well as the direct measurement of surface features such as quantitative height

information [90]. In addition, the AFM samples could be imaged in their hydrated state

and no dehydration of the sample is necessary. This renders it a non-destructive

technique. However, its capability to accurately measure the XY dimension which

defines the width of the fiber is limited [75].

Circular dichroism (CD)

Circular dichroism (CD) is defined as the phenomenon of the unequal absorption

of left-handed and right-handed circularly polarized light by some materials [91]. It is

known that any linearly polarized light wave can be obtained as a superposition of a left

circularly polarized and a right circularly polarized light waves (with electric fields

rotating clockwise ER and counterclockwise EL respectively) when their amplitude is

identical. Therefore, if linearly polarized light traverses a substance exhibiting circular

dichroism, its properties will change since the substance could absorb the two circularly

polarized components to different extents. CD is either reported in units of ∆E (the

difference in absorbance of ER and EL by an asymmetric molecule) or in degree ellipticity

(the angle whose tangent is the ratio of the minor to the major axis of the ellipse). CD is

Page 47: Self-Assembling Bioaffinity Hydrogels for Localized

31

considered as an excellent tool to rapidly determine the secondary structure or folding

properties of proteins or peptides [92-94]. Different structures of proteins or peptides

possess characteristic CD signature spectra due to the varying aligned patterns of the

chromophores from the amides of the polypeptide backbone of the proteins. For

instance, α-helix has negative bands at 222 nm and 208 nm and a positive peak at 193

nm. Well-defined β sheet has a negative band at 218 nm and positive peak at 195 nm.

Disordered proteins have very negative bands near 195 nm and very low ellipticity above

210 nm. Figure 2-1 shows the CD spectra for some representative secondary structures

in polypeptides and proteins. One of the biggest advantages of CD is that the data can be

collected and analyzed within a few hours. Samples containing only 20 ug or less of

protein in aqueous buffers under physiological conditions can be analyzed. This makes

CD the most frequently used method for studying secondary structures of peptides and

proteins.

Page 48: Self-Assembling Bioaffinity Hydrogels for Localized

32

Figure 2-1 CD spectra for polypeptides and proteins with representative secondary

structures (reprinted from reference [91]).

Fourier transform infrared (FTIR) spectroscopy

FTIR spectroscopy is another commonly used technique to study the secondary

structure of peptides in the self-assembled state. IR is a vibrational spectroscopy with

spectra observed in the wavelength range between 1 and 100 µm (102-104 cm-1).

Infrared absorption comes from a changing dipole moment in the molecules during the

vibration of interest. Therefore, it is sensitive to the changes in inter-nuclear distances

and bond angles which are decided by the conformations and configurations adopted by

the particular molecular groupings [95]. Fourier transform is a mathematical process

required to convert the raw absorption data into the actual spectrum. A major advantage

of FTIR spectroscopy is that its structural characterization does not rely on the physical

state of the sample. Samples could be easily examined in the form of aqueous or organic

solutions, hydrated gels, in homogeneous dispersions or solids. Also, it is not limited by

protein size (can be very big) or the nature of the environment. It is especially valuable

when a peptide or protein is existing in a rapidly inter-converting dynamic equilibrium

[96]. There is an empirical correlation between the frequency on FTIR spectroscopy of

the amide I and amide II absorptions of a protein and the predominant secondary

structural motif within the protein. The amide I absorption contains mainly the

contributions from the C=O stretching vibration of the amide group while amide II arises

from N-H bending and C-N stretching vibrations. Comparatively, amide I absorption is

considered to be more useful since it is purer in its composition. This empirical

correlation in amide I frequencies and various secondary structure can be explained by

Page 49: Self-Assembling Bioaffinity Hydrogels for Localized

33

dihedral angles, hydrogen bond strengths and C=O group electron density. Briefly, the

dihedral angles of a polypeptide chain determine the chain geometry, thereby dictating

the length and direction of the hydrogen bonds. Different resulting hydrogen bond

strengths further contribute to the different characteristic electron densities in the amide

C=O groups. The stronger the hydrogen bond, the lower the electron density in the C=O

group thus the lower the amide I absorption appears. As a result, while CD is sensitive to

the backbone information, FTIR is more sensitive to the degree and strength of hydrogen

bonding to the amide C=O groups. The following table (Table 2-1) provides the most

common correlations between protein secondary structure and the amide I frequency.

Table 2-1 The correlations between common protein structures and amide I frequencies

(adapted with permission from reference [97]. Copyright (1992) American Chemical

Society)

Structure Amide I frequency (cm-1)

Antiparallel β-sheet 1675-1695

α-helix 1648-1660

Unordered 1640-1648

β-sheet 1625-1640

Aggregated strands 1610-1628

Oscillatory rheology

Oscillatory rheology is a technique used to measure the viscoelastic behavior of

soft materials such as emulsions, foams, colloidal suspensions or polymer systems which

have mechanical properties that lie between a purely elastic solid and a viscous liquid

[98]. Many researchers use it to study the mechanical properties of hydrogels made of

Page 50: Self-Assembling Bioaffinity Hydrogels for Localized

34

self-assembling peptides [87, 93, 99]. An oscillatory rheometer is the instrument

designed to make measurement of these viscoelastic properties. The principle of the

instrument is to measure the resultant stress response of the material after a sinusoidal

shear deformation is applied to the sample. The viscoelastic behavior of the sample at

certain frequency of oscillation (ω) is characterized by the storage modulus G’(ω), and

the loss modulus G’’(ω). These two moduli characterize the solid-like and fluid-like

contribution to the measured stress response respectively [98]. Knowledge of the

mechanical properties of these self-assembled peptide materials through oscillatory

rheology measurement allows rational design of the peptides with appropriate mechanical

strength for specific applications in biomedical engineering or nanotechnology.

Surface tension measurement

Surface tension measurement is commonly employed to study the solution

properties of molecules with an amphiphilic structure. Axisymmetric drop shape

analysis-profile (ADSA-P) is a powerful and reliable technique to acquire surface tension

at the interface between water and air [100]. Technically, in order to get surface tension

information for a peptide amphiphile solution, a pendant drop of the solution can be

formed at the tip of a vertical Teflon needle which is connected to a motor-driven

microsyringe. The sample should be placed in a temperature-controlled environmental

chamber which is saturated with water vapor. The image of the pendant drop can be

acquired by an optical microscope and a CCD camera and then transferred to a computer

for software analysis. The surface tension value is finally generated by fitting a

theoretical curve governed by the Laplace equation of capillarity to the profile of the

pendant drop [76]. Hydrophobicity, surface activity of the peptide and kinetics of peptide

Page 51: Self-Assembling Bioaffinity Hydrogels for Localized

35

adsorption can be quantified by applying ADSA-P measurement. The conformation of

the peptide assemblies or the aggregation status can be related accordingly [101]. By

monitoring the surface tension change under various conditions or over time, we can get

a better understanding about the self-assembly process and thus speculating the self-

assembly mechanism.

Other analytical techniques

In addition to the frequently employed methods listed above, there are still a lot of

other techniques which are less commonly used due to inconvenient sample preparation,

high cost or inaccurate information. These techniques include solid-state nuclear

magnetic resonance (ssNMR) [102-104], X-ray diffraction (XRD) [104], molecular

dynamics (MD) simulations [92], scanning electron microscopy (SEM)[87] and

transmission electron microscopy (TEM) [99, 105] et al. Some of them, especially

ssNMR and XRD, can provide us with even higher resolution and more precise

information down to molecular or atomic level. They are increasingly recognized as

more useful and accurate tools to study fibril structure although high cost and limited

availability may constrain their wide application. It should be noted that, when studying

these peptide amphiphiles, multi-technique approach is required to generate a more

comprehensive and accurate picture of the self-assembly structure.

Self-assembling peptide amphiphiles and their corresponding properties

I. Ionic-complementary self-assembling peptide

Page 52: Self-Assembling Bioaffinity Hydrogels for Localized

36

Zhang et al. from Massachusetts Institute of Technology discovered the ionic-

complementary self-assembling oligopeptides [106]. These peptides are composed of

alternating hydrophilic and hydrophobic amino acid residues and tend to adopt β-sheet

conformation in solution. The hydrophilic amino acids contain both positively charged

and negatively charged residues. They are able to form complementary ionic pairs in

secondary or higher order of structures. This class of self-complementary oligopeptides

can spontaneously assemble to form large macroscopic structures/matrices when exposed

to environmental stimuli such as temperature, pH, and monovalent inorganic salts.

The first discovered self-assembling peptide of this type is EAK16-II (sequence:

Ala-Glu-Ala-Glu-Ala-Lys-Ala-Lys-Ala-Glu-Ala-Glu-Ala-Lys-Ala-Lys) (structure see

Figure 2-2). It was originally found in a region of zuotin, a putative left-handed Z-DNA

binding protein in yeast [107]. Because of the adoption of β-sheet conformation in

solution, this oligopeptide molecule can present hydrophobic alanine side chains on one

side while orienting self-complementary pairs of positively charged lysine and negatively

charged glutamic acid side chains on the other surface. It can form membrane-like

structures upon addition of monovalent metal ions. The assembly is considered to be

facilitated by side chain ionic bond interactions between glutamic acids and lysines as

well as the hydrophobic interactions among alanines in addition to the conventional β-

sheet backbone hydrogen bonding (Figure 2-3). The formed membrane has high

stability. It is resistant to protease digestion, high temperature and a wide range of pH

values (1.5-11). It is also mechanically stable and can be stained by Congo red, a dye

that preferentially stains β-pleated sheet structures. SEM observation revealed that the

membrane was composed of a fibrous structure consisting of interwoven individual

Page 53: Self-Assembling Bioaffinity Hydrogels for Localized

37

filaments (Figure 2-4) [23]. It was later found that those interwoven filaments have

diameters of 10-20nm and porous enclosures of about 50-100nm in diameter [24].

Figure 2-2 Schematic three-dimensional molecular model of EAK16-II (carbon atoms are

cyan, oxygen atoms are red, nitrogen atoms are blue and hydrogen atoms are white, A, E

and K are single letter codes for amino acids alanine, glutamic acid and lysine

respectively, reprinted from reference [108])

Page 54: Self-Assembling Bioaffinity Hydrogels for Localized

38

Figure 2-3 Initially proposed model of the membrane by Zhang et al. (a) view

perpendicular to the β-sheet; (b) stacking of β-sheets (reprinted from reference [23],

Copyright (1993) National Academy of Sciences, U.S.A.)

Page 55: Self-Assembling Bioaffinity Hydrogels for Localized

39

Figure 2-4 SEM images of membranes formed by EAK16-II with serial magnification.

The diameter of the filaments are about 10-20 nm and the distance between fibers are

about 50-80 nm. (a, 300X, b, 1200X, c, 4500X, d, 15,000X, arrows mark the same

location. Reprinted from reference [23], Copyright (1993) National Academy of

Sciences, U.S.A.)

Later, EAK16-II was systematically altered to generate other ionic

complementary peptides, such as RAD16, EFK16, KLD12 and EAR8 etc, which could

also self-assemble through similar mechanisms. Presumably, any hydrophobic amino

acid residue, such as valine (V), leucine (L), isoleucine (I) and phenylalanine (F) can be

used to substitute alanine (A). Likewise, on the hydrophilic side, arginine (R), histidine

(H) as well as lysine (K) can be included as positively charged residues while aspartic

acid (D) and glutamic acid (E) are commonly used negatively charged residues in the

alternating sequence. There are usually 8-32 amino acids in the short peptides. For

easier reference, the hydrophilic sides have been classified into several moduli according

to their varied charge distribution patterns: modulus I (-+-+-+-+); modulus II (--++--++);

modulus III (---+++); modulus IV (----++++) [109]. Here, the roman numerals indicate

the number of the same kind of charges grouped together. The total number of amino

acid residues in the peptide is indicated in the name as well. For example, EAK16-II has

Page 56: Self-Assembling Bioaffinity Hydrogels for Localized

40

16 amino acid residues and a charge distribution pattern of modulus II (--++--++).

Notably, these oligopeptides all consist of L-amino acids. Therefore, their degradation

will only yield normal amino acids which will not arouse any biocompatibility problem

in vivo [110].

This class of ionic-complementary self-assembling peptides has also attracted

significant interest from P. Chen group from Waterloo University in Canada. His group

set forth to investigate various factors that influence the self-assembling process of these

oligopeptides. Those factors include peptide concentration, pH, salt concentration,

charge distribution, etc. Their findings, together with data collected by Zhang’s group,

accelerated our knowledge about these peptides and provided us with critical information

to speculate the self-assembling mechanisms.

Self-assembling mechanisms and factors affecting the self-assembling

It was initially postulated by Zhang et al. that the β-sheet system formed by

EAK16-II may exist in two possible forms [111]. One is a single β-pleated sheet

containing many EAK16 molecules. In the middle of the single sheet, each amino acid

residue is held to its neighboring peptides by two hydrogen bonds. Meanwhile, on one

side of the sheet there is a collection of alanine side chains forming a hydrophobic

domain while on the other side of the sheet there are a number of ionic bonds between

lysine side chains on one peptide strand and glutamic acid side chains on another. The

other existing form could be two or more sheets assembled together. Specifically, there

would be layered structure in which the sheets are held together alternatively by

hydrophobic interactions and ionic interactions between the sheets (Figure 2-3b).

Page 57: Self-Assembling Bioaffinity Hydrogels for Localized

41

Without doubt, the hydrogen bonds, hydrophobic interactions and ionic interactions

stabilizing the single sheet still exist in this form.

It was experimentally proven that different conditions such as amino acid

sequence, peptide concentration, salt concentration, pH, etc. can influence the peptide

aggregation process and further development of macrostructures. Therefore, it is of

importance to investigate the influential factors in order to understand the self-assembling

mechanisms for the amphiphilic oligopeptides.

The amino acid sequence of these peptides is critical to the self-assembling

mechanism. Various sequences of the EAK family repeats such as KFE12, KIE12,

KVE12, KFE8/16 KFQ12 were designed by Caplan et al. to study the sequence influence

[112] (sequences see Table 2-2). Three aspects were taken into consideration: 1) the

hydrophobic side chains; 2) the number of repeats; and 3) the charged side chains. The

results of this study showed that increasing the hydrophobicity of the nonpolar resides

would facilitate the self-assembling which is reflected in the decrease of the critical

coagulation concentration (CCC) after increasing the hydrophobicity of the nonpolar

residues. CCC is defined as the salt concentration below which the oligopeptide does not

form a gel and above which it does form a gel. Therefore, lower salt concentration is

required to drive the gel formation if the nonpolar residues are more hydrophobic. The

phenomenon that increasing the hydrophobicity of the nonpolar residues contributes to

matrix formation was also supported by another study. It has been reported that alanine-

containing amphiphilic peptides (such as EAK) require at least sixteen-mer peptides for

salt-induced stable matrix formation. By contrast, leucine-containing amphiphilic

Page 58: Self-Assembling Bioaffinity Hydrogels for Localized

42

peptides (such as ELK) form salt-induced stable matrices from eight-mer peptides [113].

Wang et al. found that, for aliphatic residues, not only the hydrophobicity, but also the

size of these residues could introduce different steric topologies to the non-polar side of

the β-strand, thus affecting the matrix morphology of self-assembled structure [114].

Specifically, if different non-polar residues are incorporated into one peptide sequence,

diverse topologies of the β-strand non-polar side can be formed by modulating the

hydrophobic residue combinations of glycine, alanine, valine, leucine, isoleucine, and

phenylalanine. It was proven that steric size differences of non-polar residues can cause

variations in peptide secondary structures and complementary non-polar topologies

encourage cooperative binding between amphiphilic peptides which would further

facilitate matrix formation.

Table 2-2 amino acid sequences studies by Caplan et al. (reprinted from [112])

As to the number of repeats, it was found that increasing the length of the

oligopeptide will increase the intermolecular attraction as well as the entropy of the

backbone. Since self-assembly is a process facilitating intermolecular attraction while

also decreasing peptide backbone entropy, the competition of these two opposite effects

to the thermodynamic process would make the final CCC have a biphasic trend (i.e.

minimum CCC exists at intermediate length).

Page 59: Self-Assembling Bioaffinity Hydrogels for Localized

43

In addition, the charged residues in the sequence can decide the charge

distribution and overall charge carrying in the peptide molecules. Charge distribution can

make a difference to the self-assembling process even when the same amino acid

composition is concerned. In particular, different types of EAK16 (EAK16-I, EAK16-II

and EAK16-IV) with the same amino acid composition but varying charge distributions

were studied [108]. The difference in charge distributions can lead to distinct favored

conformations for each molecule. This will further affect the final self-assembly

structure observed by AFM. It was demonstrated that EAK16-I and -II favors relatively

stretched conformation and forms a fibrillar self-assembly structure while EAK16-IV

favors a β-turn conformation and forms globular self-assembly structure. On the other

hand, the overall charge carried by the peptide can be manipulated by substituting the

charged residues with other uncharged residues. This allows the control of the conditions

(especially pH) under which the oligopeptides can gel. Peptides have a higher propensity

to self-assemble when there are no repulsive charge forces between peptide molecules.

Therefore, pH can take effect by changing the charge distribution/status of the peptide

sequence [115]. For peptides having a positive overall charge at neutral pH, increasing

the environmental pH can initiate the self-assembling whereas for negatively charged

peptides at neutral pH, decreasing pH helps to protonate the negative charge, bringing the

overall charge to zero. Hence, different sequences may display totally opposite trends in

conformational change and self-assembly process when exposed to a pH change in the

solution as evidenced by M. Altman et al. [116].

Concentration is also a key factor controlling the self-assembling process. It was

proposed by Fung et al. [101] that oligopeptides like EAK16-II aggregate into

Page 60: Self-Assembling Bioaffinity Hydrogels for Localized

44

protofibrils through seeding and/or a nucleation process depending on its critical

aggregation concentration (CAC). Below CAC (~0.1 mg/ml), two forms of nanostructure

of protofibrils can be adopted, i.e. the thin filament and globular aggregates which were

postulated to consist of monolayers of β-sheets and at least 2 layers of β-sheets

respectively. Over time these protofibrils can further aggregate into fibrils at

concentration above CAC and these fibrils are composed of many globular aggregates

lining up and stacking together. It is worth noting that incubation time is another

important factor to consider in order to understand the kinetics of aggregation.

Oligopeptides at low concentration may start to aggregate at prolonged incubation time

under the condition in which it will not normally self-assemble.

The effect of salt is of particular importance because salt concentration is

physiologically relevant and is a most frequently utilized trigger to initiate the self-

assembling. They can affect the aggregation process and resulting aggregation

morphologies by electronically screening ionic interactions during the molecular self-

assembling. However, it was also found by Hong et al. [117] that the self-assembling

process of EAK16-II to form nanostructures does not strongly depend on the presence of

NaCl (i.e. electrostatic interactions) when the peptide concentration is high (>0.3 mg/ml).

On the other hand, the dimensions of the resulting self-assembled nanofibers at low

peptide concentrations (<0.1 mg/ml) appear to be dependent on NaCl concentration. The

size of the nanofibers increases until the NaCl concentration reaches about 20 mM. The

radius of the peptide fibrils decreases with further increase in salt concentration. Besides,

globular assemblies were observed in the absence of NaCl at low peptide concentration

but not observed when the electrolyte is present. This indicates that NaCl promotes

Page 61: Self-Assembling Bioaffinity Hydrogels for Localized

45

formation of fibril at low peptide concentration. It was thus speculated that the charge

screening effect of the salt promotes non-electrostatic interactions between the peptide

molecules.

In summary, consensus has been reached on the theory that these peptides

assemble in water due to the hydrophobic effect and the charged polar side chains also

play important roles in determining what conditions promote assembly and what

conformation the aggregates will preferentially adopt [118]. Specifically, the

hydrophobic effect facilitates self-assembling through the entropy increase of water

molecules in the system. On the other hand, like-charge repulsion from the charged

residuals prevent self-assembling from happening. The superposition of hydrophobic

interactions and electrostatic repulsion decide the existence state of the peptide in a

certain media. However, most of current mechanism studies are still limited to these

qualitative conclusions summarized from phenomenon observed under varying

conditions. Few publications look deep into the assembled structures down to the

molecular level and explain the process from the physicochemical or thermodynamic

perspective. Systematic research is still lacking for the big ionic self-complimentary

peptide family since most studies are concentrated on EAK16 and RAD16. More efforts

need to be invested to obtain a broad understanding about this type of peptide, thereby

generating more reliable self-assembling mechanisms.

Page 62: Self-Assembling Bioaffinity Hydrogels for Localized

46

Characteristics and applications

Having explored the self-assembling mechanisms for ionic complementary

oligopeptides, the features of the resulting self-assembly structures and their potential

applications will be discussed.

Mechanical strength of the hydrogel scaffolds is a critical property in deciding

their potential to have a satisfactory manner within the biological environment in which

they are applied [119]. It was found that higher concentration of peptide usually leads to

higher density of the fiber network which further contributes to the increase in gel

stiffness. Also, crosslinking can be an effective strategy to increase the stiffness of self-

assembling oligopeptide gels. It can be achieved, for example, by introducing cysteine

into the sequence to allow disulfide bond formation or adding glutaraldehyde to crosslink

amine groups in the residue side chains [120]. With a thorough understanding of

fundamental mechanisms governing the mechanical properties of these gels, we can

custom-design a gel scaffold with desirable stiffness for a specific biomedical

application.

It is worth noting that the well-defined nanofiber scaffold formed by ionic self-

complementary peptides could reassemble after undergoing certain types of mechanical

breakage. This is evidenced by the dynamic reassembly of RAD16-I [(RADARADA)2]

after disrupted by sonication. The smaller fragments resulting from sonication are able to

quickly reassemble into nanofibers which were indistinguishable from the original self-

assembly [120, 121]. This feature is of great value because the self-assembly scaffold

Page 63: Self-Assembling Bioaffinity Hydrogels for Localized

47

will be subject to certain mechanical strength when applied to an in vitro or in vivo

environment.

Chirality of the amino acids could also play some role for the final properties of

the self-assembly structure. It is already known that natural proteins are made of only L-

amino acids. Using the chiral counterpart –D-amino acids to generate a chiral mirror

image peptide (such as d-EAK16-II) can make the self-assembled structure more stable

and better resistant to natural enzyme degradation since the chiral peptide cannot be

recognized by natural proteases and is thus more durable [84, 122].

With so many superior features being displayed, these self-assembling

oligopeptides have been demonstrated to have the potential for diverse biomedical

applications including scaffolds for tissue repair, tissue regeneration, instant stopping of

bleeding, wound healing, drug delivery and biological surface engineering, etc. These

applications arise from their capability to self-assemble into a three-dimensional matrix

scaffold in situ under physiological conditions as well as their potential to resorb into

biocompatible degradation products.

With regard to tissue engineering, self-assembling peptide scaffolds have been

proven to support cell attachment and differentiation of plenty of mammalian primary

and transformed cells [24, 123-125]. For example, it was noted that scaffold generated

by RAD16-II peptides provided cells with a soft three-dimensional support which can

facilitate angiogenesis and allow them to form capillary-like structures [126]. This

capillary morphogenesis and endothelial cell adhesion can be affected by primary

sequence of the self-assembling peptide considering the fact that KFE8 and KLD12

Page 64: Self-Assembling Bioaffinity Hydrogels for Localized

48

cannot produce this effect while RAD16-I and RAD16-II can [127]. Moreover, the self-

assembled structure can serve as substrates for extensive neurite outgrowth and active

synapse formation in addition to supporting neuronal cell attachment and differentiation.

It did not elicit any measurable immune response or tissue inflammation when introduced

to animals [128]. It was also reported that hydrogel scaffold made from RAD16-I could

be used for instant stopping of bleeding when applied on top of multiple tissues and a

variety of wounds [115]. The results showed that the complete hemostasis can be

achieved in less than 20 seconds. The hemostatic process is directly proportional to the

length of the self-assembled nanofibers: the longer the nanofiber, the shorter the duration

of bleeding [129]. Furthermore, the nanofiber scaffolds with favorable designed

properties could be used in in-vitro 3-dimensional cell culture which mimics the

extracellular matrix (ECM) in providing support and allowing the natural flow of oxygen,

nutrients and growth factors.

In addition, functional motifs such as RGD (Arg-Gly-Asp, for cell adhesion),

BMHP (bone marrow homing peptide) [130], YIGSR and RYVVLPR (two sequences

present in laminin-1 to promote cell adhesion, migration and endothelial cell tubular

formation), and TAGSCLRKFSTM (sequence from collagen IV to bind to heparin and

promote adhesion and spreading of bovine aortic endothelial cells) [131], can be

incorporated in the peptide sequence by direct solid phase synthesis extension at peptide

termini in order to be displayed in the gel matrix for biologically active applications. In

fact, sequences with desired biological functions can be incorporated into the self-

assembling peptide with varied insertion patterns, not necessarily on the two termini. For

example, a metalloproteinase (MMP-2)-sensitive substrate can be inserted centrically

Page 65: Self-Assembling Bioaffinity Hydrogels for Localized

49

within RADA blocks with 3 RADA units flanking on each side. The hydrogel matrix

formed by this novel peptide has comparable mechanical properties and nanofiber

morphology to material built with RAD16-I alone. Exposure of the new gel to MMP-2

can lead to peptide cleavage, mimicking the remolding of natural ECM during tissue

engineering [132]. Notably, both regular and D-form (chiral counterpart) self-assembling

peptide systems could be employed to form the matrix and promote cell viability and

functionality through rational design [133]. It is encouraging that the oligopeptide

RAD16-I has already been developed as a product under the brand name PuraMatrixTM

for multiple biomedical applications [134, 135], mainly as 3D cell culture matrix.

Ionic-complementary self-assembling peptides have also drawn a lot of attention

in controlled drug delivery field due to their desirable properties. For example, they can

be easily engineered to form a variety of stable nanostructures such as fibers, rods, tubes,

nanovesicles and globules and these structures have great potential to be used as delivery

vehicles [136]. Since there are two sides on the β-sheets formed through backbone

hydrogen bonding: hydrophobic side and hydrophilic side, this class of peptide

amphiphile possesses a superior capability to incorporate both polar and nonpolar drugs

in the matrix as a drug delivery platform. A wealth of successful examples especially for

hydrogels made of EAK16-II or RAD16-I have been found for the delivery of

hydrophobic drugs [26, 136] as well as functional proteins [30], cytokines [137], growth

factors [138] and antibodies [139]. The release profiles of these molecules are usually

controlled by the self-assembling peptide concentration (nanofiber density) as well as the

electrostatic interaction between the fiber surface and diffusant molecules [140]. The

release will decrease with increasing peptide concentration (physical hindrances) and

Page 66: Self-Assembling Bioaffinity Hydrogels for Localized

50

stronger charge-induced interactions. Thus, the release kinetics can be tailored

accordingly by changing the peptide composition and concentration. It was also

demonstrated that the conformation and functionality of the proteins or antibodies will

not be affected by the peptide matrix encapsulation and the loading efficiency can reach

as high as 100% since the peptide hydrogel consists of water up to 99.5% [139].

Moreover, the self-assembling peptides may have the potential to form complexations

with some hydrophobic drugs in aqueous solution (e.g. EAR8-II with anticancer drug

pirarubicin) [141] or allow the drugs to present as different molecular states in the carrier:

protonated and crystalline (such as EAK16-II with drug ellipticine), resulting in different

therapeutic activities [27, 136]. It is also worth noting that hydrogels composed of more

than one species of peptide have been designed for programmable sustained release of

cargo. For example, RAD16-I and KLD12-I were paired to make a two-layered hydrogel

structure for human IgG release with a concentric sphere of RAD16-I as the core and

KLD12-I as the shell [139]. Controlled release of the antibody can be fine-tuned by this

system.

In the Meng lab, we have been working on EAK16-II and EAK16-II with

appended functional groups such as EAKIIH6 [19, 20] (with six consecutive histidines at

the C-terminus) and dL5_EAK (with a fluorogen activating protein dL5 attached to the

N-terminus) [50]. All of them can co-assemble with EAK16-II to form an insoluble

bioaffinity hydrogel, thus presenting the functional groups at local site of injection for an

extended period of time. For EAKIIH6, with the help of intermediate protein adaptors

(anti-H6 antibody and protein A/G), any antibody of interest could be displayed locally

by the co-assembly structure for immune modulation purposes. Its effectiveness has been

Page 67: Self-Assembling Bioaffinity Hydrogels for Localized

51

proven in vivo using two epithelial tumor cells. As to dL5_EAK, due to the fluorescence

enhancement function of the dL5 domain upon specific binding to the dye malachite

green (MG), it has a potential to be used as an in situ forming local biosensor. In both

cases, EAK16-II functions to immobilize and concentrate the functional group in situ.

Other than applications in tissue engineering and drug delivery scaffold, the self-

assembling peptides have also been used in electronics field as template for metallic

nanowire and electronic device fabrication. For example, functional Cu2+ binding group

GGH (G, glycine; H, histidine) was attached to the C-terminus of EAK16-II by H. Yang

et al. to create a bi-functional material which can form stable nanofibers and at the same

time incorporate copper into the structure via a metal binding site in the self-assembling

peptide [142].

Additionally, these peptides may serve as a good model system for studying the

mechanism of many neurodegenerative diseases such as Alzheimer’s and Parkinson’s

diseases considering that the fibrillar nanostructures resulting from these peptide

assemblies are similar to the amyloid fibrils found in those diseases. Therefore, the

mechanism or theoretical studies for this self-assembling system will give us insight into

the driving forces for the assembly of more complex endogenous proteins such as β-

amyloid and prion proteins, thereby facilitating the research trying to find the treatment

that will prevent plaque formation in vivo.

Page 68: Self-Assembling Bioaffinity Hydrogels for Localized

52

II. Self-assembling β-hairpins

Pochan, Schneider and coworkers in Delaware developed a type of peptide that

can have an intramolecular folding event before undergoing intermolecular self-assembly

into hydrogel scaffolds. This intramolecular folding can be triggered by different kinds

of stimuli such as changes in pH [143], temperature [144], ionic strength [145], light

[146] or the introduction of cell culture media [147]. This class of peptide amphiphile is

composed of a central tetrapeptide (VDPPT) which has high β-turn propensity flanked by

two extended strands containing high β-sheet propensity residues. The first peptide

designed by this group is named MAX1 (structure see Figure 2-5). The strands contain

alternating hydrophilic and hydrophobic residues, mainly lysine (K) and valine (V)

respectively. Interestingly, these strands have a similar pattern to the ionic-

complementary self-assembling peptides discussed above in terms of amino acid

organization. The biggest difference lies in the charge state of the sequence. Here,

mainly positively charged lysine is incorporated as the hydrophilic residue while

positively charged and negatively charged amino acids were alternated in the previous

case. Therefore, the overall net charge for β-hairpins at neutral pH is positive compared

to zero for the ionic-complementary peptides. This difference can affect their

corresponding self-assembling trigger conditions. Later, derivatives of MAX1 containing

other amino acid substitutes at different positions and/or with different numbers were also

created to achieve tunable stimuli functions.

Page 69: Self-Assembling Bioaffinity Hydrogels for Localized

53

Figure 2-5 proposed β-hairpin structure and amino acid sequence of MAX 1, a typical

peptide of this category (20 residues). (reprinted with permission from reference [145].

Copyright (2004) American Chemical Society)

Self-assembling mechanisms

It is not surprising that the self-assembling of β-hairpins can also be initiated by

the previously mentioned environmental cues (temperature, salt, pH, etc.) to form the

hydrogel scaffold. Many researchers have shown interest in studying the underlying

mechanisms as well. Work on different stimuli will be discussed separately.

For pH responsive hydrogels, the charge state of the hydrophilic residues plays an

important role in intramolecular folding. Since lysine is basically the hydrophilic amino

acid incorporated in the sequence of β-hairpins, increasing the pH above the intrinsic pKa

(10.8) of the lysine side chains can reduce the intrastrand charge repulsion from

neighboring lysines, thus initiating the intramolecular folding and subsequent self-

assembly to hydrogel structure. The final self-assembled structure is supported both

laterally through hydrogen bonds between adjacent hairpins and facially by hydrophobic

association of the valine-rich faces [143].

For ionic strength responsive hydrogels, the addition of salts is able to trigger the

intramolecular folding due to the formation of salt-bridge physical cross-links or charge

Page 70: Self-Assembling Bioaffinity Hydrogels for Localized

54

screening effects for the electrostatic interactions between the charged amino acids [145].

In the folded state, the single peptide molecule is facially amphiphilic, displaying all

hydrophobic residues on one face of the hairpin and hydrophilic residues on the other.

Once it is folded into a β-hairpin conformation, self-assembly can be driven by both

intra- and inter-molecular hydrogen bonding laterally and hydrophobic interactions

facially. Well-defined, bilayered fibrils can thus be formed and further developed into a

β-sheet rich hydrogel network. The final elastic property of the hydrogel can be tuned by

salt concentration. The higher the salt concentration added, the more rigid the hydrogel

formed. The network of the hydrogel is composed of dense fibrillar assemblies that are

cross-linked to each other by physical junction points. The proposed self-assembling

pathway and fibril structure for MAX1 β-hairpin molecule are shown in Figure 2-6.

Figure 2-6 (a) proposed self-assembly pathway for MAX1, (b) the structure of self-

assembled MAX1 β-hairpin molecules in a fibril and (c) TEM micrographs of MAX1

hydrogels showing the nanostructure of 0.5% wt% network, scale bar: 100 nm; insets:

magnification of MAX1 fibrils that are ≈ 3 nm in width, scale bar: 20 nm. (adapted with

Page 71: Self-Assembling Bioaffinity Hydrogels for Localized

55

permission from reference [145] and [148]. Copyright (2004) American Chemical

Society and Copyright (2007) National Academy of Sciences, U.S.A.)

For thermally responsive hydrogels, the hydrogel forms upon temperature

increase. The increase in temperature serves to dehydrate the nonpolar residues of the

unfolded peptide and the folding is hence triggered via hydrophobic collapse [144].

As to light-activated hydrogel formation, it is achieved through covalently

incorporating a photocleavable moiety into the peptide sequence. The photocleavable

moiety possesses an electrostatic feature which inhibits peptide folding. Therefore, after

exposed to light at a certain wavelength, the photocage can be released to initiate

intramolecular folding, self-assembly and finally the formation of a mechanically rigid

hydrogel [146].

It was proposed by Yucel et al. [149] that after the bilayered β-sheets are formed

through folding, the possible cross-linking mechanisms for the final hydrogelation

include (1) self-assembly defect-induced branching (i.e. kinetically trapped, imperfect

hydrophobic face collapse may lead to nucleation of a branch, see Figure 2-7) and (2)

entanglement of self-assembled fibrils. Both can lead to permanent physical cross-links.

Page 72: Self-Assembling Bioaffinity Hydrogels for Localized

56

Figure 2-7 Schematic representations for two different self-assembled cross-links: (a) a

defect-induced branch and (b) an entanglement (adapted with permission from reference

[149]. Copyright (2008) American Chemical Society)

It is worth mentioning that more than one environmental stimuli can take place

simultaneously. For example, the folding and self-assembly of the parent MAX1 peptide

can be induced by changing the solution pH to reduce the point charge of the lysine

residues, increasing the ionic strength to screen the electrostatic repulsions, or increasing

the temperature to enhance the attraction between hydrophobic residues. These changes

can work collectively to facilitate the hydrogelation process.

Since most peptides developed by this research group possess a high β-turn

propensity tetrapeptide flanked with two extended strands, predominantly an alternating

sequence of valine and lysine residues or some point substitutions with other amino acids,

the charge state of the whole strands was proven to affect the gelation kinetics of the

peptides because the electrostatic repulsion between positively charged lysine is an

important force preventing the folding and self-assembly. Therefore, the rate of the gel

formation becomes faster for point substitutions with less total charge. As a result, the

careful design of the peptide sequence can be employed to modulate the charge state and

finally the gelation kinetics of the peptides to obtain a desired profile under certain

environmental condition [150].

Characteristics and applications

Stimuli-responsive hydrogels generated from β-hairpin peptides have the potential

to be used as biomaterials due to their superior morphological and physicochemical

features. At early-time hydrogelation, peptides undergo intramolecular folding into β-

Page 73: Self-Assembling Bioaffinity Hydrogels for Localized

57

hairpins and concurrently self-assemble into branched clusters of well-defined, β-sheet

rich nanofibrils. These fibers are very uniform with monodisperse dimensions (3nm

cross section). It was postulated that the hydrogel network rigidity is boosted by the

finite interpenetration of percolating cluster with smaller clusters as well as permanent

intercluster entanglements [149].

One of the most distinguished advantages of β-hairpin hydrogels is that they can

be injected as pre-formed solids. The pre-formed solid gel can shear-thin under a proper

shear stress and consequently flow during injection; but upon removal of the stress, it

recovers back into a solid immediately. This feature allows the hydrogel system, together

with the payloads that are encapsulated during the original hydrogelation, to be delivered

via simple syringe to a targeted in vivo site and stay localized at the site of injection to

achieve the required therapeutic effect. The unwanted leakage and flow of the injected

liquid to neighboring tissue or the blood stream can be minimized by the design of higher

viscosity precursors or fast in situ gelation kinetics. The duration of shear, the shear rate

as well as the pre-shear gel rigidity all affect the gel stiffness immediately recovered after

flow and the ultimate stiffness after complete rehealing of the gel [151]. It was also

demonstrated that the shear-thin delivery process of syringe injection had little or no

effect on cell viability and three dimensional cell distribution which indicates that these

peptidic hydrogels can serve as excellent candidates for injectable therapeutic delivery

vehicles for cells [147, 151, 152].

Interestingly, the chirality of the β-hairpin peptides (MAX 1) was found to affect

the mechanical rigidity of the self-assembled hydrogels. Specifically, hydrogels

Page 74: Self-Assembling Bioaffinity Hydrogels for Localized

58

generated from enantiomeric mixtures of self-assembling β-hairpins exhibit non-additive,

synergistic enhancement in mechanical rigidity compared to gels made from either pure

enantiomer. The racemic hydrogel shows the greatest mechanical rigidity [105]. The

underlying molecular mechanism for this phenomenon is still not clear, but spectroscopy,

rheology and microscopy data suggest that it might be due to energetically favorable

interactions between the enantiomers either occurring at molecular level or hydrogel

network level or both. Therefore, biomolecular chirality has a potential to be utilized as a

tool for material design and controllably modulating the mechanical properties to the

desired level by adjusting the molar ratio of β-hairpin enantiomers.

When it comes to applications, β-hairpin self-assembling peptides are mainly used

to prepare hydrogels as injectable therapeutic delivery vehicles. Similarly to ionic

complementary peptides, a large range of drug molecules, from hydrophobic small

molecules such as curcumin [153] to hydrophilic macromolecules such as proteins [154],

could be encapsulated into the self-assembling peptide hydrogels to make an injectable

localized drug delivery system for sustained and controlled release. This is achieved by

initiating the self-assembly in the presence of drug of interest. The drug molecule

mobility within and release out of the hydrogels can be adjusted by the mesh size of the

gel, which can be modulated by altering the peptide sequence or simply varying the

concentration of the peptide. Furthermore, the electrostatic interactions between the drug

molecules and the peptides constructing the gel can affect the mobility and release of the

drug. For example, for electropositive fibrillar hydrogels, release of positively charged

and neutral molecules is controlled by the steric effect imposed by the gel network. In

Page 75: Self-Assembling Bioaffinity Hydrogels for Localized

59

contrast, negatively charged drugs interact strongly with the positively charged network,

leading to significantly slower release [155].

β-hairpin peptides self-assembly and hydrogelation can also be employed to

deliver cells through triggering the self-assembly in cell culture medium containing cells

of interest. The resulting well-controlled, homogeneous, three-dimensional encapsulation

of desired cells and the maintenance of high cell viability after the shear-thinning process

render these gels promising tools to deliver cells at target biological sites for tissue

engineering or other purposes [148].

Moreover, the electrostatic character of the gels can affect cell behavior

incorporated within. Although hydrogel networks can afford similar rheological behavior

and local network morphologies, different network electrostatic status can influence the

cell viability after encapsulation and delivery, extracellular matrix deposition, and gene

expression. For example, by comparing hydrogels made of different charge state

monomers MAX8 and HLT2 (+7 and +5 per monomer, respectively), less electropositive

HLT2 gel was found to provide a microenvironment more favorable to chondrocyte

encapsulation, delivery and phenotype maintenance. Besides, a higher cell viability was

observed for HLT2 gels compared to MAX8 gel [156].

III. Alkylated peptide amphiphiles

The Stupp research group from Northwestern University in Chicago has

developed a type of peptide amphiphile (PAs) which consist of a long linear hydrophobic

Page 76: Self-Assembling Bioaffinity Hydrogels for Localized

60

alkyl tail coupled to a polypeptide block [157]. This polypeptide block usually includes

β-sheet-forming segments, charged residues for solubility and biological epitopes, etc.

There are a large variety of possible combinations for the peptide block components.

Some only contain a β-sheet-forming segment and a charged head group while others

also include different kinds of sequences for biological signaling and functionality. In

addition, consecutive cysteine residues can be incorporated to introduce intermolecular

cross-linking for higher robustness of the supramolecular fibers and photo-cleavable

group can be employed to introduce photo-responsive property. This type of alkylated

peptide amphiphile has the potential to self-assemble into cylindrical nanofibers with

diameters on the order of a few nanometers and lengths in the micrometer scale [158].

The hydrophobic alkyl tails pack on the inside of the fiber while the peptide portions are

displayed on the surface of the fiber which makes a cylindrical micelle with an overall

conical shape from intersection [159] (see Figure 2-8). The amino acid selection and

alkyl tail modification confer nanofibers with varying morphology, surface chemistry,

and potential bioactivity [157]. The alkyl tail is usually attached to the N-terminus of the

peptide and the epitope segment is placed at the C-terminus which will be displayed on

the periphery for cell interaction after self-assembly. There are also PAs with reverse

polarity which means the alkyl tail is attached to the C-terminus and free N-terminus is

exposed to the periphery. They co-assemble with PAs with free C-terminus leading to

assemblies with unusual thermal stability when compared to assemblies composed of

only one type of PA molecules [160]. There are three main modes of induction for the

alkylated peptide amphiphiles to self-assemble into nanofibers: pH control, divalent ion

induction, and concentration induction [161].

Page 77: Self-Assembling Bioaffinity Hydrogels for Localized

61

Figure 2-8 (a) Chemical structure of one alkylated peptide amphiphile with 5 key

structure regions. Region 1: alkyl tail; region 2: four consecutive cysteine residues;

region 3: three glycine residues serving as a flexible linker; region 4: phosphorylated

charged residue; region 5: cell adhesion ligand RGD (b) Molecular model of the PA

showing the overall conical shape of the molecule going from the narrow hydrophobic

tail to the bulkier peptide region; (c) Schematic showing the self-assembly of peptide

amphiphile molecules into a cylindrical micelle; (d) Vitreous ice cryo-TEM image of the

fibers reveals the diameter of the fibers to be 7.6 ± 1 nm (adapted from reference [159])

Self-assembling mechanisms

Since there are usually more than three distinct components in one PA molecule

(such as alkyl tail, β-sheet-forming segment, tetra-cysteine region, charged residues and

biological epitopes), studying the effect of each portion to the self-assembling

phenomenon is essential for the comprehensive understanding of the self-assembling

mechanism. In order to know what components of the peptide amphiphile are

responsible for their self-assembly, different derivatives of the PAs were designed by

Hartgerink et al. [157] and three main regions of the molecules were modified: the length

of the alkyl tail (from 22 carbons to no carbon), the tetra-cysteine region (cross-linking

effect of thiol groups) and the head group region (the C-terminal end which can interact

Page 78: Self-Assembling Bioaffinity Hydrogels for Localized

62

with the environment and incorporate different cell adhesion ligands or crystal nucleation

centers). It was found that hydrophobic interactions between the alkyl tails play a key

role in self-assembly in water given the fact that the PAs without an alkyl tail cannot self-

assemble and PAs with sufficiently long alkyl tails can self-assemble more easily.

Moreover, the presence of cysteine residues is not required for self-assembly although it

can provide covalent capture of the fibers by the formation of intermolecular disulfide

bonds upon oxidation. Also, the head group of PAs, which would be exposed on the

surface of self-assembled fibers, allows a large flexibility on the sequence. In other

words, different bioactive motifs or even sequences without bioactivity can be

incorporated in this region without hurting the self-assembling propensity of the general

design. Paramonov et al. demonstrated that hydrogen bonding close to the hydrophobic

core was necessary for the nanofiber formation and had the biggest impact on the gel

properties [162].

The mechanism of peptide amphiphile self-assembly was also studied by

molecular simulation technique with a model that incorporated molecular structure and

inter- and intra-molecular interactions [163, 164]. Different morphology could be formed

due to different molecular structures or at different phases of self-assembly. These

morphologies include single β-sheet connected laterally by hydrogen bonds, stacks of

parallel β-sheets, spherical micelles, micelles with β-sheets in the corona and long

cylindrical fibers (Figure 2-9). It was found that the self-assembly of PA molecules is

driven mostly by the interplay of hydrophobic interactions between alkyl tails and

hydrogen bonds between peptide blocks. Molecular length and thermodynamic

parameters such as temperature and concentration all play important roles in self-

Page 79: Self-Assembling Bioaffinity Hydrogels for Localized

63

assembly. However, this study did not consider electrostatic interactions between epitope

head groups or with salt ions as well as the effect of the solvent which may hydrogen

bond with the peptide units. The thorough consideration of all influencing factors

requires significant computational resources and efforts.

Figure 2-9 The simulated morphologies that could exist during PA molecule self-

assembly process. A) free molecules, b) spherical micelles, c) micelles with β-sheets on

the outside forming the corona, d) long cylindrical fibers, e) stacks of parallel β-sheets,

and f) single β-sheet. (Adapted with permission from reference [163]. Copyright (2008)

American Chemical Society)

In regard to different modes of alkylated PA self-assembly, pH is one of the major

triggers. When the pH is neutral, peptides containing acidic amino acids have a net

negative charge which keeps the PAs from self-assembly due to electrostatic repulsion.

This negative charge can be eliminated upon acidification so that hydrophobic tails can

start to aggregate after pH is reduced. In addition to pH, polyvalent metal ions such as

Ca2+, Mg2+, and Cu2+ can also induce self-assembly which could be explained by the

ionic interactions between the acidic groups in PAs and positively charged counterions

Page 80: Self-Assembling Bioaffinity Hydrogels for Localized

64

(could be the formation of ion-peptide complex). However, monovalent ions are not able

to induce the formation of self-supporting gels for most PAs. Thus, it appears that

polyvalent ions are significantly more effective initiators of self-assembly than

monovalent species. Gels produced by addition of polyvalent metal ions to PA solutions

are very stable under a broad pH range (from 4-11 for some PAs) and high temperature

up to 100 °C. In contrast, gels formed by pH-triggered mechanism are only stable at pH

below 3.5. The high stability of cation-mediated PA gels is postulated to be determined

by the diversity of intermolecular interactions including hydrophobic, ionic and hydrogen

bonds [165]. In addition, concentration is important simply because sufficient fibers are

needed to form the three-dimensional networks which will further allow the gelation to

occur.

Characteristics and applications

This type of alkylated PA molecules consists of a fatty acid chain containing 10-

22 carbon atoms in length connected via an amide bond to a peptide block which includes

6-12 amino acid residues for different functions. They can self-assemble into self-

supporting gels at concentration as low as 0.25% by weight [165]. The gels are

composed of a network of cylindrical nanofibers, and these fibers formed by alkylated

PAs all possess a diameter between 5-10 nm and a length on the order of microns. As we

already discussed, the hydrophobic alkyl tails are located at the core of the fiber while the

hydrophilic peptide segments comprise the outer surface. Although all the PAs form

fibers with comparable diameter, the length and stiffness of the fibers vary significantly

depending on the sequence of these PAs [166]. It was demonstrated by FTIR studies that

the hydrogen bonding of β-sheets tends to be aligned along the long axis of fibers but

Page 81: Self-Assembling Bioaffinity Hydrogels for Localized

65

with some amount of twisting or disruptions [167] (see Figure 2-10). Since the alkyl tails

are attached to the same terminus of the peptide segment for a certain kind of PA

molecule, every peptide would have the same C-N direction in the formed nanofibers.

Therefore, it is very likely that the fibers contain only parallel β-sheets.

Figure 2-10 Schematic representation of a single β-sheet in (a) linear and (b) twisted

geometry in a peptide amphiphile nanofiber; (c) schematic of two β-sheets in the fiber

with arrows indicating the hydrogen bonds. (adapted with permission from reference

[167]. Copyright (2010) American Chemical Society)

Aside from the cylindrical shaped fibers, giant flat nanobelts could also be formed

by certain alkylated peptide amphiphiles which contain an alternating sequence with

hydrophobic and hydrophilic side chains in its peptide block [158]. The nanobelts

possess fairly mono-disperse widths on the order of 150nm and lengths of up to 0.1mm,

eliminating all curvature from the nanostructure. With variations in monomer

concentration, these PAs can generate morphology with twisted ribbons. After

appropriate functionalization, these nanostructures could offer a novel architecture to

display epitopes to cells for therapeutic applications.

Page 82: Self-Assembling Bioaffinity Hydrogels for Localized

66

It was also demonstrated that oppositely charged PA molecules with different

bioactive sequences can co-assemble into nanofibers with the simultaneous display of all

epitopes on the surface [160]. The electrostatic attractions between the oppositely

charged head groups can facilitate self-assembly by forming salt-bridged pairs at neutral

pH [168]. In this case, pH serves as a trigger to introduce the opposite charge state to the

PA molecules, thus promoting the self-assembly. In addition, since different PA

molecules are capable of co-assembly, optimized cell signaling can be achieved by

mixing a specific bioactive molecule with a different bioactive molecule or a non-

bioactive diluents molecule to vary the epitope density on the assembled nanostructure.

The stiffness of the gels formed by PAs is an important feature when the matrix is

used for regenerative medicine scaffold since cells respond to the mechanical properties

of their environment and the mechanical properties could influence cell behavior [169-

171]. The matrix stiffness can be affected by the amino acid sequence in the PAs. A

study was performed to investigate the number and position of valines and alanines in the

peptide sequence to the mechanical properties of the nanofiber gels that the PAs form

[167]. It turned out that increasing both the number and fraction of valine, which has the

highest propensity to from β-sheets, is especially effective at raising mechanical stiffness

of the gels. Moreover, the high gel stiffness is also facilitated by alignment of hydrogen

bonds along the fiber axis as well as the placement of β-sheet-forming residues near the

hydrophobic core of nanofibers.

The alkylated self-assembling PA molecules studied by the Stupp’s group were

initially designed for tissue engineering purposes and have been used for a host of

Page 83: Self-Assembling Bioaffinity Hydrogels for Localized

67

biological applications. For example, a phosphorylated residue was introduced into the

peptide segment to induce the mineralization of hydroxyapatite in order to recreate the

nanoscale structure of bone [159]. Also, with proper functionalization in the peptide

terminus, the self-assembled nanostructures offer architecture to present epitopes to cells

for varying therapeutic applications. Commonly incorporated epitopes include RGD

[172], IKVAV [173] and YIGSR[174], etc. Some bioactive epitopes can promote rapid

and selective differentiation of neural progenitor cells into neurons [175]. A heparin-

binding peptide amphiphile (HBPA) was designed by Ghanaati et al. with a heparin-

binding domain to specifically bind and self-assemble at the presence of heparin sulfate-

like glycosaminoglycans (HSGAG), allowing for the capture of angiogenic growth

factors such as VEGF and EGF-2 in a biomimetic fashion. The formed gels were

demonstrated to have good biocompatibility and be able to promote angiogenesis in vivo

which was reflected by extensive blood vessel formation in a rat cornea [176]. It can also

improve cell viability and function in the interior islets as an artificial ECM after

delivered to pancreatic islet at a very low concentration together with angiogenic growth

factor. This contributes to the treatment of type 1 diabetes through islet transplantation

[177]. Moreover, the heparin-presenting nanofiber networks were demonstrated to

capture and deliver paracrine factors therapeutically for cardiovascular diseases to elicit a

paracrine effect [178]. Another example includes bioactive, cylindrical nanofibers

bearing a cationic α-helical cytotoxic peptide (KLAKLAK)2. This novel amphiphilic

molecule is generated by attaching the peptide to a PA molecule. These (KLAKLAK)2

nanostructures formed after self-assembly, but not the peptide alone, could be readily

internalized by breast cancer cells and induce cancer cell death by membrane disruption

Page 84: Self-Assembling Bioaffinity Hydrogels for Localized

68

[179]. The cytotoxic peptide amphiphiles bearing (KLAKLAK)2 have been applied for

cancer therapy since it can self-assemble into supramolecular membranes acting either as

reservoirs for sustained release of cytotoxicity upon enzymatic degradation or just as

membranes with surface-bound cytotoxicity [180].

As we already know, the nanostructure matrix formed in the presence of

suspended cells before self-assembly can entrap cells in the matrix. This entrapment

should not arrest cell proliferation and motility and is presumably expected to improve

cell viability and promote retention after targeted application if we use PA-based

nanostructure scaffold for cell delivery. This has been demonstrated with bone-marrow

mononuclear cells (BMNCs) for ischemic tissue therapies [181]. In addition, the

entrapped cells were proven to be able to internalize the nanofibers and probably utilize

PA molecules for metabolic pathways which solves the toxicity problem and at the same

time provide nutrients for the cells [165].

Since the cylindrical nanofibers formed by these PAs containing hydrophobic

cores, hydrophobic small molecules or drug moieties have the potential to be

encapsulated within the core for desirable delivery. The encapsulation has been tested in

multiple substances such as pyrene [182], prodan [183] etc.

In addition to therapeutic applications, magnetic resonance imaging (MRI)

contrast agents could also be designed into some PA molecules to combine high-

resolution three-dimensional MR fate mapping of the scaffolds with targeting of specific

cellular receptors [184, 185].

Page 85: Self-Assembling Bioaffinity Hydrogels for Localized

69

Common features of these self-assembling peptide amphiphiles

Although the self-assembling mechanisms and peptide orientations in the fibers

are slightly different for each type of peptide amphiphiles discussed in this chapter, the

intrinsic forces contributing to the self-assembly are the same in nature. They include

hydrogen bonding, hydrophobic forces, electrostatic interactions, repulsive steric forces,

nonspecific van der Waals interactions, π-π stacking and chiral dipole-dipole interactions

[157]. All systems involve a combination of these forces which could counterbalance the

entropic cost caused by peptide aggregation.

These self-assembling peptide amphiphiles are all designed to go through sol-gel

transition after exposure to external environmental changes and they share a lot of

common features in regard to the physical structure of resultant material after self-

assembly. The interwoven fibrillar networks formed by all of these peptide amphiphiles

possess diameter and length on the same order of magnitude although their specific

values may vary due to their different lengths, compositions and external cues

(comparison see Table 2-3). These peptide-based networks can serve as novel materials

for tissue engineering simply because the hierarchical supramolecular structures and

mechanical strength those peptides generate resemble biological ECM and they can

interact with cells at the molecular level [165]. Their applications as drug delivery

scaffolds can be attributed to the restriction effect of gel matrix to the drug molecules

incorporated. The release of the drug payload is dependent on the mesh size of the

matrix as well as the electrostatic interaction between the drug molecule and peptide

scaffold. In addition, these peptide-based structures are biologically benign toward

Page 86: Self-Assembling Bioaffinity Hydrogels for Localized

70

various cell lines and will not incite immune responses in macrophages since they could

be eventually degraded to naturally occurring amino acid residues.

Table 2-3 Comparison for properties of representative peptide amphiphiles from each

category

Peptide

category

Representative

sequence

Concentration

threshold

Hydrogel

geometry

Reassembly

property

Mechanical

rigidity

Ionic-

complementary

self-assembling

peptide

EAK16-II CAC: 0.1

mg/ml (6*10-5

M)

[101]

Fiber

diameter:

10-80 nm;

pore size:

20-500 nm

[24, 84, 101]

N/A G’< 20 Pa

even at 10

mg/ml

[133]

RAD16-I

(PuraMatrixTM)

N/A Fibers

width: 10-20

nm;

length:100-

1000 nm

[115];

pore size: 5-

200 nm

[130, 131]

Dynamic

reassembly

completes

within time

scale for

hours [120,

121]

G’= 36.81

Pa [130] /

1541.7 Pa

[132] / 4.7

kPa [120] for

10 mg/ml

RAD16-I

(not

consistent)

β-hairpins (VK)4VDPPT(VK)4

(MAX 1)

N/A Fiber width:

3 nm

Fiber height:

2 nm [145]

Fiber length:

10-200 nm

[148]

Pore size

18-50

nm[155]

Gel

recovers

immediately

after

cessation of

shear [151]

mechanically

rigid gel

G’>2500Pa

(2 hours

after

initiating

self-

assembly for

2% MAX1)

[147]

Alkylated

peptide

amphiphiles

C16H31O-NH-

CCCCGGGS(P)RGD

(Generally: 10-22

carbon chain

attached to 6-12

amino acids)

Gelation

threshold:

0.25% (wt%)

[157]

Fiber

diameter:

7.6nm

(5-10nm in

general)

Fiber length:

several

micrometers

[157, 167]

N/A stiffness

varies

considerably

depending

on sequence

(G’: from

<100 Pa to >

10 kPa)

[165, 167]

CAC: critical aggregation concentration; N/A: not available

Page 87: Self-Assembling Bioaffinity Hydrogels for Localized

71

Moreover, most of the peptides can be easily chemically synthesized through

solid-state phase peptide synthesis technology and this technology can produce peptide

amphiphiles in relatively large quantities within short time. One exception is alkylated

PAs which need an additional manual coupling reaction to attach the peptide portion

generated from solid phase synthesis to long-chain aliphatic acid via the formation of an

amide bond. Alternatively, with the help of genetic engineering, epitopes with different

functionalities can be incorporated into the peptide sequence to introduce adhesion,

differentiation or other specific biological effects [186]. Supramolecular self-assembling

biomaterials could thus be easily controlled in terms of architecture, shape and

dimensions as well as density of bioactive signals.

Owing to the easy manipulation of both peptide composition as well as external

conditions, the properties (including gelation kinetics, gel stiffness, network mesh size) of

the self-assembled structure (usually cross-linked fibrils) can be readily controlled and

modified by the adjustment of amino acid sequence, peptide concentration or external

environmental conditions such as solution ionic strength, pH or temperature etc. in order

to achieve desired cell encapsulation or controlled release of therapeutics [152].

Altogether, common features shared by these peptide amphiphiles such as low

toxicity and high biocompatibility, easy production and manipulation, and wide

biomedical applications have made the study of these self-assembling peptides the focus

interest of multiple research groups. So far, great strides have been made in the study of

multifaceted aspects of these peptides, from underlying self-assembly mechanism to

macrostructure characteristics favoring their practical use in diverse fields. Continuing

Page 88: Self-Assembling Bioaffinity Hydrogels for Localized

72

effort is still needed to design novel materials and explore their applications in different

aspects.

Conclusions and future prospects

Extensive research has been done on the self-assembling peptides in the last three

decades due to their favorable features and extensive potential applications in biomedical

engineering and nanotechnology. These peptide-based molecules are present in different

forms as single molecules as discussed above, however, they all tend to form similar

fibrillar-entangled macrostructures after self-assembly. This self-assembling process

usually needs to be initiated by exterior environmental stimuli and the final resulting

structure has a semi-solid gel-like behavior with a certain mechanical strength. Those

unique characteristics of the materials make them ideal material candidates as tissue

engineering scaffolds or drug delivery vehicles. Despite high levels of similarity in

macrostructures, the differences in self-assembling mechanisms and material

microstructures still exist. Therefore, it is challenging but beneficial to distinguish the

differences among them and find the optimal application for each category.

Comparison and property predictions

Although considerable effort has been invested in the study for each type of self-

assembling peptide amphiphile as well as their corresponding biomedical applications,

little comparison has been made to find their relative suitability at a given condition. For

example, if three-dimensional culture for a particular cell line is desired, little information

is available for us to target the best fit material for this situation. Therefore, a systematic

prediction system for the performance of each peptide amphiphile under certain condition

Page 89: Self-Assembling Bioaffinity Hydrogels for Localized

73

is expected to be created to take the most advantage of the existing resources. In order to

accomplish this, a structure-property correlation for these self-assembling peptides needs

to be established. With the knowledge of this relationship, more rational molecular

design for these peptides is made possible by simply modifying their

sequences/compositions or stimulus conditions. Hence, not only comparison within each

type of peptide is valuable, inter-category comparison would further facilitate the

selection of the best possible material for each specific situation. A more thorough

understanding of relationship between peptide molecular structure and network

biomedical properties can ultimately lead to a better bio-molecular network behavior in

general. This could be one of the most compelling directions for future research.

Standardization of analytical techniques

To get a better comprehension of the self-assembling systems, the analytical

techniques employed to determine their structures and speculate their self-assembling

mechanisms are of critical importance. So far, CD and FTIR are the most frequently

used methods to determine the secondary structures of these peptides in self-assembling

process while AFM and SEM are the most popular microscopic methods to observe the

dimensions and morphologies for the self-assembled structure. However, most

researchers use only two or three techniques for each peptide studied. This may lead us

to less accurate conclusion since each technique has its own limitations. Another

drawback is that the results obtained by one group lack consistency to the results

observed from other groups who use different techniques [187]. For example, the fiber

diameter measured by one group was 10-20nm [23] but 40-80nm [117] by another group

for the same peptide. Therefore, the speculation for its organization in the self-assembled

Page 90: Self-Assembling Bioaffinity Hydrogels for Localized

74

structure could be significantly different. That explains why controversy still exists in

terms of peptide molecules orientation in the self-assembled fiber structures for ion-

complementary self-assembly peptides. Some proposed that these oliogopeptides are

either staggered or organized in registry to form intermolecular interactions (i.e.

hydrophobic and ionic interaction between the polar and nonpolar side chains) and fiber

is formed by the extension of these interactions where the direction of these

intermolecular interactions is parallel to the fiber axis [23]. Others speculated that fibers

are developed in length through the extension of intermolecular backbone hydrogen

bonding where hydrophobic interactions only help to form the β-sheet bi-layer which

determines the height of the filament [112]. Since the fiber dimensions are usually used

to speculate molecular organization in the structure and the self-assembling mechanisms

are derived accordingly, different values obtained from different groups may lead to

contradictory conclusions with regard to self-assembling mechanisms. Therefore,

multiple other technologies with higher resolutions such as X-ray crystallography, solid

state nucleic magnetic resonance (ssNMR) and even computer simulation could be

employed together with CD, FTIR, and microscopy to generate a more comprehensive

picture. More detailed information on the molecular level of these peptides can be

provided by X-ray fiber diffraction [188] and ssNMR [189] to confirm those speculations

and facilitate the development of more accurate mechanisms. For example, ssNMR

studies (together with 13C isotopic labeling) performed on RAD16-I recently [189]

confirmed that the nanofibers are composed of two stacked β-sheets stabilized by a

hydrophobic core formed by alanine side chains. More detailed studies should be

conducted in future work to derive more convincing and precise mechanisms/theories for

Page 91: Self-Assembling Bioaffinity Hydrogels for Localized

75

peptide self-assemblies. This will promote the fundamental understanding of self-

assembling principles and practically avoid unnecessary contradictory conclusions drawn

by different groups, thereby facilitating rational design of peptide-based self-assembling

materials for specific applications.

Page 92: Self-Assembling Bioaffinity Hydrogels for Localized

76

Chapter 3

Genetic Engineering of a Fluorogenic Module in Bioaffinity

Hydrogels for Antibody Delivery

A biomaterial design in which a fluorogen-activating protein (FAP) is used to

mediate loading and release of antibodies is described in this chapter. The system

consists of the FAP dL5, a dimerized single-chain variable fragment (scFv), appended

with three repeats of the amphiphilic sequence AEAEAKAK via a glycine-serine peptide

linker. This fusion generates a polypeptide called dL5_EAK. Intermixing dL5_EAK

with the -fibrillizing peptide AEAEAKAKAEAEAKAK (commonly known as EAK16-

II) results in a functionalized composite referred herein as dL5 hydrogel. The bioaffinity

hydrogel illuminates fluorescence upon addition of malachite green (MG), the

fluorogenic partner of dL5. MG was conjugated with protein A/G (pAGMG) as an adaptor

that can capture IgG to the dL5. Using confocal microscopy and dose-response titration,

we demonstrate that loading of IgG into the hydrogel was mediated by binding of pAGMG

to dL5. In in vitro experiments, it was observed that only 26% of the IgG were released

from hydrogel embedded with pAGMG after five days, compared to close to 90% in

hydrogel without pAGMG. Close to 60 % of the antibody recovered over time remained

bioactive. Denaturation may be further reduced by modulating the ratio of IgG to the

multivalent pAGMG, identified in the computational simulations as a principal parameter.

In mouse footpads, dL5 hydrogel delivered IgG were retained locally, with distribution

through the lymphatics impeded. The ability to monitor bioaffinity sites in vitro and in

vivo, orthogonal to IgG release, will aid the development and optimization of local

antibody delivery systems.

Page 93: Self-Assembling Bioaffinity Hydrogels for Localized

77

Background and Rationale

Drug discovery efforts are increasingly turning toward immunomodulatory

monoclonal antibodies aimed for chronic ailments [190]. While most antibodies are

formulated for systemic infusion, opportunities exist in developing delivery systems to

concentrate drugs in affected tissues and to target regional lymph nodes in order to affect

local immune milieu [191]. Outcomes of inflammation are governed in part by the

phenotypes of granulocytes, myeloid cells, and lymphocytes trafficking between injured

tissues and draining lymph nodes. As such, there are opportunities for developing novel

systems for local delivery of biologics. In situ-forming hydrogels can create depots of

antibody drugs in subcutaneous space. Local administration of immune modulators can

enhance efficacy and reduce adverse events. In situ-forming hydrogels or coacervates

formed by cross-linked polymer chains that remain physically intact with high water

intake are particularly attractive because of their ability to adapt to tissue space [192].

Drug loading can be accomplished by introducing the biologic during polymer cross-

linking. The pore sizes (5-100 nm) and the resulting high permeability of such depots

[193], however, preclude the ability to concentrate antibodies at injection sites and

accumulate in draining lymph nodes. Rapid diffusion of drugs out of the network (“burst

release”) can be limited by increasing cross-linking density [194] or through bioaffinity

retardation [17]. A less explored challenge is that the location of the depot cannot be

ascertained in vivo, making it difficult to determine the rate-limiting mechanism of drug

localization.

The EAK16 series of low molecular weight, -sheet fibrillizing peptides (FP)

can be formulated into hydrogels for delivering small molecule drugs, proteins, and cells

Page 94: Self-Assembling Bioaffinity Hydrogels for Localized

78

[195-197]. The FP EAK16-II, comprised of AEAEAKAKAEAEAKAK (with A as

alanine, E as glutamic acid, and K as lysine), self-assembles into cross-linking fibrils in

high ionic strength aqueous environment [195]. When administered through

conventional needled syringes, EAK16-II forms stable networks in vivo, adapting and

adhering to the local tissue contours [198]. Due to the high water content (~99.5% w/v)

[195], IgG can be loaded into the gel matrices with high efficiency. The amphiphilic

EAK16-II, carrying no net charge at pH 7.4, presumably interacts with IgG via weak

non-specific interactions, resulting in a large fraction of the protein molecules remained

unbound [199].

We have used a coassembly strategy to engineer biochemical functionalities into

EAK16-II fibrillar matrices [198, 200-203]. Using the analogue EAKIIH6, His-tags are

incorporated into EAK16-II fibrils, driven by the FP domain for -sheets. The two

peptides co-assemble efficiently, with greater than 90% of EAKIIH6 incorporated into

the composite [198] generated from excess EAK16-II. By embedding recombinant

protein A/G (pA/G) via anti-His-tag antibodies, IgG can be loaded and release over time

[201-203]. Because Protein A and Protein G bind to IgG with relative high affinities

(with apparent binding constants reported at 2.9 x 107 and 1.13 x 108 M-1 [204],

respectively), pA/G provides a stable and extended affinity mechanism [205]. In vivo,

diffusion of IgG from the material is driven by concentration gradient between the depot

and the surrounding tissues, including the lymphatics. The system renders retention of

IgG molecules in subcutaneous injection sites for up to ten days [201]. The EAK16-

II/EAKIIH6 system, when adapted with pAG and anti-His-tag antibodies, was designed

as antibody-functionalized scaffolds for cytokines and cells [201, 202]. The ability to

Page 95: Self-Assembling Bioaffinity Hydrogels for Localized

79

monitor binding sites in vivo would enhance correlation with in vitro or in vivo

optimization.

In this chapter, we describe an in situ-forming hydrogel platform on which the

binding sites for IgG, pA/G, can be monitored using an activation-dependent fluorescent

module. The hydrogel are generated by intermixing EAK16-II with its analogue

dL5_EAK, a polypeptide consisting of a FAP with two light chains, called dL5, fused to

three repeats of AEAEAKAK via a glycine-serine spacer, (GGGGS)3 [200] (Figure 3-1a).

In high ionic strength environment, EAK16-II and dL5_EAK undergo solution-gel phase

transition and coassembly, apparently simultaneously. The resultant materials are

illuminated in the presence of malachite green (MG), the fluorogen that can be activated

by dL5 [206]. Relative to free MG, the fluorescence of dL5-bound MG is enhanced by

up to 20,000 folds, due to constraining of the rotatable bonds in MG by two L5 proteins.

Crystallographic studies have shown that MG mediates dimerization of two L5 proteins

to form a fluorescent ternary complex [207]. We hypothesized that the system of MG-

conjugated pAG (pAGMG), dL5_EAK and EAK16-II, when deployed together (called

dL5 hydrogel, see Figure 3-1b), serves as a self-illuminating, injectable formulation for

local deposition of IgG. The tight binding of dL5 to MG ensures a stable intermediate

[207]. The exceptionally low fluorescence background renders the fluorogenic module

an efficient reporter of binding sites, independent of model IgG drugs [206]. We

characterized and modeled the release kinetics in vitro, and found that the dL5 hydrogel

retained IgG at the injection site and draining lymph node in vivo.

Page 96: Self-Assembling Bioaffinity Hydrogels for Localized

80

Figure 3-1 Diagrams showing (a): the construct of dL5_EAK and (b) pAGMG embedded

dL5 gel displaying IgG molecules.

Materials and Methods

EAK16-II and expression of dL5_EAK

EAK16-II was custom synthesized by American Peptide Company (Sunnyvale,

CA) at greater than 95% purity with the N-terminus acetylated and C-terminus amidated.

The sequence was confirmed by mass spectrometry. The peptide, received lyophilized,

was reconstituted in sterile deionized water (18.2 M at 25C) as needed and stored at -

80°C until use. Construction of plasmids encoding dL5_EAK was described previously

[200]. Specifically, the plasmid pET21 GST HRV dL5 was constructed by PCR

amplification of L5 (E52D L91S, referred to as L5 in this work) monomer and ligation

into the bacterial expression vector pET21 GST HRV. An annealed DNA insert

encoding the amino acid repeat (G4S)4 was ligated downstream from the L5 subunit,

followed by the ligation of PCR amplified DNA encoding an additional L5 subunit. The

Page 97: Self-Assembling Bioaffinity Hydrogels for Localized

81

plasmid pET21 GST HRV dL5_EAK was constructed by ligation of double stranded

DNA encoding 3 additional repeats of (G4S) followed by 3 repeats of the amino acids

AEAEAKAK downstream of the dL5 sequence of pET21 GST HRV dL5. Plasmid

sequences were verified by DNA sequencing. Expression and purification of dL5_EAK

was carried out by transforming pET21 GST-HRV-dL5_EAK into calcium competent

Rosetta-Gami 2(DE3) E. coli. (Novagen, Madison, WI). The transformed bacteria were

grown in 500 ml volumes and induced using 500 µM IPTG and 0.4% glucose overnight

at 20ºC. Proteins in cell lysates were captured using Ni-NTA agarose columns (Qiagen,

Valencia, CA) and eluted by HRV3C protease. Fractions containing the protein were

combined and dialyzed against PBS, with the resultant concentrate quantified using UV-

Vis ( = 280 nm). Details of plasmid construction, protein expression and purification,

can be found in Saunders et al. [200].

Conjugation of protein A/G

Protein A/G (pA/G) and malachite green isothiocyanate (MG-ITC) were

purchased from Life Technologies (Carlsbad, CA). Conjugation of pA/G to MG-ITC

was carried out in alkaline aqueous buffers (pH = 8.5 to 9.5), with the resultant conjugate

separated from unreacted MG-ITC using PD-10 columns (Sephadex G-25 M, GE

Healthcare, Aurora, OH). The degree of labeling of the eluent was calculated based on

the molar ratio of dye concentration to protein concentration. Each mole of pA/G was

conjugated with 0.67 mole of MG, as determined by bicinchoninic acid assay for the

former and UV-Vis absorbance at 633 nm of the latter.

Page 98: Self-Assembling Bioaffinity Hydrogels for Localized

82

Gel electrophoresis

Different amounts of EAK16-II (10 L, 20 L, 40 L, 80 L or 160 L, at 5

mg/ml) were added to 8.3 L of dL5_EAK (72 M) to prepare mixtures containing

EAK16_II to dL5_EAK at 50:1, 100:1, 200:1, 400:1, and 800:1 molar ratios. Each

sample was suspended in a total volume of 500 l, followed by incubation at 37°C

overnight. The samples were centrifuged at 16,000 RFC for 6 minutes with the

supernatants analyzed using denaturing Novex NuPAGE 4-12% Bis-Tris gels. To each

sample, the supernatant (5 µL) was added with 5 µL of NuPAGE SDS buffer and 10 µL

of PBS, with the resulting solution heated at 70 °C for 10 minutes before loading. All

samples were run at 200 V for 35 min in NuPAGE MES SDS running buffer. Proteins

were visualized using the SilverQuestTM staining kit for which a detection limit of 0.3 ng

has been reported. All electrophoretic reagents were purchased from Life Technologies

(Carlsbad, CA). Images were captured using the Kodak 440 Image Station (Kodak,

Rochester, NY), and band intensities were quantified using the NIH Image J software.

Percent dL5_EAK incorporated into the insoluble fraction was quantified by subtracting

the intensity of the protein band in samples containing EAK16-II from the intensity of the

samples without EAK16-II.

Viscosity measurement

Stock solutions of EAK16-II and dL5_EAK were prepared in distilled water at the

level of 10 mg/mL and 12.2 M respectively. Samples of EAK16-II with or without

dL5_EAK were prepared in PBS (pH = 7.4) and drew into positive displacement pipettes

(RheoSense Inc., San Ramon, CA). EAK16-II was prepared in different concentrations

Page 99: Self-Assembling Bioaffinity Hydrogels for Localized

83

(0.01–2 mg/ml) in two solvents (PBS and distilled water). The EAK16-II with dL5_EAK

in PBS was prepared at the ratio 400:1 and the final concentrations were 1.5 mg/ml and

2.3 µM, respectively. Measurements were performed using the VISC unit (RheoSense

Inc.) equipped with a temperature control module, which was set at 25C. Samples were

subjected to different shear rates by changing the flow rate through the capillaries (depth

= 50 m) in the sensor cartridge (HA01-01). The instrument is capable of measuring

viscosity between 0.2 to 100 mPa·s from shear rates 6.5 to 5,850 s-1 at within 2%

accuracy according to the manufacturer.

Titration of pAGMG to dL5_EAK gel

Hydrogels composed of EAK16-II and dL5_EAK (400:1) were established in

Corning FluoroBlok 96-well plate (Thermo Scientific, Inc.). Different amounts of

pAGMG were added to the hydrogel to generate samples containing dL5_EAK to pAGMG

molar ratios of 1:8, 1:4, 1:2, 1:1, 2:1, 4:1, 8:1, or 16:1. Sufficient PBS was added to each

well to attain a final total volume of 200 l. The solvated hydrogels were incubated

overnight at 37 °C before measuring MG/dL5 fluorescence using a Tecan Infinite M1000

microplate reader, with excitation set at 635 nm and emission at 670 nm. EAK16-II

with pAGMG, or EAK with a PEGylated form of MG [206] were used as the controls

(n=3).

Laser-scanning confocal microscopy

Samples were prepared for confocal imaging as described below. To 20 L of 5

mg/ml EAK16-II, 40 L of 10 M dL5_EAK were added and mixed to produce a gel

Page 100: Self-Assembling Bioaffinity Hydrogels for Localized

84

with a molar ratio of 150:1. To this mixture, 12.2 µL of 11 M pAGMG, 4 µL of 0.05

mg/mL FITC-labeled rat anti-mouse CD23 IgG antibody were added in a confocal dish

(MatTek Corp., Ashland, MA). The sample was brought up to 100 µL with PBS and

incubated overnight. The hydrogel was collected by centrifugation and washed with 200

µL of fresh PBS five times before transferring to a glass bottom microwell dish. Imaging

was performed using a Zeiss Axioplan 2 imaging microscope equipped with a mercury

lamp. Fluorescein fluorescence was collected a 470/530 nm filter, and Cy5 fluorescence

was obtained using a 710/75 nm filter (Chroma, Lititz, PA). Control samples without

dL5_EAK, without pAGMG, or containing only EAK16-II and fluorescein-labeled IgG,

were also imaged.

In vitro IgG release

The kinetics of release of a fluorescein-labeled antibody (Pierce goat anti-human

IgG) from pAGMG-embedded dL5 hydrogel were determined by monitoring fluorescence

over time. The IgG-loaded materials were established in MicroSert (inner diameter =

4.15 mm, maximum volume = 500 µL) by instilling a mixture of EAK16-II (60 nmol),

dL5_EAK (150 pmol), pAGMG (150 pmol), and IgGFITC (75 pmol). The 50 l mixture

was incubated overnight at 37°C before adding 400 l of PBS (0.1% BSA, 0.025%

NaN3) as the release medium. The insert was placed in its receptacle vial and capped

before incubation at 37°C at the start of the release. At various time points, 200 l of the

release medium was collected for measurement (excitation = 485 nm; emission = 535

nm) using the Tecan Infinite M1000. Aliquots of fresh release medium were added to

restore the system to its initial volume.

Page 101: Self-Assembling Bioaffinity Hydrogels for Localized

85

Modeling Methodology

To delineate the parameters driving IgG release, a mathematical model is

described and implemented here. The data for this comparison is taken directly from the

in vitro case described in Figure 3-10, in which diffusion of the antibody IgG from a

tailored hydrogel with and without pAGMG binding sites is compared. Six components in

the system with four possible interactions were considered:

k1

pAGMG + dL5 ⇌ pAGMG·dL5

k2

k3

IgG + pAGMG·dL5 ⇌ IgG·pAGMG·dL5 (1)

k4

k5

IgG + pAGMG ⇌ IgG·pAGMG

k6

k7

IgG. pAGMG + dL5 ⇌ IgG·pAGMG·dL5

k8

To understand the affinity interactions on the overall release dynamics, we

considered the systems of equations previously described for other affinity systems

designed for sustained release of proteins and drugs [17, 32, 42, 208]. These models

contain parameters analogous to those describe the current system in which two sets of

reversible interactions operate: binding of pAGMG to dL5_EAK, and binding of IgG to

pAGMG. Free pAGMG would bind IgG, generating IgG-pAGMG complexes that bind to

dL5 associated with EAK16-II fibrils. All dL5_EAK are assumed remained in the main

EAK16-II fibrils throughout. The relative fractions of these components are governed by

Page 102: Self-Assembling Bioaffinity Hydrogels for Localized

86

the rate constants associated with identifiable processes: with k1 > k2, and k7 > k8.

Because pAGMG was found to be relatively immobile compared to the IgG (Figure 3-8b),

the concentration of pAGMG in the sample was held constant. This allows for the isolated

study of IgG diffusion for describing the behavior of the system, focusing on the third

reaction in Equation set (1).

The diffusion of unbound IgG from the hydrogel is first modeled through Fick’s

law in a heterogeneous medium as shown in Equation (2), where IgG is the molar

concentration of unbound IgG, and D is the variable diffusion coefficient. Diffusion is

assumed to be one-dimensional, and the hydrogel and aqueous regions are assumed to be

divided neatly into two cylindrical volumes as depicted in Figure 3-10a.

2

2diff

IgG IgGD

t x

(2)

The diffusion equation is solved using central-order finite-difference methods.

The hydrogel and release medium are discretized with respect to x with even grid spacing

(100 points evenly distributed across the length of the vial). The change in IgG

concentration at each node is then estimated and integrated through finite difference

approximations (Equation (3)) [209]. Zero-flux boundary conditions are enforced at

either end of the vial. Since the diffusion coefficient is assumed to vary with location, the

averaged value between the neighboring nodes is used.

Page 103: Self-Assembling Bioaffinity Hydrogels for Localized

87

21 2 1 21 1

2 2

21

2

i ii i i i

t t

i i

D IgG IgG D IgG IgGIgG

x x

IgGIgG IgG t

x

(3)

The binding-release behavior of IgG from the pAGMG sites is assumed to follow

Hill’s law [210], commonly used for ligand binding with multiple sites. This equation

was selected due to uncertainty of the exact nature of the binding sites, by combining

them into a single association constant with a variable describing their interactions. Each

pAGMG possesses six binding sites for IgG (4 on protein A, 2 on protein G), which may

exhibit cooperative or competitive effects between the sites as binding progresses. Here

n is Hill’s coefficient, where n > 1 corresponds to positive cooperativity and n < 1

corresponds to negative cooperativity. Y is the ratio of bound sites to total sites, and K is

the association constant.

1

n

n

K IgGY

K IgG

(4)

As the unbound IgG diffuses away from the hydrogel and into the surrounding

medium, the equilibrium value for the bound pAGMG complexed with IgG shifts as well,

releasing a portion of the IgG to replace some of the diffused components. Therefore, the

total change in IgG concentration may be modeled as a combination of diffusion and

release. The diffusion component is captured through Equations (2, 3), and the release

component may be calculated by taking the derivative of Equation (4) with respect to the

concentration of unbound IgG and multiplying by the total change in concentration [211,

212]. This produces the final system of equations summarized as Equation (5).

Page 104: Self-Assembling Bioaffinity Hydrogels for Localized

88

2

2

5

51

total diff release

diff

MG

release total

MG

difftotal

IgG IgG IgG

t t t

IgG IgG IgGDD

t x x x

pAG IgG dlIgG IgG

t dt IgG

pAG IgG dlIgG IgG

t IgGt

(5)

Zero-flux boundary conditions are enforced, and the total concentration of IgG within the

vial is integrated and tracked throughout the simulation to estimate and control the error.

The resulting model calculates the concentrations of both bound and unbound IgG with

respect to time for various conditions and association constants and cooperativities. The

apparent association constant K is estimated by assuming that the concentration of bound

and unbound IgG within the hydrogel at t = ∞ are equal, and K is recalculated for each

value of the Hill coefficient n from Equation (6).

5

5nMG

final

IgG dlK

pAG IgG dl IgG

(6)

Mice and in vivo studies

Six to eight week old female (certified-virus-free) BALB/c mice were purchased

from Charles River Laboratories (Wilmington, MA) and housed in the Duquesne

University Animal Care Facility certified by the USDA. Animals were handled in

accordance with protocols approved by Institutional Animal Care and Use Committee.

The injection mixtures contained 60 nmol (in 20 µL) of EAK16-II together with 0.15

Page 105: Self-Assembling Bioaffinity Hydrogels for Localized

89

nmol of dL5_EAK, 0.15 nmol of pAGMG, and 75 pmol of Dylight-800 4X PEG

conjugated IgG (Thermo Scientific). Mixtures, some without pAGMG, were drawn into

25 G’ needled syringes and injected subcutaneously into the footpads of mice while

anesthetized with isoflurane (induction 5%, maintenance 3%). For retention studies

(Figure 3-11), each mouse was injected with the system and control, but on different

footpads. Four replicates were performed. For distribution studies (Figure 3-12), each

mouse received either solutions containing the intact system or control, in both footpads.

Three replicates were performed for these experiments. In both studies, footpads and/or

lymphatic organs were excised 72 hours after injection and imaged using an Odyssey

imager (Li-Cor, Inc., Lincoln, Nebraska). All images were captured at low intensity (2.0)

and the same focus offset (3.2 mm), and quantified using the software Image Studio (Li-

Cor, Inc.). Table 3-1 shows the formulations used in the in vivo experiments.

Table 3-1 Formulations used in in vivo experiments

Intact systema Control: retentionb

(g)

Control:

distributionc

EAK16-II (µg) 100 100 -

dL5_EAK (µg) 4.3 4.3 -

pAGMG (µg) 7.7 - -

IgG (µg) 11.3

Fluorescein

11.3

Fluorescein

11.3

Fluorescein Total volume

(l) 49.4 49.4 49.4

a“System” formulation used in experiments shown in Figure 3-11 and Figure 3-12.

b“Control” formulation used in experiments shown in Figure 3-11.

c“Control” formulation used in experiments shown in Figure 3-12.

Page 106: Self-Assembling Bioaffinity Hydrogels for Localized

90

Results

In this chapter, we aimed to develop a bioaffinity-driven system in which the

release of IgG from a fibrillar matrix was retarded in order to concentrate therapeutic

antibodies locally at injection sites and in regional lymph nodes. By concentrating

antibodies in lymph nodes to which inflamed tissues drained toward, dosage of the

injection can be reduced. Techniques of parenteral injections targeting selected lymph

nodes are established [213]. The design hinges on engineering a fluorogenic module into

the network of FP within which binding sites for IgG can be monitored optically. Two

dye-labeled IgG molecules, IgGFITC and IgG800, were used as model payloads in in vitro

and in vivo studies, respectively. Expression and purification of dL5_EAK were

enhanced by that the polypeptide does not self-assemble. The identity of dL5_EAK was

verified using SDS-PAGE and mass spectrometry [200]. Building upon the prototype

described in Saunders et al. where the coassembly of dL5_EAK and EAK16-II were

visualized by the dye Congo red and probed with SEM (Figure 3-2), we determined here

that intermixing EAK16-II with dL5_EAK at 400:1 molar ratio resulted in 94% of the

latter being incorporated into the final composite (Figure 3-3b). Unless noted otherwise,

the experiments described herein were conducted at the same molar ratio.

Page 107: Self-Assembling Bioaffinity Hydrogels for Localized

91

Figure 3-2 Morphological features of EAK16-II and dL5_EAK/EAK16-II hydrogels as

revealed in (a) Optical images (40X; scale bar=50 µm) captured using a cooled CCD

camera immediately after solutions containing 5 µL of EAK16-II (left panel), 5 µL of a

1:100 ratio of dL5_EAK to EAK16-II (middle panel) and 5 µL of 58 µM dL5_EAK

alone (right panel) were pipetted into an equal volume of PBS with 0.01% Congo red and

spotted on glass slides. β sheet structures formed by self-assembly of EAK16-II domains

are visible by Congo red incorporation. (b) SEM images (5000X magnification) of fixed

and lyophilized membrane samples. The particles (circles and pointed by a red arrow in

the image) formed on the membrane is indicative of dL5 domain.

Page 108: Self-Assembling Bioaffinity Hydrogels for Localized

92

Figure 3-3 Co-assembly and association of dL5 gel system components. (a) Schematic

depiction of the affinity hydrogel for release of IgG. (b) Analysis of incorporation of

dL5_EAK into EAK16-II fibrils using SDS-PAGE. Increasing amounts of EAK16_II

were added to the same amount of dL5_EAK to make the ratio of the two from 50:1-

800:1. The bands represent unincorporated, soluble fraction in the mixture after

centrifugation. (c) Viscosity of EAK16-II mixed with dL5_EAK at different shear rates

(n = 3). Measurements were made at 25°C using a µVISC unit equipped with a

temperature control module.

To characterize the mixing effect of adding dL5_EAK to EAK16-II fibrillar

networks, viscosity of solutions containing both was measured using capillary rheometry

by which pressure gradient of the sample is measured across microfluidic channels [214].

Consistent with results reported by the Chen group [215], viscosity of diluted EAK16-II

Page 109: Self-Assembling Bioaffinity Hydrogels for Localized

93

peaked at approximately 25% of PBS (Figure 3-4), indicating that the method is sensitive

to the same material behavior of the FP. Mixtures containing EAK16-II and dL5_EAK

were tested in a range of shear rates at 25C in PBS. Solutions dissolved with only

dL5_EAK exhibited baseline viscosity, at near 1 mPa·s (Figure 3-3c), confirming our

previous observation that the polypeptide does not self-assemble at the concentrations

tested. At shear rate 4,000 s-1, the viscosity values of EAK16-II and EAK16-II mixed

with dL5_EAK both fell between 1 to 1.5 mPa·s, indicating that the materials lack shear

resistance. Scanning EAK16-II in PBS at lower shear rates (1,000-3,000 s-1) revealed a

sharp transition at 1,800 s-1 for EAK16-II (Figure 3-3c); the downward trend from 2.1

mPa·s to 1.5 mPa·s suggests a pseudo-elastic material. The transition was attenuated in

the presence of dL5_EAK, with viscosity measured between 1.3 to 1.5 mPa·s within the

range of shear rates tested. Intermixing the two peptides does not diminish staining by

Congo red [200]. These results show that the microscopic organization in EAK16-II

appeared sensitive to perturbation by dL5_EAK, but the macroscopic structure appeared

intact. The lack of mechanical strength of the materials is not unexpected because the

low-molecular weight and the low concentration of the FP used. The mixture can be

injected without clogging, for the shear rate applied in 25 gauged syringe needles is

estimated approximately 8000 s-1. The deformable hydrogels can adapt to tissue space

when injected subcutaneously.

Page 110: Self-Assembling Bioaffinity Hydrogels for Localized

94

Figure 3-4 Viscosity of EAK16-II at different salt concentrations. Measurements were

made using VISC with shear rate set at 1,600 s-1 at 25°C. Symbols represent

independent experiments conducted.

pAGMG conjugate

Malachite green (MG) was conjugated to pA/G by using an amine reactive form

of MG –MG isothiocyanate (ITC). Under slightly alkaline pH, the lone pair electron on

the amine group of amino acid residues of pA/G is attacking the electron deficient carbon

in the ITC group and forms the thiourea product pAGMG. SDS-PAGE verified the slight

increase in molecular weight of pA/G after the conjugation (Figure 3-5).

Page 111: Self-Assembling Bioaffinity Hydrogels for Localized

95

Figure 3-5 Electrophoretic analysis of native pAG and pAGMG using SDS-PAGE.

Localization of IgG in pAGMG-embedded dL5 hydrogel

The conjugate pAGMG was tested for binding to composites of dL5_EAK mixed

with molar excess of EAK16-II (referred to as “dL5 hydrogel” hereafter) pre-formed on

plastic surface. KD between pAGMG and dL5 gel was determined to be 90.9 nM,

calculated based on titration experiments using MG fluorescence as the readout (Figure

3-6a). Saturation was observed at molar ratio of 1:1 between pAGMG and dL5. No

fluorescence was observed in PBS containing pAGMG, or in water dissolved with

EAK16-II without dL5_EAK (Figure 3-7a), all indicating specific binding of pAGMG to

dL5.

Page 112: Self-Assembling Bioaffinity Hydrogels for Localized

96

Figure 3-6 Binding and localization of pAGMG in dL5 hydrogel. (a) Titration of pAGMG

to established hydrogel (n = 3). KD was determined using the one site-specific binding

model (y axis title: FI, fluorescent intensity; a.u., arbitrary unit). No washing steps were

needed because no background fluorescence was generated from unbound MG. (b) Top

view of IgGFITC-loaded pAGMG-dL5 gels, the intact system, showing fluorescence

associated with IgG (green), MG-dL5 (red), and the two combined. Scale bar represents

20 μm. (c) Side view of the intact systems, showing fluorescent components along the

vertical axis of the gel. Gels were deposited into glass bottom microwells immediately

prior to imaging. Controls examined included gels made with components used for

making the intact system without pAGMG or dL5_EAK (Figure 3-7b).

We next determined if pAGMG can indeed mediate IgG loading in dL5 hydrogel.

As the EAK16-II fibrillar network contains large aqueous void volume (99.5% w/v), we

did not attempt to saturate the system, but determined that at least 70 g of IgG could be

loaded per mg of the peptides. Even this sub-saturating dose may suffice for delivering

Merged

Side view

IgG-FITC MG

Top view(b)

(c)

(a)

Page 113: Self-Assembling Bioaffinity Hydrogels for Localized

97

antibodies for local effects, for example with an anti-tumor necrosis factor antibody

[216]. Using laser scanning confocal microscopy, we observed scattered non-specific

IgG adsorption (green) on the surface, but the extensive co-localization (yellow)

underneath suggests dL5-bound pAGMG captured the majority of the antibodies in the

matrix (Figure 3-6). Without pAGMG, IgGFITC in the matrix was significantly reduced, as

evidenced by the relatively low fluorescence (Figure 3-7b). These data support the

fluorescence-generating loading mechanism.

Figure 3-7 Loading and co-localization of IgGFITC in dL5 hydrogels. Images were

obtained from the same experiment described in Figure 3-6 (b and c). (a) MG

fluorescence in solutions containing different components. (b) Images obtained using

(a)

(b)

Page 114: Self-Assembling Bioaffinity Hydrogels for Localized

98

confocal microscopy of control samples, including the intact system (EAK16-II,

d_5_EAK, pAGMG, and IgGFITC) without dL5_EAK (top panel), without pAGMG (second

panel) and without IgGFITC (third panel). The last panel shows image of a sample made

with EAK16-II and IgGFITC without the other components.

In vitro IgG release from pAGMG-embedded dL5 hydrogel

Release of IgGFITC from pAGMG-embedded dL5 hydrogel was monitored in vitro

over two months. The experiment was initiated by mixing the antibody with the hydrogel

components in a small cylindrical vessel with defined dimensions. In the control

experiment in which hydrogel were made without pAGMG, almost 90% (mean value,

n=3) of the antibody loaded were released in the first five days (Figure 3-8a). In contrast,

a biphasic pattern was seen with hydrogel embedded with pAGMG; a slow initial phase,

during which only 26% were released by day 5, was followed by a second phase during

which approximately 50% were released over 2 months. At the end of the eight-week

period, 22.7% of the IgG still remained, suggesting that the release would continue. The

cumulative amount released might suffice the dosage needed for local antibody drug

delivery [216]. The IgGFITC, an anti-human IgG antibody, released from the hydrogel at

the end of the 2-month period were shown to bind its cognate antigen (human IgG) at

58% capacity: 1.7 g/ml of the 2.9 g/ml anti-human IgG released recognized plastic-

adsorbed human IgG (Figure 3-9). Dissociation of pAGMG from dL5 appeared slow, as

evidenced by that more than 60% of the initial MG fluorescence remained in the gels at

the end (Figure 3-8b). These results demonstrate that the affinity binding delayed the

release of IgG from the matrix, and that pAGMG-bound dL5 is a relatively stable

component.

Page 115: Self-Assembling Bioaffinity Hydrogels for Localized

99

Figure 3-8 Monitoring in vitro IgG release and pAGMG – dL5 interaction. (a) Cumulative

release of IgGFITC from dL5 gels with (n = 3) or without pAGMG (n = 3). Each time point

represents measurement from triplicate samples in gels setup in parallel. (b) Images and

quantification showing MG fluorescence in release vessels at the start and the amount

remained at the end of the period (n = 3 for all groups). “Start” refers to day 0, “End”

refers to day 60, “Control” was the same in (a), and “Blank medium” refers to vials

containing PBS with no other components added.

Page 116: Self-Assembling Bioaffinity Hydrogels for Localized

100

Figure 3-9 Biological activity of released anti-human IgGFITC tested by binding to

immobilized human IgG (y axis title: FI, fluorescent intensity; a.u. arbitrary unit). (a)

Standard curve established for quantifying the anti-human IgGFITC (b) Titration of the

anti-human IgGFITC to plate-bound human IgG as target antigen with concentrations

determined using the standard curve. (c) Fluorescence measurements of fresh anti-human

IgGFITC and the same antibody incubated at 37°C for 60 days to validate the stability of

FITC fluorescence as the readout for IgG release amount quantification.

Modeling and simulation of release

We sought to understand the parameters governing the kinetics of IgG release

from the hydrogel. The initial phase of the in vitro release was fit to the non-steady state

Fickian model (Figure 3-10a and Table 3-2). Apparent diffusion coefficients (Dapp) were

(a)

(b)

(c)

Page 117: Self-Assembling Bioaffinity Hydrogels for Localized

101

determined by relating Mt/M as a function of t0.5. Dapp of IgG in hydrogel without

pAGMG was an order of magnitude higher than Dapp of IgG in hydrogel embedded with

pAGMG. Release of IgG was impeded in both systems, as Dapp of IgG in aqueous solution

is estimated much higher [29]. While the analysis does not encapsulate the entire release

profile, the results show that without pAGMG, a large fraction of loaded IgG would escape

by rapid diffusion shortly after the components coalesce.

Figure 3-10 Modeling and simulation of IgG release in vitro. (a) Short time

approximation of release of IgGFITC in vitro (see Figure 3-8) using Fick’s second law.

Page 118: Self-Assembling Bioaffinity Hydrogels for Localized

102

Because the small vial diameter relative to the vial length, the release is assumed 1-D

along the vial length into the saline medium. No IgG escapes the sides and bottom of

from the gel placed in the cylindrical glass vials. The volume of the gels was assumed

constant, as FP matrices do not swell. IgG was assumed distributed in the matrices

homogenously and released in one-dimensionally, with IgG concentration at boundary

assumed zero at the start. The height of the gel is 3.7 mm, with the base inner diameter

equals to 4.15 mm (per manufacturer specification). (b) The evolution of unbound IgG

with respect to time for non-cooperative binding (n = 1.0). Initially the IgG is contained

exclusively within the fibrillar matrix. Upon exposure to the release media, the IgG

begins to diffuse into the simulated vial, accompanied by a gradual release of the bound

IgG from the gel. (c) The percentage of IgG released into the vial is plotted with respect

to time with varied binding characteristics through the Hill equation. The release profiles

show the effect of the binding characteristics, and an increase in sustained release

achievable through cooperative hexavalent binding.

Table 3-2 Apparent diffusivities (Dapp) determined using approximation of early releasea

Dapp (10-10 m2s-1) R2 IgG remained in gel (%)

Control 0.075 0.92 1.4

System 0.0086 0.82 22.7

IgG in aq. 0.40b N/A N/A

aFick’s second law was used in the approximation, with Dapp being apparent diffusion

coefficient, t as time in hours, H is one-dimensional height of matrix (3.7 mm), Mt being

the amount of IgG released at time = t, and M∞ being the theoretical maximal amount

released at time infinity. Dapp were determined by regression of the data to the question:

𝑀𝑡

𝑀∞=(

16 𝐷𝑎𝑝𝑝. 𝑡

𝜋 𝐻2 )0.5

bCalculated based on the Stokes-Einstein approximation of IgG in water at 37°C [29].

The numerical model outlined in modeling methodology section was used to

further examine the experimental results, expanding on this observed decrease in

apparent diffusion with the bioaffinity hydrogel. Figure 3-10b shows the predicted

evolution of the unbound IgG along the length of the vial, assuming 1-D diffusion.

Initially all IgG is assumed to exist only in the gel, and the concentration along the length

Page 119: Self-Assembling Bioaffinity Hydrogels for Localized

103

of the vial is approximated through a sharp sigmoidal function at the boundary of the

fibrillar matrix. The diffusion coefficient is assumed to vary in a similar fashion between

the Control value and the IgG in aqueous values in Table 3-2, and the difference between

the measured Control and system values is accounted for by the binding behavior within

the model. The concentration is calculated by dividing the initial IgG concentration by

the volume of the matrix, assuming an even distribution. Equation (4) is then used to

divide the concentration into bound and unbound concentrations. As the model IgG

diffuses out of the hydrogel into the release media, additional IgG is dissociated from

pAGMG, thereby replenishing in the matrix the IgG lost to diffusion.

The concentration of unbound IgG outside of the matrix region was integrated

with respect to time and divided by the saturation concentration. This is calculated as the

original number of moles of IgG within the vial divided by the total volume of the vial (

n

finalIgG ), and remains constant between each case. Four cases are recreated; one

without pAGMG, and three with varying levels of cooperativity between the binding sites

on pAGMG, n=0.5 (negative cooperative binding), 1.0, and 2.0 (positive cooperative

binding) as defined in Equation (4) and (6). The apparent association constant K is

estimated by assuming that the concentration of bound and unbound IgG within the gel at

t = are equivalent, and K is recalculated for each value of the Hill coefficient n from

Equation (4). The values for the association constant K are calculated as 1.08, 62.54, and

2.08 * 105 for n = 0.5, 1.0, and 2.0 respectively.

The molar concentration of IgG outside of the gel is integrated across the length

of the vial and divided by the equilibrium concentration, where the moles of IgG are

Page 120: Self-Assembling Bioaffinity Hydrogels for Localized

104

evenly distributed across the vial volume. The total fluid volume considered is 250 L to

better approximate the original fluid left in the vial between measurements. This

methodology does not take into account the experimental removal and replenishment of

the buffer solution, as this removal of IgG concentrations introduced undesirable

instability in the finite difference method. Consequently, the model under predicts the

diffusion rate for the Control case, likely due to a combination of the

removal/replenishment behavior combined with the potential convective flow introduced

as a result. However, the complete system simulations in which IgG is bound to the

fibrillar matrix provides a favorable result, indicating that the binding between IgG and

pAGMG is responsible for the controlled rate of diffusion, and suggesting that the nature

of the binding is cooperative. If the diffusion for the complete system is enhanced

experimentally by the removal/replenishment of the buffer solution, then the binding is

likely to further favor the cooperative case.

Simulated cases containing a positive cooperativity show a more sustained

release. Comparing the results to Figure 3-8a, it is demonstrated that the improved and

controlled release offered by the bioaffinity gel may be predicted through a numerical

model, and this model predicts that the IgG-pAGMG binding is cooperative. The profile

of the release is intrinsically tied to the hexavalent binding between IgG and pAGMG, and

this binding is crucial for the controlled and sustained release observed in experiments.

From the plots given, the estimated values for the pAGMG-IgG binding are an association

constant of K 2.08 * 105, and a Hill coefficient of n 2.0.

Page 121: Self-Assembling Bioaffinity Hydrogels for Localized

105

Deposition of dL5 gel formulated IgG in vivo

To test the notion that dL5 hydrogel serves to retain IgG in vivo, dL5_EAK and

EAK16-II mixed with pAGMG and IgG800 were injected subcutaneously into footpads of

mice and imaged ex vivo (Figure 3-11a). The results indicate that gelation occurred

beneath the skin (Figure 3-11b). Compared to IgG800 formulated without pAGMG, the

intact hydrogel system rendered superior retention at the injection site, as reflected by the

relatively strong fluorescence emitted in the 800 nm channel (green). The concentrated

fluorescence sustained for at least 72 h, after which the signals in both sets of mice

dissipated to background level. Movement of animals would no doubt perturb the

materials located in the subcutaneous space. To this end, we probed the localization of

the pAGMG-dL5 hydrogels in footpads of living mice. In a separate experiment, it was

detected at the injection site (after 48 h) that IgG800 (green) co-localized with pAGMG-

bound dL5 (red; 700 nm) (Figure 3-11c). No MG fluorescence was detected in the

control footpad administered with free IgG800. Note that in footpads injected with the

intact system, IgG was also detected in the draining lymph node (indicated by arrow).

No MG fluorescence was detected in the same lymph node, consistent with the in vitro

observation that the pAGMG/dL5 complex is a stable component in the matrix. These

results demonstrate the consistent coalescence of IgG800, pAGMG, dL5_EAK, and

EAK16-II in vivo.

Page 122: Self-Assembling Bioaffinity Hydrogels for Localized

106

Figure 3-11 Retention of IgG and binding sites in footpad. Retention of IgG delivered

with dL5 gel at injection site. (a) Diagram depicting the in vivo experiments. Each

mouse was injected subcutaneously with solutions containing all hydrogel components

(system) in one footpad and the same except without pAGMG (control) in the other. The

mixture was drawn into a 25’ G needled syringe as a low viscosity liquid by keeping in

low ionic strength buffer. The mice were sacrified at 72 hours after the injection. (b)

Images of four replicates showing fluorescence associated with IgG800 in footpads of each

mouse three days after injection. (c) Images taken from a separate experiment 48 hours

after injection, showing fluorescence in mouse injected with IgG800 injected in saline (left

panel) and the same antibody injected in the same mouse with all components in the

system (right panel).

(a)

(b)

(c)

control system control system

control system

mouse #1

control system

mouse #2

control system

mouse #3

control system

mouse #4

NIR800 MG

Page 123: Self-Assembling Bioaffinity Hydrogels for Localized

107

In a separate set of mice, distribution of IgG800 was traced by fluorescence in

regional and distant lymphoid organs (Figure 3-12). Mice injected with the antibody in

plain PBS (into both footpads) resulted in detectable fluorescence in sciatic, inguinal,

brachial lymph nodes, and to a lesser extent, the spleen. Distribution of IgG800 delivered

with the system (i.e. pAGMG, dL5_EAK and EAK16-II) was mostly sequestered to sciatic

lymph nodes, draining from the footpads; inguinal and brachial nodes showed lower

fluorescence, with spleen at near background level of intensity. In mice receiving the

intact system, the low fluorescence in distant lymph nodes is congruent with the strong

fluorescence remained at the injection site. These results indicate the dL5 hydrogel

impeded the distribution of IgG deposited at the injection site.

Page 124: Self-Assembling Bioaffinity Hydrogels for Localized

108

Figure 3-12 Distribution of IgG800 injected in footpads 72 h after injection. (a) Mice were

injected with either control or system mixtures in both footpads. Unlike the experiments

described in Figure 3-11, here each mouse was injected with the same formulation. (b)

The images show fluorescence associated with IgG800 in (1 and 1’) sciatic, (2 and 2’)

inguinal, (3 and 3’) brachial lymph nodes, the spleen, and footpads from mice

administered with the intact gel system mixed with IgG800, or IgG800 in PBS (control).

1

2

3 spleen

footpads

control mouse #1system mouse #1

control mouse #2system mouse #2

control mouse #3system mouse #3

1’

2’

3’

same injection (system or

control) in both footpads

Page 125: Self-Assembling Bioaffinity Hydrogels for Localized

109

Discussion

This bioaffinity hydrogel design was motivated by the need for a tractable

injectable system by which antibodies could be concentrated and monitored in vivo.

Potential applications include intratumoral antibody injection [20], and topical delivery of

anti-cytokine antibodies [217]. The inflammatory mediator TNF driving non-healing

cutaneous wound is a case in point. In addition to releasing soluble TNF, macrophages

infiltrated into the wound tissues express transmembrane TNF [218]. Together with

neutrophils and adipocytes, macrophages impose local inflammation via autocrine and

paracrine actions of both forms of TNF. To maximize effectiveness, topically delivered

anti-TNF antibody must penetrate into tissues underneath the wound bed. Macrophages

and neutrophils transverse on and below the wound surface, trafficking through the

dermis and draining lymph nodes, and cluster around adipocytes underneath. Thus,

design of topical formulations should aim at rendering sustained anti-cytokine antibodies

releasing into local-regional tissues, yet without excessive drug partitioning into the

blood circulation.

The results presented herein demonstrate that the dL5 hydrogel system can

accumulate IgG at the site of injection and in draining lymph nodes. The study is a step

towards modular engineering of local antibody delivery. The binding of pAGMG to dL5

serves two purposes, in rendering IgG binding sites, and generating fluorescence that can

be detected in subcutaneous tissues. We presented data demonstrating that the system

delayed release of IgG. In vitro, approximately 50% of IgGFITC were released in the last

55 days of the two-month period with dL5 hydrogel embedded with pAGMG, without

which near 90% of the antibody were released in the first five days (Figure 3-8). The

Page 126: Self-Assembling Bioaffinity Hydrogels for Localized

110

adaptor pAGMG mediated IgG loading (Figure 3-6) and was found to be a relatively

immobile component in the system (Figure 3-8b). In vivo, distribution of IgG was

impeded by the dL5 hydrogel, restricting a large fraction of the dose at the injection site

and in the immediate draining lymph node (Figure 3-11 and Figure 3-12). The prototype

delivered a limited dose over 24 hours, but such formulation may be suffice for local

immune modulation [216]. The computational simulations suggest cooperative binding

of IgG to pAGMG (Figure 3-10).

The hydrogel is generated through coassembly for functionalized FP [210, 219-

222]. We have shown that the EAK16-II/EAKIIH6 composite prolongs in vivo retention

of IgG in inflammatory tissues, in mammary tumors implanted in syngeneic mice [201],

and in tissues underneath skin allografts [202]. By separating affinity binding and

gelation, the former can be adjusted independently of the latter. The ability to vary the

density of dL5 through intermixing the polypeptides allows mutual optimization of bulk

physical properties and IgG loading and release kinetics. Because the pore size and

tortuosity of the hydrogel matrix can be tuned by changing EAK16-II concentration,

other agents, including small molecule drugs, or clusters of cells [223], may be

encapsulated without impeding the ability to regulate IgG loading and release. In the

presence of excess EAK16-II, dL5_EAK incorporates efficiently into the network,

without disrupting the integrity of the fibrils [200]. But excess dL5_EAK in vivo may

self-aggregate (similar to EAKH6 [198]), creating ectopic clusters of pAGMG outside the

hydrogel matrix, diverting IgG from the main depot. This can be mitigated by

formulating the molar fractions of the two polypeptides such that greater than 90% of

dL5_EAK is incorporated (Figure 3-3). Thus the ratio of the two polypeptides is an

Page 127: Self-Assembling Bioaffinity Hydrogels for Localized

111

important parameter in engineering functionalities within the EAK16-II fibrillar matrix

[21].

After incubation for two months in vitro, IgG released from dL5 gels retained

58% of antigen-binding capacity (Figure 3-9). The high water content of EAK16-II

hydrogels should preserve protein native conformations [224], but the long incubation

period in vitro (Figure 3-8) may have promoted denaturation of the IgG. Through Fc

binding to pAGMG, IgG in dL5 hydrogel are oriented [225-227] and patterned such that

random adsorption and denaturation can be reduced. This is supported by scanning

electronic microscopic images of EAK16-II/EAKIIH6, which show what appear as

individual islands of IgG/pAG complexes [201]. The templated antibodies are active

against soluble [201] and cell-bound antigens [203]. Denaturation can be minimized by

optimizing key parameters; the simulations (Figure 3-10) show cooperativity of IgG

binding in the system driven by the hexavalent pAGMG on which two or more IgG may

be captured. Altering the density of pAGMG in the hydrogel, guided by MG fluorescence,

could minimize the loss of protein activity.

Potential systemic toxicities are mitigated by the fact that the gel is largely

retained locally at the injection site. Because pAGMG remained largely with the hydrogel,

the amount released into the systemic lymphatic and blood circulation should be low.

Phagocytes will likely opsonize the fluorogen-protein conjugate. We have evidences

showing that EAK16-II appears non-inflammatory; neither EAK16-II nor pAGMG elicited

IL-6 production from freshly isolated splenocytes cultured in vitro (data not shown). As

with other synthetic biomaterials, the hydrogel is likely be degraded by the foreign body

Page 128: Self-Assembling Bioaffinity Hydrogels for Localized

112

reaction (FBR) [228]. The process is triggered by adsorption of endogenous proteins,

followed by adhesion of macrophages, which may fuse into foreign body giant cells

(FBGC). This pathway, which may take weeks to consummate, serves to sequestrate

foreign substances.

The system described herein is unique in that the bioaffinity module has a built-in

imaging feature. In dL5-pAGMG, the hydrogel can be imaged in superficial tissues.

Compared to conventional fluorescent reporters, MG-dL5 has enhanced photo-stability.

The FAP protects the fluorogen from oxidative damage as crystallographic studies show

that a single MG molecule is flanked by two L5 domains [207]. This feature was

exploited here to monitor the stability of the IgG binding sites, as pAGMG would escape

from the gel matrix to become non-fluorescent. Degradation of dL5, either through

extracellular or intracellular means, would diminish the signal, as unbound MG has low

intrinsic fluorescence. The fluorescence generated by MG and dL5 interaction could

serve as a surrogate for the extent of in vivo degradation, such as phagocytosis by

macrophages [229].

While the present study used dL5-MG as the fluorogenic module, other FAPs

have been synthesized for additional fluorogens. FAP-fluorogen pairs can be selected

based on binding affinities and fluorescent spectra suited for specific applications.

Fluorescence generated from dL5-MG can be detected optically in superficial tissues, but

other FAP systems, such as far-red fluorogenic dyes currently being explored, may be

used for internal imaging [230]. There are also promiscuous FAPs that can bind more

than one fluorogen [231], with alternate excitation and emission wavelengths and varying

Page 129: Self-Assembling Bioaffinity Hydrogels for Localized

113

affinity constants for fluorogen binding. There may be biological scenarios in which

higher or lower binding affinities are called for, such as using a low affinity FAP-

fluorogen combination as a more rapid but controlled release mechanism, through

disassociation of the pA/G-dye conjugate from the FAP itself. Additionally, other dyes

which generate higher quantum yields and signal such as cyanine-tagged MG [232] could

be used to improve signal-to-noise ratio for hydrogel with low FAP incorporation

efficiency. These other FAPs can be genetically fused with FP domains such as

AEAEAKAK as functional modules in excess EAK16-II, as demonstrated here.

Conclusions

We demonstrated herein that biochemical and imaging functionalizing of a βFP-

based injectable platform. This depot platform can be used to deliver IgG locally of a

variety of specificities. The delayed release demonstrated in vitro may be translated into

in vivo prototypes through iterative optimization, guided by in situ detection of MG-dL5

fluorescence. This design could be used to deliver antibodies for suppressing cutaneous

inflammation and targeted immune modulation in draining lymph nodes.

Page 130: Self-Assembling Bioaffinity Hydrogels for Localized

114

Chapter 4

A Facile Fc-binding Composite in Bioaffinity Hydrogel for Antibody

Delivery

The ability to congregate high concentrations of multiple antibodies in diseased

tissues for prolong durations renders a method to engage diverse molecular targets in a

sustained manner. In this chapter we sought to determine the extent to which IgG

molecules could be concentrated in subcutaneous tissues and allogeneic skin transplants

while limiting systemic exposure using a facile Fc-binding composite. This biomaterial

design centers on the novel polypeptide pG_EAK, consisting of a truncated protein G

fused with repeats of the amphiphilic sequence AEAEAKAK (“EAK”). The direct

attachment of Fc binding domain protein G to EAK reduced the number of noncovalent

interactions involved in the bioaffinity hydrogel compared to the pAGMG-dL5 gel

described in the previous chapter, rendering a design with a better translational prospect.

Intermixing pG_EAK with molar excess of EAK generated fibrillar gels with multivalent

Fc-binding sites. Capture of IgG was verified in gels prepared from admixing pG_EAK

and EAK16-II (pG gel). Injection of a dye-conjugated IgG formulated with pG gel into

skin allografts resulted in detectable fluorescence at the site in mice for at least six days,

at which time only background signal was observed in grafts injected with the same

antibody delivered in saline or non-Fc binding EAK gels. The advantage coincided with

weaker fluorescence in the liver, spleen and lymph nodes, relative to those in graft-

bearing mice injected with the antibody formulated in saline or non-Fc binding EAK gels.

In conclusion, pG gel was shown to display IgG locally, thus a potential platform for

modulating the cellular and molecular mechanisms in target tissues.

Page 131: Self-Assembling Bioaffinity Hydrogels for Localized

115

Background and rationale

Described in the preceding chapter was a system in which IgG capture was made

possible by two non-covalent interactions and one conjugation reaction. The complexity

of the design complicates the effort to optimize performance of such system. To reduce

the complexity and improve the translational prospect, a second generation bioaffinity

hydrogel was constructed by incorporating the IgG capture domain directly in the EAK

gel matrix, without the need for conjugating pA/G with MG [233]. This simplified

system consists of only one interaction between IgG and Fc-binding domain, in this case

a truncated Streptococcal protein G appended with EAK repeats, hereafter referred as

“pG_EAK”. Functionalized gels can be prepared by intermixing excess EAK with

pG_EAK, thereby encasing the latter.

Protein G (along with protein A) is widely used in lab-scale and commercial

purification of antibodies. Staphylococcal Protein A (SpA) has been developed into a

systemic therapeutic agent (PRTX-100) approved for use in humans [234]. Native

protein G variants are cell wall components isolated from Streptococcal strains (e.g.

G148) [45]. The optimal binding of protein G spans over a broad pH range, from 4 to 9

for human IgG [48]. Compared with SpA, protein G reacts with IgG from broader range

of species, and in particular binds to all four subclasses of human IgG [47, 235]. The

equilibrium binding and kinetic constants of protein G for IgG have been determined

(Table 4-1), and the binding was found to be more stable than protein A-IgG [48]. Since

the affinity interaction between protein G and IgG is not as pH and salt-dependent as it

between protein A and IgG and protein G has a wider and stronger reactivity profile

compared to protein A [236], protein G was selected to be directly attached to

Page 132: Self-Assembling Bioaffinity Hydrogels for Localized

116

amphiphilic self-assembling peptide EAK for the bioaffinity gel construction. Through

coassembly of the novel construct pG_EAK with excess parent polypeptide EAK16-II,

protein G could be directly embedded in the hydrogel for IgG loading, thereby

independently replacing dL5 bound pAGMG in the gel design described in the previous

chapter.

Table 4-1 Equilibrium constants comparison between protein G and protein A with IgG

from various species (Determined by Scatchard plots using immobilized IgG and

radioiodinated protein G or protein A. Republished with permission from reference [48],

permission conveyed through Copyright Clearance Center, Inc.)

IgG Protein G Protein A

M-1 (X 10-9)

Human 67.4 44.1

Rabbit 69.8 52.3

Goat 14.3 Not detected

Mouse 41.5 25.7

Rat 1.4 0.018

Human IgG1 2.0 1.3

Human IgG2 3.1 2.1

Human IgG3 6.1 Not detected

Human IgG4 4.7 3.4

Native protein G contains multiple domains, including three homologous regions

to the amino-terminus that have affinity for albumin [237, 238]. Toward the C-terminal

end of protein G are three Fc-binding domains (C1, C2 and C3) followed by a cell

membrane anchoring domain [47, 237]. In constructing the pG_EAK polypeptide, only

the Fc-binding track of sequence (C1-D1-C2-D2-C3) was included, with three GS linkers

Page 133: Self-Assembling Bioaffinity Hydrogels for Localized

117

connecting to three repeats of AEAEAKAK at the C-terminal (Figure 4-1 (a)). The

rationale for the order is that the Fc-binding and amphiphilic domains are physically

connected yet operate independently. Removal of the albumin binding sites reduces the

extent of cellular internalization via opsonization [239]. The pG_EAK polypeptide was

expressed in E. Coli transformed with a plasmid encoding the amino acid sequence of the

truncated pG domains, the GS linker, and the EAK tract, as well as GST and His-tag for

purification. Composites of pG_EAK and EAK were characterized in vitro and in vivo

for stability and kinetics of association with IgG molecules. As with the dL5 gel system,

the purpose is to construct an antibody-delivery system for local or topical applications in

models of inflammatory diseases. The diagram for the proposed design is shown in

Figure 4-1 (b).

Figure 4-1 Diagrams showing (a): the construct of pG_EAK and (b) coassembled pG gel

displaying IgG molecules.

Page 134: Self-Assembling Bioaffinity Hydrogels for Localized

118

Materials and Methods

Peptides and reagents

EAK16-II was custom synthesized by American Peptide Company (Sunnyvale,

CA) at greater than 95% purity with the N-terminus acetylated and C-terminus amidated.

The sequence was confirmed by mass spectrometry by the manufacturer. The peptide,

received lyophilized, was reconstituted in sterile deionized water (18.2 M at 25°C) as

needed and stored at -80°C until use.

Expression and purification of pG_EAK

Expression and purification of pG_EAK was performed after transforming

plasmid (pET21) engineered with the nucleotide sequences encoding both protein G and

EAK, together with 10X His GST HRV cleavage site into BL21(DE3) competent

Escherichia Coli (Invitrogen One Shot®). The diagram showing the complete procedure

is depicted in Figure 4-2. Full protein sequence of pG_EAK after HRV protease

cleavage is as follows:

GPSKLTYKLILNGKTLKGETTTEAVDAATAEKVFKQYANDNGVDGEWTYDDAT

KTFTVTEKPEVIDASELTPAVTTYKLVINGKTLKGETTTEAVDAATAEKVFKQYA

NDNGVDGEWTYDDATKTFTVTEKPEVIDASELTPAVTTYKLVINGKTLKGETTT

KAVDAETAEKAFKQYANDNGVDGVWTYDDATKTFTVTEGGRGGGGSGGGGS

GGGGSAEAEAKAKAEAEAKAKAEAEAKAK.

Specifically, the starter culture of transformed bacteria was grown in 5 ml of LB

media with 100 μg/ml ampicillin, then transferred to 500 mL of LB+ media (LB media

Page 135: Self-Assembling Bioaffinity Hydrogels for Localized

119

with 100 mM phosphate, 20 mM succinic acid and 0.4% gylcerol) and grown until its

optical density at 600 nm reached 0.6. The temperature was then changed to 20°C and

the culture was grown for another 1h before adding 500 μM IPTG and 0.4% glucose

(final concentration) to start induction for expression of pG_EAK. After overnight

culture at 20°C, cells were pelleted at 4°C, 6000 rcf for 20 min, resuspended in 30 ml of

buffer (50 mM Tris HCl, 750 mM NaCl, 50 mM imidazole, 0.1% TritonX-100 and

0.025% Tween 20) and frozen at -20ºC for at least 4 hours. Thawed cell pellet was lysed

by sonicating on ice for 15 min (level 3, 10s on, 5s off). Lysate was clarified by

centrifugation at 13000 rpm, 4°C for 30 min. Lysate was then added to 0.6 ml of Ni-

NTA agarose (washed with 1X Ni-NTA bind buffer: 300 mM NaCl, 50 mM NaH2PO4,

10 mM imidazole, pH 8.0) and incubated on shaker in cold room for 2 hours. Bound

agarose was washed with 21 mL of 1X Ni-NTA wash buffer (300 mM NaCl, 50 mM

NaH2PO4, 20 mM imidazole, pH 8.0) and 14 ml of 1X HRV 3C cleavage buffer (150

mM NaCl, 50 mM Tris-HCl, pH7.5). The column was capped on bottom after which 120

µL of 2U/µL HRV 3C protease and 1.2 ml of 1X HRV 3C cleavage buffer were added to

the Ni-NTA agarose. The column was capped on top as well and incubated overnight on

shaker in cold room. Eluted proteins were collected the following day by adding 50 µL

of Ni-NTA agarose to bind potential excess HRV 3C protease, followed by incubation in

cold room for another 2 hours. Column was then uncapped and eluent was collected.

Additional 0.8 ml of 1X HRV 3C protease cleavage buffer was passed over the column

twice to collect any additional digested protein. Purified pG_EAK was quantified by 280

nm absorption measurements.

Page 136: Self-Assembling Bioaffinity Hydrogels for Localized

120

Figure 4-2 Diagram showing the procedure flow for pG_EAK plasmid construction,

transformation, expression and purification.

MALDI-TOF mass spectrometry

Purified pG_EAK obtained from above procedure was diluted 10 times with a final

concentration of 5.46 µM. Combine 40 ul of this diluted solution with 10 µL of 0.5%

TFA in MilliQ-water and apply the protein directly to target of Applied Biosystems

Voyager DE-STR MALDI-TOF mass spectrometer.

Coassembly experiments

Coassembly of pG_EAK into EAK16-II fibrils was studied using gel

electrophoretic (SDS-PAGE), Congo red binding, and viscosity measurement which were

described separately below.

Page 137: Self-Assembling Bioaffinity Hydrogels for Localized

121

Gel electrophoresis

Different amounts of EAK16-II (10 µL, 20 µL, 40 µL or 80 µL, at 5 mg/ml) were

added to 7.34 ul of pG_EAK (82.2 μM) to prepare mixtures of EAK16-II and pG_EAK at

different molar ratios (50:1, 100:1, 200:1 and 400:1) of the two components. Each

mixture was q.s to a total volume of 500 µL and incubated overnight at 37°C. The

samples were centrifuged at 16000 rfc for 6 min with 5 µL of the supernatant analyzed

using denaturing Novex SDS-PAGE gel electrophoresis (NuPAGE 4-12% Bis-Tris

Electrophoresis system). For sample preparation, 5 µL of the supernatant was mixed

with 5 µL of NuPAGE sample buffer and 10 µL of PBS. The mixtures were heated at

70°C for 10 min before loading while 5 µL of the molecule maker was mixed with 10 µL

of sample buffer and 25 µL of PBS without heating before loading. All samples were run

at 200 V for 35 min in NuPAGE MES SDS running buffer. Fast silver staining

procedure (SilverQuestTM staining kit) was used to visualize the protein bands. Gel

images was captured using Kodak 440 image station and band intensities were quantified

by Image J software. Percentage of pG_EAK incorporated into EAK coassemblies was

calculated by subtracting the intensity of the protein band in samples containing EAK16-

II from the control sample with only pG_EAK in solution.

Congo red binding

The stock solution of 0.005% Congo red was prepared by diluting 30 µL of 0.1%

of its aqueous solution using 570 µL of PBS. The absorbance spectra of Congo red were

scanned from 400 nm to 700nm with Tecan Infinite M1000 microplate reader after

adding 40 µL of the stock solution to each of the samples either containing 20 µL of

Page 138: Self-Assembling Bioaffinity Hydrogels for Localized

122

EAK alone at the concentration of 5 mg/ml, 20.8 µL of pG_EAK alone at the

concentration of 7.2 μM or the mixture of the two components. For the sample

containing only one component before adding Congo red, either water was added to

substitute EAK or PBS was added to substitute pG_EAK. The molar ratio of EAK and

pG_EAK was at 400:1. Incubation was performed for 1 hour after adding Congo red and

before the absorbance scan. The absorption peak of each sample was registered for

comparison as an indication for fibril formation confirmation.

Viscosity measurement

Stock solutions of EAK and pG_EAK were prepared at 10 mg/ml in water and

9.26 μM in PBS respectively. Samples of EAK with or without pG_EAK were prepared

in PBS and drew into positive displacement pipettes (RheoSense Inc., San Ramon, CA).

The ratio of EAK to pG_EAK was prepared at 400:1 and the final concentrations were

1.5 mg/ml and 2.3 μM, respectively. Measurements were conducted using the μVISC

unit (RheoSense Inc) equipped with a temperature control module, which was set at

25°C. Samples were subjected to different shear rates (1200-4000 s-1) by changing the

flow rate through the capillaries (depth = 50 μm) in the sensor cartridge (HA01-01).

Controls including pG_EAK alone and distilled water alone were measured at the same

conditions.

Page 139: Self-Assembling Bioaffinity Hydrogels for Localized

123

IgG binding experiments

Surface plasmon resonance

Biacore T200 was used to conduct experiments with a CM5 chip at 25°C. pG

(25.3 kDa, 0.92 mg/mL stock concentration) were used as ligand to immobilize onto the

CM5 sensor surface. Ab-1 (mouse monoclonal IgG, 150 kDa, 3.3 µM stock

concentration), Ab-2 (goat monoclonal IgG, 150 kDa, 6.7 µM stock concentration), and

Ab-3 (rabbit IgG monoclonal, 150 kDa, 3.3 µM stock concentration) were used as

analytes to flow over the ligand immobilized surface. Flow Cell (FC) 1 was used as

reference for FC2, and FC3 for FC4. pH scouting of ligand was done and pH 4.0 was

selected for pG as the immobilization pH value. pG was diluted (1:500 dilution, 1.84

µg/mL diluted concentration) in 10 mM sodium acetate buffer at pH 4.0 and immobilized

onto FC4 to a level of ~300 RU, using standard amine coupling chemistry. PBS-P (20

mM phosphate buffer pH 7.4, 2.7 mM KCl, 137 mM NaCl, 0.05 % (v/v) surfactant P20)

was used as the immobilization running buffer. Based on the captured ligand response

values, theoretical Rmax values were calculated. The Rmax values assumed 1:1

interaction mechanism. Overnight kinetics were performed for all analytes binding to the

ligands. The kinetics experiments were performed in the presence of PBS-P. The flow

rate of all analyte solutions was maintained at 50 µL/min. Association and dissociation

times used for all analytes were 60 s and 600 s, respectively. One 20s pulse of Glycine

pH 1.5 was used to regenerate the sensor surface. Injected compound concentrations

were 0 µM, 3.125 nM, 6.25 nM, 12.5 nM, 25 nM, 50 nM, and 100 nM. Sensorgrams

from the overnight kinetics were evaluated using 1:1 steady state affinity or 1:1 kinetics

model fitting.

Page 140: Self-Assembling Bioaffinity Hydrogels for Localized

124

Loading of IgGFITC on pG hydrogel

The binding of IgG to pG bioaffinity hydrogel was measured by determining the

loading of IgG on the hydrogel. Fluorescein (FITC) labeled IgG (Invitrogen) was used as

the model IgG to provide the readout in the form of FITC fluorescence. 11.25 μg (75

pmol) of IgGFITC was added to gel formed with 20 µL of 5 mg/ml EAK (60 nmol) and

24.5 µL of 6.16 μM pG_EAK (150 pmol) (molar ratio of EAK : pG_EAK = 400 : 1).

After incubated for 15 min, the sample was brought up to 100 µL with PBS. The IgGFITC

loaded hydrogel was collected by centrifugation and washed with 1ml of fresh PBS five

times before re-suspended and transferred to black 96 well plate for FITC fluorescence

measurement. Control samples including IgGFITC loaded in hydrogel made of only EAK,

IgGFITC mixed with only pG_EAK, or only IgGFITC were also prepared by replacing EAK

with water and pG_EAK with PBS when either one is missing.

Paper chromatography

The binding of IgG to pG gel was examined using a thin layer chromatography

based on the trapping of IgG on the path way of its migration driven by capillary action

on a chromatography paper (WhatmanTM GE healthcare) and concentration gradient.

Rabbit IgG800s (NOVUS biologicals, 1 mg/ml) were spotted horizontally (numbered 1-3

from left to the right) near the bottom of the paper (leaving about 3 cm to the edge of the

bottom) with different samples including PBS, EAK (5 µL of 5 mg/ml), or EAK and

pG_EAK (5 µL of 5.46 μM) mixture spotted above the IgG spotting positions

corresponding to each IgG spot vertically. After the drying of the spots, the bottom of the

chromatography paper was immersed in a reagent reservoir filled with PBS. Allow the

Page 141: Self-Assembling Bioaffinity Hydrogels for Localized

125

liquid to migrate upward through capillary action and take the paper out of the PBS when

the water front move up to the top of the paper. Let the paper dry completely before

imaging the migration of IgG800 in the path of convective flow using an Odyssey imager

(Li-Cor, Inc., Lincoln, Nebraska) set at low intensity (L2.0) and focus offset of 3.2mm.

Multi-IgGs loading on pG hydrogel

The loading of three different IgGs on the pG hydrogel was verified by detecting

the fluorescence spectra from three different dyes labeled to each of these three IgGs

respectively. Specifically, equal molar amount (25 pmol) of PerCP-Cy5.5 labeled rat

anti-mouse CD4 IgG (BD Pharmingen), PE labeled rat anti-mouse IgG (BD pharmingen)

and FITC labeled mouse IgG (Southern Biotech) were added to EAK and pG_EAK

coassemblies (20 µL of 5 mg/ml EAK mixed with 24.5 µL of 6.16 μM pG_EAK) and

EAK without pG_EAK as the control. After one hour of incubation for the binding to

reach equilibrium, the samples were washed five times with 1ml of fresh PBS and

transferred to black 96 well plate to get the emission spectra from (495 nm to 800 nm)

under the excitation of 488 nm. The combination of the three IgG solutions was also

included as the positive control to show the desirable peak locations as reference.

FITC-IgG self-quenching after loaded to pG hydrogel

The distribution of loaded IgG in pG hydrogel could be speculated by studying

the quenching effect of FITC dye labeled on the loaded IgG. Specifically, 22.5 µL of 0.5

mg/ml mouse FITC-IgG (Southern Biotech) was added to the pG_EAK and EAK

coassemblies (20 µL of 5 mg/ml EAK mixed with 24.5 µL of 6.16 μM pG_EAK)

established at the bottom of black 96-well plate. The fluorescence from FITC-IgG was

Page 142: Self-Assembling Bioaffinity Hydrogels for Localized

126

scanned from 500nm-800nm after excitation at 494 nm and compared to the fluorescence

spectra from the same amount of FITC-IgG in aqueous solution or loaded in EAK gel.

The single point measurement at its emission peak (520 nm) was also scanned separately

for better comparison.

In vitro IgG release

The release kinetics of a fluorescein-labeled IgG (Pierce goat anti-human IgG)

from pG bioaffinity hydrogel were determined by monitoring the cumulative

fluorescence in release medium over 3 months. IgG-loaded hydrogels were established at

the bottom of MicroSert (inner diameter = 4.15 mm, maximum volume = 500 µL) by

instilling a mixture of EAK (60 nmol), pG_EAK (150 pmol) and IgGFITC (75 pmol). The

55 µL mixture was incubated overnight at 37°C before adding 400 µL of PBS containing

0.1% BSA, 0.024% NaN3 as the release medium. The insert was placed in its receptacle

vial and capped to prevent the medium evaporation. The vial was put under 37°C for

incubation. At various time points, 200 µL of the release medium were collected for

measurement using plate reader (Perkin Elmer 1420 Multilabel Counter VICTORTM,

excitation = 485 nm; emission = 535 nm). Aliquots of fresh release medium were added

to restore the system to its initial volume.

Mice and in vivo studies

Six to eight week old female BALB/c or C57BL/6 mice were purchased from

Hilltop Lab Animals (Scottdale, PA) and housed in the Duquesne University Animal

Care Facility certified by the USDA. Animals were handled in accordance with protocols

approved by Institutional Animal Care and Use Committee.

Page 143: Self-Assembling Bioaffinity Hydrogels for Localized

127

Retention after foot pad injection (preliminary test, no replicates)

The injection mixture containing 60 nmol (in 20 µL) of EAK together with 150

pmol of pG_EAK and 75 pmol of Dylight 800 4X PEG conjugated goat anti-rabbit IgG

(Thermo Scientific) was drawn into 25G’ needled syringe and injected subcutaneously

into one foot pad of the mouse while free IgG800 control was injected into the other foot

pad of the same mouse when the mouse was anesthetized with isoflurane (induction 5%;

maintenance 3%). The mouse was sacrificed 2 days after the injection and foot pads

were excised and imaged using an Odyssey imager (Li-Cor, Inc., Lincoln, Nebraska).

The image captured from Odyssey was set at low intensity (L2.0) and focus offset of

3.2mm. This preliminary test was conducted to get a sense of the performance of pG gel

in retaining IgG in foot pad (as tested in dL5 gel) before moving forward to the

investigation of its performance in other subcutaneous location or inflammatory

conditions.

Stability after back injection

Inject onto the back of C57BL/6 mice IgG800 (45 μg in 1 mg/mL goat anti-rabbit

Dylight 800 4X PEG conjugate) or IgG800 displayed by pG hydrogel (80 µL of 10 mg/mL

EAK and 82.4 µL of 14.556 μM pG_EAK) (4 replicates each). The NIR signal at 800nm

on the back was monitored at the local injection site using Li-Cor Pearl live imager

before and over 2 days period after the injection. At the end of day 2, all the mice were

sacrificed for liver, spleen and lymph nodes scan using Odyssey imager at low intensity

(L2.0) and the same focus offset (3.2mm).

Page 144: Self-Assembling Bioaffinity Hydrogels for Localized

128

Stability after undergraft injection

To demonstrate the stability of the bioaffinity hydrogel under inflammatory

conditions, we used a murine allogeneic skin transplantation model (donor: BALB/c;

recipient: C57BL/6) and tracked its undergraft stability over 6 days after injection. The

procedure has been approved by the University IACUC. Six- to eight-week old mice was

purchased as certified pathogen-free from Hilltop Lab Animals (Scottdale, PA). Mice

were quarantined for at least five days prior to use in experiments. Mice between the

ages of 8-12 weeks were randomized into three groups in the transplant experiments.

The procedure entailed harvesting skin grafts by euthanizing donor mice, making incision

at the base of the ear and splitting the dorsal side from the base. The dorsal layer was

rinsed with sterile saline prior to the transplantation to serve as the donor. Then a 0.5 cm

incision was made in the left flank of a C57BL/6 mouse for placement of donor skin.

Excess donor skin was trimmed so that transplant lies flatly within graft bed. The

recipient mouse was wrapped in sterile bandage and secured using a 7mm stainless steel

wound clip (Reflex 7). Recipient mice were recovered under heat lamp while waking up

from anesthesia before placement in a clean cage. Optical imaging of the mice was

scanned using a Pearl Impulse (Li-Cor) small animal imaging system. Pre-injection

images were taken by placing the animals in the instrument chamber. Post injection

images were taken at specified time intervals. To image in the near infrared range, the

mice were kept under anesthesia and exposed to a 700 nm laser, followed by an 800 nm

laser, and then a white light image was taken. Grafting of BALB/c skins onto C57BL/6

mice is a well-characterized model in studying acute rejection of allografts. The full-

MHC (class I and II) mismatch renders consistent rejection in >90% of cases.

Page 145: Self-Assembling Bioaffinity Hydrogels for Localized

129

In this in vivo undergraft stability test, IgG800 (11.24 μg in 1 mg/mL goat anti-

rabbit Dylight 800 4X PEG conjugate), IgG800 displayed by EAK hydrogel (20 µL of 10

mg/mL EAK) or IgG800 displayed by pG hydrogel (20 µL of 10 mg/mL EAK and 5.5 µL

of 54.577 μM pG_EAK) was injected subcutaneously underneath the skin graft (3

replicates each) two days after the transplantation surgery. Signals retained under the

skin graft was recorded in real time using a Li-cor Pearl Impulse Imager before, right

after and 2, 4, 6 days after the injection for each mouse. NIR fluorescence intensity at

different time points was quantified using Li-Cor Pearl imager software.

Dual delivery of Fc-CTLA-4 and Fc-TGF-β1 in skin transplantation

Skin transplantation surgery was performed in the same way as described in the

previous undergraft stability test. 33 μg of TGF-β1 Fc (Chimerigen laboratories, 1

mg/ml) and 33 μg of CTLA-4 (CD152) Fc non-lytic recombinant protein (ProSci, 1

mg/ml) displayed by pG gel (20 µL of 10 mg/ml of EAK mixed with 4.34 µL of 69.2 μM

of pG_EAK) or only the pG gel were injected underneath the skin graft two days after the

surgery. The same injection was repeated 3 days after the first injection. Four days after

the second injection, the mice were euthanized to collect the spleen. The splenocytes

were processed for in vitro re-stimulation with allogeneic antigen presenting cells from

BALB/c spleen and cultured for 3 days before flow cytometry analysis for Treg cell using

mouse regulatory T cell staining kit (eBioscience).

Page 146: Self-Assembling Bioaffinity Hydrogels for Localized

130

Results

Confirmation of pG_EAK molecular mass

pG_EAK recovered from nickel resins were further purified using endotoxin

removal spin column (Pierce®, Thermo Scientific, 0.5ml). The final eluent was analyzed

using denaturing SDS-PAGE and mass spectrometry. Electrophoretic image indicated a

single band near the 35 kDa mark, and MALDI-TOF showed a singly-charged species

with Mr of 25.3 kDa (Figure 4-3), consistent with the theoretical Mr of pG_EAK

(predicted using ExPASy Bioinformatics Resource Portal). In conjunction with

functional studies that will be shown later in this chapter, it was concluded that the

predicted protein pG_EAK was expressed and purified.

Figure 4-3 MALDI-TOF mass spectroscopy of pG_EAK protein.

Page 147: Self-Assembling Bioaffinity Hydrogels for Localized

131

Co-assembly

Gel electrophoresis was used to determine the fraction of pG_EAK incorporated

into excess EAK. The unincorporated portion of pG_EAK retained in the sup was

reflected in the pG_EAK band on the SDS-PAGE gel after intermixing EAK16-II with

pG_EAK at different ratios (Figure 4-4). The ratio of 400:1 for EAK and pG_EAK

resulted in 92% of the latter being incorporated into the final composite. Therefore,

except where stated otherwise, the experiments described in this chapter were conducted

using gels admixed at the 400:1 molar ratio.

The absorbance spectra for different components mixed with Congo red, the dye

found to have a shift in its absorbance peak after binding to fibrils formed by β-sheet

structure, were obtained. After Congo red was added to samples containing pG_EAK

alone, the samples exhibit the same absorbance peak as free Congo red at 486nm,

indicating that pG_EAK does not self-assemble to form fibril itself. In contrast, both

Congo red added to EAK or the mixture of EAK and pG_EAK at ratio of 400:1 exhibit a

red shift in absorbance peak, at 510nm and 506 nm respectively, indicating the fibril

formation in those two samples.

Page 148: Self-Assembling Bioaffinity Hydrogels for Localized

132

Figure 4-4 Coassembly between EAK16-II with pG_EAK (a) SDS-PAGE analysis of

incorporation of pG_EAK into EAK16-II fibrils. The protein bands in image represent

the unincorporated, soluble fraction in the mixture of EAK : pG_EAK coassemblies at

different ratios of the two. (b) Congo red absorbance spectra after added to different

samples containing EAK, pG_EAK alone or the mixture of these two components at the

ratio of 400:1

Injectable composite

The viscosity of the composite materials was determined using a microfluidic-

based measurement where the sample is injected at a constant flow rate through the flow

channel, a rectangular slit constructed of borosilicate glass, of a measuring cell or chip.

Samples of EAK, EAK intermixed with pG_EAK and pG_EAK alone were analyzed

using a VISC module where multiple pressure sensors were mounted within the base to

monitor the pressure drop from the inlet to the outlet of the channel. The viscosity was

calculated based on the shear rate and shear stress at the boundary wall which are

correlated with the pressure drop, geometry of the rectangular slit and the flow rate.

Solution dissolved with only pG_EAK exhibited baseline viscosity, at near that of

water (1 mPa·s), confirming our previous observation that pG_EAK do not self-assemble

Page 149: Self-Assembling Bioaffinity Hydrogels for Localized

133

at the concentrations tested. At shear rate of 4,000 s-1, the viscosity values of EAK and

EAK mixed with pG_EAK registered 2.3 0.56 mPa·s and 2.5 0.24 mPa·s, respectively,

indicating that the materials lack shear resistance and can be injected using conventional

needled syringes. At the low shear rate of 1200 s-1, EAK mixed with pG_EAK deviated

from EAK, at 2.9 0.87 mPa·s and 4.4 0.32 mPa·s, respectively. Yet pG gel holds

sufficient mechanical strength to be suspended and adsorbed to a silica vessel and

remained suspended for at least 24 h (Figure 4-5b).

Figure 4-5 Viscosity profile of pG gel and its physical appearance visualized by Congo

red (a) Viscosity of EAK16-II mixed with pG_EAK (400:1) at different shear rates

compared with pG_EAK, EAK or deionized water alone. (b) pG gel stained with Congo

red at the bottom of silica vessel inverted for 24 hours.

Capture of IgG in pG gels in vitro

The affinity and kinetic constants of IgG for pG_EAK were measured using

surface plasmon resonance (Table 4-2). Sonograms indicate that the fusion polypeptide

Page 150: Self-Assembling Bioaffinity Hydrogels for Localized

134

bound to IgG at high affinity. The association (kon = 9.435 x 105 M-1s-1) and dissociation

(koff = 21.81 x 10-3 s-1) constants were also found in the same magnitudes of documented

for the native proteins. The apparent binding constant (Ka) for mouse IgG has been

reported as strong (1.13 x 108 M-1), with association rate kon = 3.29 x 104 M-1s-1 and

dissociation rate koff = 2.90 x 10-4 s-1 [49]. The Kd (8.8 nM) is close to the Kd tested in

our measurement for mouse IgG (23.1 nM). The differential affinities across the species

documented in the literature were reproduced in that pG_EAK exhibited strong affinities

for goat and rabbit IgG, but relatively weak for mouse IgG1. These results show that

pG_EAK associated and dissociated with IgG through specific interactions.

Table 4-2 Binding and kinetic constants of pG_EAK

Kd (nM) kon (1/Ms) koff(1/s)

Goat IgG polyclonal 0.005 3.658 * 105 2.0 * 10-6

Rabbit IgG polyclonal 0.1 3.225 * 105 3.817 * 10-5

Human IgG polyclonal 4.1 1.273 * 105 5.265* 10-4

Mouse monoclonal IgG1 23.1 9.435 * 105 2.181 * 10-2

To investigate the extent to which pG_EAK gel could display solvent-accessible

Fc-binding sites, an in vitro pull-down assay was used to test capture of IgG molecules.

A 3-fold higher fluorescent intensity was observed in samples in which IgGFITC was

mixed with pG_EAK and EAK compared with EAK alone (Figure 4-6b). Mixing the

antibody with pG_EAK resulted in low amounts of fluorescent aggregates collected by

centrifugation, consistent with the Congo red and viscosity measurements showing that

Page 151: Self-Assembling Bioaffinity Hydrogels for Localized

135

the lack of fibrillization of the polypeptide itself. In addition, association of IgG and pG

gel was examined in paper chromatograph under non-equilibrium conditions (Figure

4-6c). IgG800 accumulated at the pG_EAK/EAK spot in the pathway driven by capillary

action verified the specific association between IgG and protein G.

Figure 4-6 Capture of IgG in pG gels in vitro (a) Schematic depiction of the affinity

hydrogel for retention of IgG and assay procedure. (b) Loading of FITC-labeled antibody

on pG bioaffinity hydrogel compared with that on incomplete systems. (c) paper

chromatography showing IgG800 migration driven by capillary action in paper

chromatography with pG_EAK spot in path of convective flow

As expected, pG_EAK gel can be used to capture three different antibodies in the

same mixture simultaneously (Figure 4-7). The crystal structures of protein G indicate a

single α-helix stabilized by four β-strands. One of the β-strands in protein G appears to

interact with the β-strand in CH1 of IgG. It may be deduced that in pG_EAK a trivalent

(C1, C2 and C3) Fc-binding module displaying independent helix-β-strand structures.

The data indicate that at least one of the C1, C2 and C3 segments was solvent accessible

in the pG_EAK fibrillar gel network.

Page 152: Self-Assembling Bioaffinity Hydrogels for Localized

136

Figure 4-7 Emission scan of three different dye labeled (FITC, PE and PerCP-Cy5.5)

IgGs on the pG hydrogel.

A self-quenching phenomenon was observed for FITC-IgG loaded on pG gel

(Figure 4-8). Fluorescence quenching is the phenomenon where the quantum yield of a

fluorophore is decreased, which is induced by the molecular interaction with quencher

molecules. Self-quenching takes place when the quencher molecule is the same

fluorophore. When the fluorophore molecules are in close proximity to each other, the

manifested fluorescence intensity would be decreased due to self-quenching. The

observation of lower fluorescence intensity in sample where FITC-IgG was loaded to pG

gel compared to the other two samples where the same amount of FITC-IgG was added

to either aqueous solution or EAK gel, indicates that FITC-IgGs were highly

concentrated by the IgG binding domain pG embedded in the gel matrix, causing the self-

quenching effect of the dye molecules labeled to IgG. In contrast, free FITC-IgG

exhibited the highest fluorescence intensity while the fluorescence of FITC-IgG loaded to

non-specific EAK gel lies in between free FITC-IgG and FITC-IgG loaded in pG gel.

The specific binding between FITC-IgG and pG was thereby verified.

Page 153: Self-Assembling Bioaffinity Hydrogels for Localized

137

500 600 700 800 9000

1000

2000

3000

4000

FITC-IgG in pG gel

FITC-IgG in EAK gel

Free FITC-IgG

wavelength (nm)

Flu

ore

scen

ce In

ten

sit

y

Emission at 520 nm

FITC-Ig

G in

pG g

el

FITC-Ig

G in

EAK g

el

Free

FITC-Ig

G

0

10000

20000

30000

40000

Flu

ore

scen

ce In

ten

sit

y

Figure 4-8 Self quenching phenomenon of FITC-IgG loaded on pG gel (a) Emission

fluorescence scan for FITC-IgG established in black 96-well plate together with pG

hydrogel and compared that with free FITC-IgG or FITC-IgG nonspecifically trapped in

EAK gel. (b) Emission fluorescence intensity of FITC-IgG at its maximum emission

wavelength under the same setup as (a).

In vitro IgG release

To understand the microscopic properties of the bioaffinity interaction, release of

antibodies from pG_EAK gel were monitored in vitro over a three-month period (Figure

4-9). The experiment was initiated by mixing IgGFITC with pG_EAK and EAK in a

cylindrical vessel with defined dimensions. In control experiments in which gels were

made with only EAK, 85.2% ± 2.3% of the antibody loaded were released in the first 5

days (Figure 4-9a). In contrast, a slow rate of release was observed with pG gel; by day

5, 2.7% ± 0.5% were released. It was calculated that 82.7% of the IgG remained at the

end of the 12-week period. These data show that the pG_EAK module can be used to

delay the release of antibodies from EAK gels.

Page 154: Self-Assembling Bioaffinity Hydrogels for Localized

138

Figure 4-9 In vitro release of IgG from the gel. (a) Cumulative release of IgGFITC from

pG gels (n = 3) or EAK gels (n = 3). Each time point represents measurement from

triplicate samples in gels setup in parallel. (b) initial release phase fitting to 1D unsteady

state form of Fick’s second law of diffusion: 𝑀𝑡

𝑀∞=(

16 𝐷𝑎𝑝𝑝. 𝑡

𝜋 𝐻2 )0.5

by plotting fractional

release vesus 𝑡0.5 for small values of t.

The apparent diffusivity (Dapp) of IgG in pG gel was estimated by fitting the

initial release to the non-steady state Fickian equation, relating Mt/M∞ as a function of

t0.5. Dapp of IgG in pG gel was determined to be 5.6 x 10-15 m2 s-1, three orders of

magnitude lower than Dapp of IgG embedded in the EAK gel (Table 4-3). The analysis

shows that incorporating pG_EAK served to retain a large fraction of the IgG loaded in

the EAK matrix.

Page 155: Self-Assembling Bioaffinity Hydrogels for Localized

139

Table 4-3 Calculated apparent diffusion coefficients for both IgG displayed by EAK gel

and pG gel and the IgG remaining in the gel at the end of release study.

Dapp (10-10 m2s-1) R2 IgG remained in gel (%)

Control 0.0563 0.83 2.4

System 0.0000564 0.77 82.7

Mice and in vivo studies

pG_EAK gel extends IgG localization in vivo

The in vivo fate of antibodies co-injected with pG_EAK gel components into the

subcutaneous space on the back was tracked in C57BL/6 mice real-time (Figure 4-10a).

Superior local fluorescence was observed in mice injected with IgG800 mixed with

endotoxin removed pG_EAK and EAK (test), compared to mice injected with the

antibody in saline (9-fold on day one and 12-fold on day two) (Figure 4-10b). Systemic

distribution of the antibody was analyzed on day 2, based on fluorescent intensities in the

liver, spleen and lymph nodes recovered from euthanized mice and measured ex vivo.

IgG800 delivered in saline appeared in the liver, spleen, and brachial, inguinal and sciatic

lymph nodes, while the same antibody injected in pG gel was restricted to the immediate

draining branchial lymph nodes (Figure 4-10c). pG_EAK also conferred superior IgG

retention where the components were injected into mouse footpads (Figure 4-10d), which

is consistent with the effects observed in the torso. In addition, no immediate

inflammatory reactions were observed in mice injected with pG_EAK, and low levels of

IL-6 were detected incubating the peptide with freshly isolated mouse splenocytes in

Page 156: Self-Assembling Bioaffinity Hydrogels for Localized

140

vitro (Figure 4-11). These results indicate pG gel impeded the draining and distribution

of IgG deposited at the injection site.

Figure 4-10 IgG retention in subcutaneous space after injection into the back of the mice.

(a) Real time images of C57BL/6 mice injected with free IgG800 (control) and IgG800

displayed by pG bioaffinity hydrogel (system) (n=4). (b) Quantification of the remaining

NIR signal on the back of the mice over 2 days. (c) NIR signal distribution in liver,

spleen and lymph nodes (axillary, inguinal and sciatic) at the end of day 2. (d) Image of

mouse foot pads 2 days after injection with control (free IgG800) and test (IgG800

displayed by pG hydrogel) hydrogel.

Page 157: Self-Assembling Bioaffinity Hydrogels for Localized

141

Figure 4-11 Interleukin 6 production. “EAK” represents cells added with EAK16-II (100

µg/ml) alone. Different concentrations (200 ng/ml, 2 µg/ml and 20 µg/ml) of endotoxin

(LPS) were used as positive controls. Splenocytes were harvested from a naïve C57BL/6

mouse and cultured (106 cells/ml) for 24 h at 37°C in complete RPMI medium

supplemented with 10% fetal bovine serum, and the substances indicated. The BD

Mouse IL-6 ELISA set was used to determine the concentration.

Having established that the self-assembly and bioaffinity interactions took place

in vivo in normal subcutaneous tissues, the capacity of pG_EAK gel to concentrate IgG

in skin allografts was examined. IgG co-injected into skin allografts with pG gel were

retained locally in mice for at least 6 days (Figure 4-12), concentrating in the center of

the wound bed. This resulted in an average 2-fold higher exposure in terms of AUC

compared to IgG delivered in saline control, with 229.1 v.s. 131.6, respectively. The

higher graft retention coincided with reduced distribution to liver, lymph nodes, and

spleen (Figure 4-12c). Thus, the pG gel platform can be deployed in vivo for displaying

antibodies in inflammatory tissues.

Page 158: Self-Assembling Bioaffinity Hydrogels for Localized

142

Page 159: Self-Assembling Bioaffinity Hydrogels for Localized

143

Figure 4-12 Retention of subcutaneously injected IgG800 under inflammatory condition

(skin transplantation). (a) Real time images of C57BL/6 mice injected underneath the

skin graft with free IgG800 (control 1), IgG in EAK gel (control 2) and IgG800 displayed

by pG bioaffinity hydrogel (system) (n = 3). (b) Quantification of the remaining NIR

signal under the skin graft of the mice over 6 days. (c) NIR signal distribution in liver,

spleen and lymph nodes (brachial, axillary and inguinal) at the end of day 6.

Dual delivery of Fc-CTLA-4 and Fc-TGF-β1 in skin transplantation

It has been shown that TGF-β1 and CTLA-4 mutually support

immunosuppression. The capacities of Tregs to regulate naïve CD4+ T cells and DCs are

primarily attributed to TGF-β1 and CTLA-4 [240]. These factors are best function in

concert, and tolerance is acquired only if Tregs accumulate in skin allografts [241]. TGF-

β1, signature cytokine secreted by Tregs, convert naïve T cells to iTregs (induced Treg)

by inducing Foxp3 [242]. In combination with IL-10, TGF-β1 facilitate the generation of

tolerogenic DCs (tDCs) [243]. Host tDCs in turn secret IL-10, thereby driving cycles of

suppressive milieu, promoting tolerance in direct and indirect recognition of alloantigens

Page 160: Self-Assembling Bioaffinity Hydrogels for Localized

144

[244]. Tregs constitutively express on the surface CTLA-4, which competes with CD28

for B7 (CD80/CD86) molecules on DCs [245]. CTLA-4 ligation of B7 downregulates

the costimulatory molecules, and suppresses production of IL-, TNFα and IL-6 in DCs

[246]. Mice lacking the CTLA-4 gene display phenotypes that resemble Foxp3-deficient

mice [247, 248]. We therefore postulate that sustained supply of CTLA-4 and TGF-β1 in

grafts will raise the thresholds of Th1 activation, disrupt Th17 differentiation, increase

Tregs, and generate tDCs [243, 244].

Here we seek to recapitulate Treg functions in a cell-free system by concentrating

Fc-tagged TGF-b and CTLA-4 in skin allografts shortly after transplantation during

which T cells and DCs are polarized [243, 244]. Non-lytic, mouse Fc fusion proteins of

TGF- (TGF-1/Fc, Mr=190 kDa), and CTLA-4 (CTLA-4/Fc, Mr=220) have been

validated functionally in mice [249-254] and are available at high purity with endotoxin

below 0.06 EU/µg (Chimerigen Laboratories, San Diego, CA). While infusing all factors

into patients would lead to detrimental immunosuppression, simply injecting Fc-proteins

into skin grafts will not result in retention, as we have shown [202]. The targeted,

localized approach proposed will increase the bioavailability of these factors at the

transplant site, thereby sparing patients from systemic toxicities. The clustered factors in

protein G gel is expected to engage signaling pathways more efficiently by cross-linking

cell surface receptors [255]. In this preliminary in vivo study for the demonstration of pG

gel application, Fc fused CTLA-4 and TGF-β1 was multiplexed on pG gel and

administered underneath the skin graft after the transplantation surgery. An upregulation

of regulatory T cells was observed in the flow cytometry analysis for the splenocytes

collected from treated mice and restimulated with allogeneic antigen-presenting cells

Page 161: Self-Assembling Bioaffinity Hydrogels for Localized

145

(Figure 4-13). The increase in the Treg population suggests sustained signaling of

CTLA-4 and TGF-1 in the transplant microenvironment, most likely rendered by the

affinity capture mechanism of the pG gel. Additional experiments will be necessary to

confirm this preliminary finding.

Figure 4-13 representative flow cytometry plots of in vitro allogeneic re-stimulated

splenocytes from mice treated with pG gel displayed Fc-CTLA and Fc-TGF-β1 (exp) or

only pG gel (control) after allo skin transplantation surgery. (a) forward and side scatter

plots for the gating of lymphocyte population P1; (b) histogram for the gating of CD4

positive population M2; (c) dot plots showing Foxp3-A CD25 double positive cells gated

on P1 and M2 indicating induced Treg population.

Page 162: Self-Assembling Bioaffinity Hydrogels for Localized

146

Discussion

The data presented in this chapter support the notion of generating antibody

depots in situ by intermixing IgG and pG_EAK with molar excess of EAK. The system

was designed around the concept of multivalent display of antibodies in vivo, employing

a direct, specific and reversible binding interaction between the materials and antibodies.

Derived from the partial sequence of Streptococcal protein G, consisting of the IgG-

binding C and D domains, without the albumin binding A and B domains, pG_EAK has a

molecular weight of 25.3 kDa, a big reduction in molecular weight from dL5 complex

with dL5_EAK of 28.4 kDa and pAGMG of 51 kDa. The non-covalent interactions were

also reduced from two to one. This simplified system has the potential to increase the

translational prospect due to the reduction of potential immunogenicity concerns and

modeling complexity.

The platform can be armed with multiple therapeutic antibodies, thereby enabling

targeting distinct pathways simultaneously in pathogenic tissues and their draining lymph

nodes. In combination therapies, the prolonged concentration of antibodies in draining

lymph nodes would enhance the synergistic impacts on multiple cellular and molecular

entities driving pathogeneses. The novelty of the system rests on the extent to which an

antibody can be concentrated by varying the ratio of IgG and pG_EAK in the gels. The

simplified design facilitates the modeling of IgG release from the system by using

reaction-diffusion model proposed by Lin and Metters [17, 256]. Two release

mechanisms, the diffusion of the IgG from the hydrogel and the specific affinity

interaction between the IgG and protein G, are controlling the release rate of IgG from

the gel system. This is manifested by three important parameters: IgG diffusivity (D),

Page 163: Self-Assembling Bioaffinity Hydrogels for Localized

147

IgG-pG dissociation constant (Kd) and IgG debinding rate constant (Koff). Specifically,

the release of free IgG from the bioaffinity pG gel is governed by Fick’s second law of

diffusion and binding affinity between pG and IgG. Kd determines the fraction of bound

and free IgG in the gel throughout the release process and Koff determines the debinding

rate. The following equation could be used to describe the change in the IgG

concentration within the hydrogel due to both diffusion and reversible binding to pG

where 𝑥 is the spatial coordinate in the gel perpendicular to the gel surface (assuming

one-dimensional release) and t is the time.

𝜕[𝐼𝑔𝐺]

𝜕𝑡= 𝐷 ∗

𝜕2[𝐼𝑔𝐺]

𝜕𝑥2− 𝑘𝑜𝑛[𝐼𝑔𝐺][𝑝𝐺] + 𝑘𝑜𝑓𝑓[𝐼𝑔𝐺 − 𝑝𝐺]

With specific values of these constants possibly obtained from SPR experiments

for different antibody species and subtypes (example evidenced in Table 4-2), a time-

dependent distribution of [IgG], [pG] and [IgG-pG] within the gel should be able to be

generated for a given IgG through numerical solution of the equation using finite-

difference method when the initial and boundary conditions describing the concentration

changes of free IgG, unbound ligand and bound IgG-pG during the period of IgG release

from the gel matrix is provided.

The discordance between the in vitro release kinetics (Figure 4-9) and the in vivo

retention profiles of IgG (Figure 4-10 and Figure 4-12) in pG gel can be attributed to

several factors. The ambient movements of the animals can disrupt the integrity of the

localized soft gel, thereby increasing the rate of release. The lack of a strong buffering

system in the subcutaneous space may allow the pH to drift, perturbing the affinities of

Page 164: Self-Assembling Bioaffinity Hydrogels for Localized

148

pG for IgG in inflammatory tissues acidified by lactate [257]. The ionic strength may be

depleted in vivo compared to the in vitro setting as the former may not be fully hydrated.

pG_EAK and EAK may be degraded through proteolysis in the extracellular space and in

intracellular organelles upon phagocytosis by macrophages, neutrophils and/or dendritic

cells. Multiple cleavage sites are embedded within the pG_EAK sequence for peptidase-

cleavage. The low viscosity of the gel is conducive to such cellular uptake. Eventually

the materials will disintegrate after approximately 21 days [21], most likely due to a

combination of proteolytic degradation and macrophage phagocytosis [228].

For macromolecules such as IgG, lymphatic uptake is the main route of

absorption from subcutaneous depots into the systemic circulation. Preclinical studies

indicate that diffusion of IgG through the extracellular matrix (ECM) is the rate-limiting

step, given the relatively rapid transport from the lymph network to blood circulation. In

the ECM IgG are degraded by proteolytic enzymes [68]. This elimination pathway is

compensated in part by endosomal recycling of IgG via neonatal Fc receptors (FcRn)

expressed by local phagocytes, in which a fraction of the administered IgG is protected

from proteases. This is evidenced in that high doses of IgG (e.g. 40 mg/kg), however,

can saturate FcRn in local tissues, thereby increasing the fraction vulnerable to

proteolysis [69]. Mager and coworkers have demonstrated this by monitoring

subcutaneous-administered rituximab in rats, in which they observed a reduction of

bioavailability from 31% to 18% corresponding to increasing the dose from 10 mg/kg to

40 mg/kg [258]. We postulate that the pG_EAK gel depot mimics the protective effects

of FcRn recycling in limiting the diffusion of IgG through the ECM. The system may be

Page 165: Self-Assembling Bioaffinity Hydrogels for Localized

149

analogous to biomolecular condensates discovered in eukaryotic cells, phase-separated in

the cytoplasm, concentrating proteins and nucleic acids in a small finite volume [259].

Conclusion

The second generation bioaffinity hydrogel built on pG_EAK was demonstrated

to be a simpler IgG local delivery platform involving less binding interactions in the

design while retaining stable IgG binding capability to provide prolonged retention of

IgG at both normal and inflammatory subcutaneous space. The release rate could be

modulated by varying the ratio of IgG and pG in the formulation. The delivery of Fc

fused immunosuppressive cytokines by pG gel underneath transplanted allogeneic skin

graft has the potential to prolong the survival of the graft skin through upregulation of

immunosuppressive T cell populations—regulatory T cells, which generally

downregulate the induction and proliferation of effector T cells.

Page 166: Self-Assembling Bioaffinity Hydrogels for Localized

150

Appendix:

Content for appendix includes 2 sections. The first section is the continuous

miniaturization of IgG binding domain in the bioaffinity hydrogel design and the

preliminary data generated regarding its application in neutrophil inhibition. The second

part is focused on the application of these bioaffinity hydrogels as three-dimensional cell

culture scaffold. The construction of thymus epithelium cells aggregates could be

facilitated by the gels to promote T cell development in immunomodulatory deficient

animals.

I. Miniaturization of IgG Binding Domain in the Bioaffinity Hydrogel Design

Z34c functionalized EAK gel

The design of bioaffinity hydrogels has undergone progressive miniaturization in

our lab, in using lower molecular weight Fc-binding domains, and reducing the number

of non-covalent interactions, in order to simplify the optimization and reduce the

potential immunogenicity. Toward this goal, the construction and characterization of

pG_EAK has been described in chapter 4. Built on top of this pG_EAK design was the

construction of another fusion protein Z34c_EAK for bioaffinity hydrogel formation

(Figure 5-1). Z34c_EAK consists of the B domain mimetic (4.18 kDa) of the

Staphylococcal protein A (MW = 42 kDa) [260]. Generated through phage display

randomization and point mutations from the earlier protein A mimetic Z38, Z34c is a

disulfide-linked cyclic peptide consisting of two helices, analogous to helix-1 and helix-2

of domain B in the native protein A [261]. The binding constants for IgG of Z34c and

native protein A are comparable, both at approximately 20 nM [260], yet Z34c is

Page 167: Self-Assembling Bioaffinity Hydrogels for Localized

151

significantly smaller in size (4.18 kDa) compared to protein G domain (22 kDa) in

pG_EAK. Z34c_EAK was expressed in E.coli as well and Z34c is predicted to have low

immunogenicity, with few potential MHC-II ligands [92] compared with other Fc binding

domains (Table 5-1). Specifically, Z34c_EAK has only one predicted ligand with

smaller than 10th percentile while native protein A has 88 and pG_EAK has 62 predicted

ligands with smaller than 10th percentile and smaller percentile rank indicates higher

potential immunogenicity.

Figure 5-1 Diagrams showing (a): the construct of Z34c_EAK and (b) coassembled Z34c

gel displaying IgG molecules.

Page 168: Self-Assembling Bioaffinity Hydrogels for Localized

152

Table 5-1 immunogenicity prediction for different IgG binding domains based on MHC-

II ligand scanning using immune epitope database (IEDB)

Allele

Predicted

ligands

10th

percentile

Peptide

Percentile

rank

(consensus

method)

Native

protein G

HLA-

DRB1*0401 137 VSDYYKNLINNAKTV 0.41

Native

protein A

HLA-

DRB1*0401 88 QNAFYQVLNMPNLNA 0.49

pG_EAK HLA-

DRB1*0401 62 GKTLKGETTTKAVDA 1.88

Z34c_EAK HLA-

DRB1*0401 1 QRRFYEALHDPNLINE 9.71

Intermixing Z34c_EAK with molar excess of EAK resulted in gels that bind IgG

antibodies which was reflected by the superior PE fluorescence signal from PE-IgG

loaded in Z34c gels compared to other incomplete systems as well as the colocalization

of IgG with the hydrogel fibril under fluorescence microscope by PE fluorescence from

IgG and ThT fluorescence stained with EAK fibrils (Figure 5-2).

Page 169: Self-Assembling Bioaffinity Hydrogels for Localized

153

Figure 5-2 IgG loading on Z34c bioaffinity hydrogel and its colocalization with gel fibrils

(a) Loading of PE-labeled antibody on Z34c bioaffinity hydrogel compared with that on

incomplete systems. (b) florescence microscopy showing the colocalization of loaded

IgG and gel fibrils reflected by the overlapping signal from PE fluorescence from labeled

IgG and ThT fluorescence after binding to gel fibrils.

In addition, Z34c is effective in increasing retention of IgG at host-graft interface

in murine skin transplantation model after subcutaneous injection of IgG loaded on

Z34c_EAK/EAK composite underneath the skin graft. IgG displayed by Z34c gel

retained better than free IgG control which diffuses away from the injection site fast.

Relatively strong signal remained at the center of the graft for Z34c gel displayed IgG 2

days after the injection at the host-graft interface of BALB/c donor skin on C57BL/6

recipient mouse Figure 5-3. Thus, we stabilized the antibody on the graft bed through

Z34c gel.

Page 170: Self-Assembling Bioaffinity Hydrogels for Localized

154

Figure 5-3 The fluorescence decay over 2 days after injection of IgG800 with (test) or

without (control) Z34c gel at the interface of host-skin graft.

Inhibition of neutrophil trafficking in skin transplant model using anti-neutrophil

surface marker IgGs displayed by Z34c gel

One possible application of Z34c gel is to inhibit neutrophil trafficking in skin

transplant model. Neutrophil trafficking inhibition at the host-graft interface in allograft

transplantation patients may play an important role in prolonging survival of skin

allografts. It could potentially limit fluid loss, prevent infection, decrease pain and

promote wound healing. We expect that neutrophil migration be impeded by antibody-

induced oligomerization of the cell-specific surface molecules Ly6G and CD11b. The

antibodies against those surface molecules (anti-Ly6G and anti-CD11b IgGs) can be

displayed by our Z34c gel. The 4.16 kDa mimetic of Staphylococcal protein A (Z34c), is

a disulfide-linked cyclic peptide consisting of two helices. It mimicks helix-1 and helix-2

of domain B in the native protein A and possesses comparable affinity toward IgG to

native protein A, but retains low immunogenicity and reduced steric hindrance during

Page 171: Self-Assembling Bioaffinity Hydrogels for Localized

155

coassembly with EAK16-II due to its relatively low molecular weight after fused to EAK

(6.9 kDa). Therefore, by displaying anti-Ly6G and anti-CD11b IgGs on Z34c gel,

activated neutrophils can be intercepted locally and its infiltration to draining lymph

nodes will be reduced, thus suppressing the following immune response and prolonging

allograft survival (Figure 5-4).

Figure 5-4 Schematic diagram showing the design of anti-neutrophil Z34c gel delivered

at the interface of host-graft in skin transplant model

The infiltration of neutrophils at the graft site were verified both ex vivo and in

situ by detecting neutrophil elastase activity as the indicator, assuming more neutrophil

infiltration would cause higher production of neutrophil elastase at the graft site. For ex

vivo assay (Figure 5-5a), neutrophil elastase was extracted from transplanted skin graft

two days after transplantation and incubated with the fluorogenic elastase substrate

MeOSuc-AAPV-AMC. The fluorescence intensity generated after incubation was

compared to that generated from elastase extracted from naïve ear skin without

transplantation. Higher elastase activity was detected in skin graft recovered from a

Page 172: Self-Assembling Bioaffinity Hydrogels for Localized

156

transplanted mouse compared to fresh ear skin reflected by higher fluorescence intensity

detected from the former at varying time points resulting in a higher slope of the line

fitting the data, which indicates a higher catalytic rate facilitated by more neutrophil

elastase. For in situ assay (Figure 5-5b), local administration of NE680, the agent which

contains an elastase substrate that relieves the quenching of two conjugated dye when

cleaved, was performed on the graft bed of transplanted mouse one day after the

transplantation surgery. The yielding fluorescence in the near infrared range on the graft

bed after removal of the skin graft was monitored for 2 days and compared to the control

fluorescence generated on the open wound on the same location without transplantation.

Higher elastase activity detected on allo-transplanted mouse compared to control open

wound is in agreement with the in situ assay.

Figure 5-5 verification of neutrophil infiltration after skin transplantation surgery at the

graft site (a) ex vivo assay determining the fluorescence intensity of the fluorogenic

neutrophil elastase substrate after incubation with transplanted skin graft or naïve ear skin

without transplantation. (b) in situ assay determining the near infrared signal after

delivering signal generating neutrophil elastase substrate NE680 to the transplantation

site and compared to open wounds without allo skin transplantation surgery.

Page 173: Self-Assembling Bioaffinity Hydrogels for Localized

157

After the elevated neutrophil infiltration level at the graft site after skin

transplantation surgery was confirmed, the effect of anti-neutrophil antibody treatment

displayed by the Z34c bioaffinity module was tested in vivo by injecting neutrophil

specific, NIR fluorescent imaging agent cFLFLF-PEG-Cy7, which binds to formyl

peptide receptor on activated neutrophil, through tail vein the second day after

transplantation surgery and anti-neutrophil antibody treatment. The signal accumulation

at the graft site was compared to the control setup with the administration of only Z34c

gel. From the in vivo imaging scan, less neutrophil accumulation at transplantation site

was observed in treated mouse, suggesting that this treatment could effectively impede

continuous neutrophil infiltration after immobilizing first wave of neutrophil infiltration.

Figure 5-6 Graft homing-neutrophil impediment at host-graft interface (a) diagram

showing the treatment and injection regime after skin transplantation surgery. (b) NIR

signal tracking after injection of neutrophil specific, NIR fluorescent imaging agent

cFLFLF-PEG-Cy7.

Meanwhile, this strategy can be coupled with anti-neutrophil elastase particles to

counter tissue-destructive mechanism elicited by neutrophil elastase. Alvelestat (AZD

9668), a selective elastase inhibitor, can be loaded in PLGA particles and the particles

could be trapped into the EAK gel matrix to work synergistically with the displayed

Page 174: Self-Assembling Bioaffinity Hydrogels for Localized

158

antibodies to constrain neutrophil effect. The EAK gel could trap the particles and

slowdown the release of the particles. Transplant necrosis is expected to be reduced

through this method. DiR dye loaded PLGA particles was used as the model particles to

demonstrate their entrapment effect in EAK gels. The NIR dye DiR (ex/em:

750nm/780nm) can provide a readout signal representative of the particle location. The

entrapment of DiR loaded PLGA particles in EAK was successfully verified by

comparing the NIR signal distribution in systems with or without EAK. A localized and

concentrated NIR signal can be observed in system incorporating EAK (Figure 5-7).

Figure 5-7 Odyssey scan (800nm channel) for particle distribution before and different

time points after the addition of release medium to the system. DiR particles were set up

at the bottom of the vials initially with (left) or without (right) EAK gels (first panel) and

200 ul of PBS was added to each sample before inversion of the vials (system

incorporating EAK gel shows on the right side for later time points)

This is the demonstration of potential combination of anti-neutrophil antibodies

treatment and anti-neutrophil elastase therapy facilitated by miniaturized bioaffinity

hydrogel and small molecule weight drug incorporated particles trapped in the gel

respectively. The experiments described in this section are mainly preliminary data, most

of which are still lack of replicates. Further verification of the results would be required

in the future by adding replicates. In addition, the proposed combination use of

bioaffinity hydrogel and particles is proof of concept where the entrapment of particles in

Page 175: Self-Assembling Bioaffinity Hydrogels for Localized

159

the gel was realized by using dye loaded particles. The effectiveness of the

combinational therapy needs further investigation by using real drug loaded particles.

Page 176: Self-Assembling Bioaffinity Hydrogels for Localized

160

II. Bioaffinity Hydrogels for Cell Therapeutics

3D cell culture facilitated by hydrogels

Another application of the functionalized EAK based hydrogels is to facilitate

three-dimensional (3D) cell culture for cell therapeutics in tissue engineering. 3D cell

culture facilitated by hydrogels is a commonly exploited strategy in tissue engineering

field to provide the cells with the mimetic microenvironment when they actually reside in

tissue. Hydrogel affords necessary geometry and provide similar environment to

extracellular matrix (ECM) to facilitate cell survival or proliferation, allowing the cells to

behave more reflective of their in vivo response [262]. In addition, the spatial

organization of the cell surface receptors involved in the interactions of neighboring cells

could be influenced and physical constraints will be exerted to the cells with the hydrogel

facilitated 3D cell culture. The potential application of the bioaffinity hydrogel was

explored in the generation of mini functional thymic units which is essentially 3D culture

of thymic epithelial cell (TEC) clusters. The TEC clusters lie in the stroma of the

thymus, the essential gland for adaptive immunity which is responsible for continuous

generation of new T-lymphocytes to effectively respond to evading pathogens, while

remaining unresponsive to self. The epithelial cells in thymus is unique in that their

functional and survival are relying on their 3D organization because TECs cultured in a

2D monolayer display altered patterns of expression of key genes for TEC function and

growth, thereby losing their ability to support T cell development [263].

Page 177: Self-Assembling Bioaffinity Hydrogels for Localized

161

Construction of mini thymus unit through bioaffinity hydrogel assisted 3D

aggregates of thymus epithelium cells (TECs)

The construction of mini artificial thymic units to support the development of

mature T cells in immunomodulatory disorders can be achieved by 3D aggregation of

thymus epithelium cells (TECs). Formation of 3-D TEC clusters facilitated by anti-

EpCAM (TECs surface marker) IgG displayed EAK gel helps to retain the molecular

properties of those cells, thus generating a functional mini thymus and supporting

functional T cells development. Anti-EpCAM antibodies could be displayed by any Fc-

binding domain functionalized EAK gel described previously including Z34c, protein G

(pG), or protein A/G (pAG) either attached to EAK sequence directly or with the help of

adaptor complex formation. Figure 5-8 shows an example of one design incorporating

coassembly of EAK16-II and its histidinylated analogue EAKIIH6, and the formation of

adaptor complexes (pA/G loaded with both anti-His and anti-EpCAM IgGs) [264].

Figure 5-8 Schematic drawing showing the strategies to generate the TECs mini thymus

unit. The tri-component adaptor complex was formed by mixing anti-EpCAM antibody

Page 178: Self-Assembling Bioaffinity Hydrogels for Localized

162

(Ab), anti-His Ab and recombinant protein A/G at 3:3:1 ratio. Isolated thymic epithelial

cells (TECs) were stained with the adaptor complexes, and mixed with EAK16-II and

EAKIIH6 mixture in 10% sucrose. Gelation was initiated by adding complete RPMI-10

medium to the cell material mixture.

As a demonstration of bioaffinity hydrogel application in cell therapeutics,

primary medulla thymic epithelial cells were delivered in an anti-EpCAM antibody-

armed bioaffinity gel into nude mice. The cells are expected to congregate into

functional units in vivo, leading to emergence of CD4 positive and CD8 positive T cells

in the mice. Toward this goal, the 3D mini thymus unit was constructed in vitro by using

the bioaffinity hydrogels described in Figure 5-8 to provide geometrical organization of

the TECs in the 3D gel scaffold, thereby helping to establish proper TEC-TEC and TEC-

thymocyte interaction for successful thymopoiesis.

In this design, the amphiphilic peptide EAK16-II co-assembles with its C-

terminal histidinylated analogue EAKII-H6 (AEAEAKAKAEAEAKAKHHHHHH) to

generate the Histidine (His)-tags integrated, ionic strength responsive bioaffinity

hydrogel. This functionalized bioaffinity 3D scaffold provides a docking mechanism,

with antibodies of interest being anchored to the gel with the help of a linker complex

comprised of anti-His antibodies and the Fc-binding, recombinant protein A/G. It was

utilized to induce 3D aggregation of TECs for in vitro culture through the display of anti-

EpCAM IgGs which target the surface marker of TECs. This bioaffinity hydrogel design

based on EAKII-H6 was systemically characterized by previous member in our lab and is

out of the scope of this dissertation. However, its application in the cell therapeutics was

explored here since the same principle could be applied to all the bioaffinity hydrogel

designs described in this dissertation.

Page 179: Self-Assembling Bioaffinity Hydrogels for Localized

163

The high affinity interaction between the His-tag embedded in the bioaffinity

hydrogel matrix and the anti-His IgG served as a molecular bridge to anchor the EpCAM

expressing TECs on the EAK16II/EAKIIH6 structure through another high affinity

interaction between protein A/G and anti-EpCAM/anti-His IgGs. TECs formed clusters

of 20-30 cells in the hydrogel matrix as shown in Figure 5-9a, suggesting that the

EAK16-II and EAKIIH6 bioaffinity hydrogel system could induce the aggregation of

TECs. The viability of TECs embedded in the EAK16-II/EAKIIH6 bioaffinity hydrogel

was examined using calcein-AM and ethidium homodimer-1 (EthD-1) staining. The

presence of green calcein-AM signal indicates live cells with intracellular esterase

activity, whereas red-fluorescent EthD-1 is retained in dead cells with compromised

plasma membrane integrity. Most of the TECs remained alive after 3 days culture within

the gel as shown in Figure 5-9b. These image results, together with detection of

expression of genes specific to thymic epithelium (EpCAM) and different TEC subsets

(Krt5 and Krt14 for mTECs and Krt8 for cTECs) as well as transcriptions of TEC-

specific transcription factors (FoxN1 and Tbata) and expression of signaling molecules

(CCL25) important for lympho-stromal interaction through RT-PCR, suggest that TECs

embedded in the bioaffinity hydrogel as clusters not only can maintain their thymic

epithelium identity in vitro, but also be able to retain their functionalities to support T cell

development in vivo. The EAK16-II/EAKIIH6 based mini-thymus units were

transplanted underneath the kidney capsules of athymic B6.FoxN1-/- nude mice and flow

cytometry analyses revealed the presence of circulating CD45+CD3+ T cells 5 weeks post

operation. Both CD3+CD4+ and CD3+CD8+ T cells were found to be present in

secondary lymphoid organs. Those T cells are capable of reacting to alloantigens but not

Page 180: Self-Assembling Bioaffinity Hydrogels for Localized

164

syngeneic antigens which was reflected by their proliferating response after mixed

leukocyte reaction (MLR) assay where drastic proliferation was observed for T cells

mixed with allogeneic antigen presenting cells while only limited, background levels of

proliferation were observed for T cells when syngeneic antigen presenting cells were

supplemented. Therefore, the functionality of the mature T cells in the thymus

reconstructed B6. Nude mice was confirmed to be similar to that of wild-type B6 mouse

while staying self-tolerant. The RT-PCR experiments in verifying the gene, transcription

factors and signaling molecules expression levels in vitro, and flow cytometry analyses

for in vivo T-lymphopoesis verification after transplantation of mini thymus units into

athymic mice, were conducted by our collaborators at Allegheny General Hospital and

documented in the paper published on clinical immunology. Details of those experiments

could be found in reference [264].

Figure 5-9 Generation of TECs mini thymus complex in vitro. (a) representative phase

contrast microscopic image (50X) of TEC aggregations (arrows) in EAK16-II/EAKH6

hydrogel. (b) representative fluorescent microscopic images of TEC aggregates in

EAK16-II/EAKH6 hydrogel cultured in vitro for 3 days. Live cells were discriminated

from the dead cells by their intracellular esterase activities to generate green fluorescen

calcein-AM (green) and their capabilities to exclude the red fluorescent ethidium

homodimer-1 (EthD-1, red) from entering the nucleus. Left panel, low magnification

image (100X) of the TEC aggregates (white arrows). Right panel, high magnification

image (400X) of the TEC aggregates.

Page 181: Self-Assembling Bioaffinity Hydrogels for Localized

165

Taken together, it was demonstrated that TECs anchored to EAK16-II/EAKIIH6

bioaffinity hydrogel with anti-EpCAM IgG containing tri-component adaptor complex

could self-assemble into 3D aggregates. This 3D configuration could retain, to a certain

degree, the properties of the TECs in vitro, and enable them to function as mini thymus

units to support the development of mature T cells when engrafted into athymic nude

mice. The construction of mini thymus units, which is essentially TECs organized in the

3D configuration, mimicking TECs cultured in the thymus stroma, could be extended to

other bioaffinity hydrogel designs, including the ones described in chapter 3 and chapter

4.

Page 182: Self-Assembling Bioaffinity Hydrogels for Localized

166

References

[1] C. A. Janeway, Jr., P. Travers, et al., Immunobiology: The Immune System in Health

and Disease, (2001).

[2] B.J. Underdown, J.M. Schiff, Immunoglobulin A: Strategic Defense Initiative at the

Mucosal Surface, Annual Review of Immunology, 4 (1986) 389-417.

[3] R. Geisberger, M. Lamers, G. Achatz, The riddle of the dual expression of IgM and

IgD, Immunology, 118 (2006) 429-437.

[4] D.W. Grainger, Controlled-release and local delivery of therapeutic antibodies, Expert

Opinion on Biological Therapy, 4 (2004) 1029-1044.

[5] A. Ferrante, L.J. Beard, R.G. Feldman, IgG subclass distribution of antibodies to

bacterial and viral antigens, The Pediatric Infectious Disease Journal, 9 (1990) 516-524.

[6] A.B. Tolstrup, T.P. Frandsen, S. Bregenholt, Development of recombinant human

polyclonal antibodies for the treatment of complex human diseases, Expert Opin Biol

Ther, 6 (2006) 905-912.

[7] N.S. Lipman, L.R. Jackson, L.J. Trudel, F. Weis-Garcia, Monoclonal Versus

Polyclonal Antibodies: Distinguishing Characteristics, Applications, and Information

Resources, ILAR Journal, 46 (2005) 258-268.

[8] H.H. Cai, Therapeutic Monoclonal Antibodies Approved by FDA in 2016, MOJ

Immunology, 5 (2017).

[9] https://www.accessdata.fda.gov/scripts/cder/daf/index.cfm. 29 Mar. 2018

[10] J.R. Adair, Engineering antibodies for therapy, Immunological reviews, 130 (1992)

5-40.

[11] M. Suzuki, C. Kato, A. Kato, Therapeutic antibodies: their mechanisms of action

and the pathological findings they induce in toxicity studies, Journal of Toxicologic

Pathology, 28 (2015) 133-139.

[12] M.A. Postow, R. Sidlow, M.D. Hellmann, Immune-Related Adverse Events

Associated with Immune Checkpoint Blockade, New England Journal of Medicine, 378

(2018) 158-168.

[13] G. Tiwari, R. Tiwari, B. Sriwastawa, L. Bhati, S. Pandey, P. Pandey, S.K.

Bannerjee, Drug delivery systems: An updated review, International Journal of

Pharmaceutical Investigation, 2 (2012) 2-11.

[14] C. Bee, Y.N. Abdiche, J. Pons, A. Rajpal, Determining the binding affinity of

therapeutic monoclonal antibodies towards their native unpurified antigens in human

serum, PLoS One, 8 (2013) e80501.

[15] S. Mohammadi-Samani, B. Taghipour, PLGA micro and nanoparticles in delivery of

peptides and proteins; problems and approaches, Pharmaceutical Development and

Technology, 20 (2015) 385-393.

[16] H.K. Makadia, S.J. Siegel, Poly Lactic-co-Glycolic Acid (PLGA) as Biodegradable

Controlled Drug Delivery Carrier, Polymers, 3 (2011) 1377-1397.

[17] C.C. Lin, A.T. Metters, Hydrogels in controlled release formulations: network

design and mathematical modeling, Adv Drug Deliv Rev, 58 (2006) 1379-1408.

[18] L. Yu, J. Ding, Injectable hydrogels as unique biomedical materials, Chemical

Society Reviews, 37 (2008) 1473-1481.

Page 183: Self-Assembling Bioaffinity Hydrogels for Localized

167

[19] Y. Zheng, Y. Wen, A.M. George, A.M. Steinbach, B.E. Phillips, N. Giannoukakis,

E.S. Gawalt, W.S. Meng, A peptide-based material platform for displaying antibodies to

engage T cells, Biomaterials, 32 (2011) 249-257.

[20] Y. Wen, H.R. Kolonich, K.M. Kruszewski, N. Giannoukakis, E.S. Gawalt, W.S.

Meng, Retaining Antibodies in Tumors with a Self-Assembling Injectable System,

Molecular Pharmaceutics, 10 (2013) 1035-1044.

[21] Y. Wen, S.L. Roudebush, G.A. Buckholtz, T.R. Goehring, N. Giannoukakis, E.S.

Gawalt, W.S. Meng, Coassembly of amphiphilic peptide EAK16-II with histidinylated

analogues and implications for functionalization of beta-sheet fibrils in vivo,

Biomaterials, 35 (2014) 5196-5205.

[22] Y. Wen, W. Liu, C. Bagia, S. Zhang, M. Bai, J.M. Janjic, N. Giannoukakis, E.S.

Gawalt, W.S. Meng, Antibody-functionalized peptidic membranes for neutralization of

allogeneic skin antigen-presenting cells, Acta Biomaterialia, 10 (2014) 4759-4767.

[23] S. Zhang, T. Holmes, C. Lockshin, A. Rich, Spontaneous assembly of a self-

complementary oligopeptide to form a stable macroscopic membrane, Proceedings of the

National Academy of Sciences, 90 (1993) 3334-3338.

[24] S. Zhang, T.C. Holmes, C.M. DiPersio, R.O. Hynes, X. Su, A. Rich, Self-

complementary oligopeptide matrices support mammalian cell attachment, Biomaterials,

16 (1995) 1385-1393.

[25] P. Homayoun, T. Mandal, D. Landry, H. Komiskey, Controlled release of anti-

cocaine catalytic antibody from biodegradable polymer microspheres, The Journal of

pharmacy and pharmacology, 55 (2003) 933-938.

[26] C. Keyes-Baig, J. Duhamel, S.-Y. Fung, J. Bezaire, P. Chen, Self-Assembling

Peptide as a Potential Carrier of Hydrophobic Compounds, Journal of the American

Chemical Society, 126 (2004) 7522-7532.

[27] S. Lu, H. Wang, Y. Sheng, M. Liu, P. Chen, Molecular binding of self-assembling

peptide EAK16-II with anticancer agent EPT and its implication in cancer cell inhibition,

Journal of Controlled Release, 160 (2012) 33-40.

[28] S. Koutsopoulos, L.D. Unsworth, Y. Nagai, S. Zhang, Controlled release of

functional proteins through designer self-assembling peptide nanofiber hydrogel scaffold,

Proc Natl Acad Sci U S A, 106 (2009) 4623-4628.

[29] S. Koutsopoulos, S. Zhang, Two-layered injectable self-assembling peptide scaffold

hydrogels for long-term sustained release of human antibodies, J Control Release, 160

(2012) 451-458.

[30] S. Koutsopoulos, L.D. Unsworth, Y. Nagai, S. Zhang, Controlled release of

functional proteins through designer self-assembling peptide nanofiber hydrogel scaffold,

Proceedings of the National Academy of Sciences, 106 (2009) 4623-4628.

[31] S.E. Sakiyama-Elbert, J.A. Hubbell, Controlled release of nerve growth factor from

a heparin-containing fibrin-based cell ingrowth matrix, J Control Release, 69 (2000) 149-

158.

[32] S.E. Sakiyama-Elbert, J.A. Hubbell, Development of fibrin derivatives for controlled

release of heparin-binding growth factors, J Control Release, 65 (2000) 389-402.

[33] S.J. Taylor, J.W. McDonald, 3rd, S.E. Sakiyama-Elbert, Controlled release of

neurotrophin-3 from fibrin gels for spinal cord injury, J Control Release, 98 (2004) 281-

294.

Page 184: Self-Assembling Bioaffinity Hydrogels for Localized

168

[34] M.D. Wood, S.E. Sakiyama-Elbert, Release rate controls biological activity of nerve

growth factor released from fibrin matrices containing affinity-based delivery systems, J

Biomed Mater Res A, 84 (2008) 300-312.

[35] M.D. Wood, D. Hunter, S.E. Mackinnon, S.E. Sakiyama-Elbert, Heparin-binding-

affinity-based delivery systems releasing nerve growth factor enhance sciatic nerve

regeneration, J Biomater Sci Polym Ed, 21 (2010) 771-787.

[36] S.E. Sakiyama-Elbert, Incorporation of heparin into biomaterials, Acta Biomater, 10

(2014) 1581-1587.

[37] B. Soontornworajit, J. Zhou, Z. Zhang, Y. Wang, Aptamer-functionalized in situ

injectable hydrogel for controlled protein release, Biomacromolecules, 11 (2010) 2724-

2730.

[38] B. Soontornworajit, J. Zhou, M.P. Snipes, M.R. Battig, Y. Wang, Affinity hydrogels

for controlled protein release using nucleic acid aptamers and complementary

oligonucleotides, Biomaterials, 32 (2011) 6839-6849.

[39] M.R. Battig, B. Soontornworajit, Y. Wang, Programmable release of multiple

protein drugs from aptamer-functionalized hydrogels via nucleic acid hybridization, J Am

Chem Soc, 134 (2012) 12410-12413.

[40] M.R. Battig, Y. Huang, N. Chen, Y. Wang, Aptamer-functionalized superporous

hydrogels for sequestration and release of growth factors regulated via molecular

recognition, Biomaterials, 35 (2014) 8040-8048.

[41] K. Vulic, M.S. Shoichet, Tunable growth factor delivery from injectable hydrogels

for tissue engineering, J Am Chem Soc, 134 (2012) 882-885.

[42] K. Vulic, M.M. Pakulska, R. Sonthalia, A. Ramachandran, M.S. Shoichet,

Mathematical model accurately predicts protein release from an affinity-based delivery

system, J Control Release, 197 (2015) 69-77.

[43] K. Vulic, M.S. Shoichet, Affinity-based drug delivery systems for tissue repair and

regeneration, Biomacromolecules, 15 (2014) 3867-3880.

[44] M.M. Pakulska, K. Vulic, M.S. Shoichet, Affinity-based release of chondroitinase

ABC from a modified methylcellulose hydrogel, J Control Release, 171 (2013) 11-16.

[45] L. Björck, G. Kronvall, Purification and some properties of streptococcal protein G,

a novel IgG-binding reagent, The Journal of Immunology, 133 (1984) 969-974.

[46] T. Moks, L. Abrahmsen, B. Nilsson, U. Hellman, J. Sjoquist, M. Uhlen,

Staphylococcal protein A consists of five IgG-binding domains, European journal of

biochemistry, 156 (1986) 637-643.

[47] B. Guss, M. Eliasson, A. Olsson, M. Uhlén, A.K. Frej, H. Jörnvall, J.I. Flock, M.

Lindberg, Structure of the IgG-binding regions of streptococcal protein G, The EMBO

Journal, 5 (1986) 1567-1575.

[48] B. Akerstrom, L. Bjorck, A physicochemical study of protein G, a molecule with

unique immunoglobulin G-binding properties, The Journal of biological chemistry, 261

(1986) 10240-10247.

[49] K. Saha, F. Bender, E. Gizeli, Comparative Study of IgG Binding to Proteins G and

A:  Nonequilibrium Kinetic and Binding Constant Determination with the Acoustic

Waveguide Device, Analytical Chemistry, 75 (2003) 835-842.

[50] M.J. Saunders, W. Liu, C. Szent-Gyorgyi, Y. Wen, Z. Drennen, A.S. Waggoner,

W.S. Meng, Engineering Fluorogen Activating Proteins into Self-Assembling Materials,

Bioconjugate Chemistry, 24 (2013) 803-810.

Page 185: Self-Assembling Bioaffinity Hydrogels for Localized

169

[51] A. Tajima, W. Liu, I. Pradhan, S. Bertera, R.A. Lakomy, W.A. Rudert, M. Trucco,

W.S. Meng, Y. Fan, Promoting 3-D Aggregation of FACS Purified Thymic Epithelial

Cells with EAK 16-II/EAKIIH6 Self-assembling Hydrogel, (2016) e54062.

[52] A. Burd, T. Chiu, Allogenic skin in the treatment of burns, Clin Dermatol, 23 (2005)

376-387.

[53] G. Benichou, Y. Yamada, S.H. Yun, C. Lin, M. Fray, G. Tocco, Immune recognition

and rejection of allogeneic skin grafts, Immunotherapy, 3 (2011) 757-770.

[54] Y. Su, A. Richmond, Chemokine Regulation of Neutrophil Infiltration of Skin

Wounds, Adv Wound Care (New Rochelle), 4 (2015) 631-640.

[55] T. El-Sawy, M. Miura, R. Fairchild, Early T cell response to allografts occurring

prior to alloantigen priming up-regulates innate-mediated inflammation and graft

necrosis, Am J Pathol, 165 (2004) 147-157.

[56] M. Miura, T. El-Sawy, R.L. Fairchild, Neutrophils mediate parenchymal tissue

necrosis and accelerate the rejection of complete major histocompatibility complex-

disparate cardiac allografts in the absence of interferon-gamma, Am J Pathol, 162 (2003)

509-519.

[57] M. Surquin, A. Le Moine, V. Flamand, K. Rombaut, F.X. Demoor, I. Salmon, M.

Goldman, D. Abramowicz, IL-4 deficiency prevents eosinophilic rejection and uncovers

a role for neutrophils in the rejection of MHC class II disparate skin grafts,

Transplantation, 80 (2005) 1485-1492.

[58] J.V. Dovi, A.M. Szpaderska, L.A. DiPietro, Neutrophil function in the healing

wound: adding insult to injury?, Thromb Haemost, 92 (2004) 275-280.

[59] H.R. Hampton, J. Bailey, M. Tomura, R. Brink, T. Chtanova, Microbe-dependent

lymphatic migration of neutrophils modulates lymphocyte proliferation in lymph nodes,

Nat Commun, 6 (2015) 7139.

[60] H.R. Hampton, T. Chtanova, The lymph node neutrophil, Semin Immunol, 28

(2016) 129-136.

[61] A. Iwasaki, R. Medzhitov, Control of adaptive immunity by the innate immune

system, Nat Immunol, 16 (2015) 343-353.

[62] T. El-Sawy, J.A. Belperio, R.M. Strieter, D.G. Remick, R.L. Fairchild, Inhibition of

polymorphonuclear leukocyte-mediated graft damage synergizes with short-term

costimulatory blockade to prevent cardiac allograft rejection, Circulation, 112 (2005)

320-331.

[63] K. Lim, Y.M. Hyun, K. Lambert-Emo, T. Capece, S. Bae, R. Miller, D.J. Topham,

M. Kim, Neutrophil trails guide influenza-specific CD8(+) T cells in the airways,

Science, 349 (2015) aaa4352.

[64] M.F. Fransen, M. Sluijter, H. Morreau, R. Arens, C.J. Melief, Local activation of

CD8 T cells and systemic tumor eradication without toxicity via slow release and local

delivery of agonistic CD40 antibody, Clinical cancer research : an official journal of the

American Association for Cancer Research, 17 (2011) 2270-2280.

[65] M.F. Fransen, T.C. van der Sluis, F. Ossendorp, R. Arens, C.J. Melief, Controlled

local delivery of CTLA-4 blocking antibody induces CD8+ T-cell-dependent tumor

eradication and decreases risk of toxic side effects, Clinical cancer research : an official

journal of the American Association for Cancer Research, 19 (2013) 5381-5389.

[66] C. Wang, Y. Ye, Z. Gu, Local delivery of checkpoints antibodies, Human Vaccines

& Immunotherapeutics, 13 (2017) 245-248.

Page 186: Self-Assembling Bioaffinity Hydrogels for Localized

170

[67] W.F. Richter, B. Jacobsen, Subcutaneous absorption of biotherapeutics: knowns and

unknowns, Drug metabolism and disposition: the biological fate of chemicals, 42 (2014)

1881-1889.

[68] L. Kagan, Pharmacokinetic modeling of the subcutaneous absorption of therapeutic

proteins, Drug metabolism and disposition: the biological fate of chemicals, 42 (2014)

1890-1905.

[69] Y. Vugmeyster, X. Xu, F.-P. Theil, L.A. Khawli, M.W. Leach, Pharmacokinetics

and toxicology of therapeutic proteins: Advances and challenges, World Journal of

Biological Chemistry, 3 (2012) 73-92.

[70] C. Tang, F. Qiu, X. Zhao, Molecular Design and Applications of Self-Assembling

Surfactant-Like Peptides, Journal of Nanomaterials, 2013 (2013) 1-9.

[71] P. Chen, Self-assembly of ionic-complementary peptides: a physicochemical

viewpoint, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 261

(2005) 3-24.

[72] C. Ha, C.B. Park, Template-directed self-assembly and growth of insulin amyloid

fibrils, Biotechnology and Bioengineering, 90 (2005) 848-855.

[73] C. Pike, D. Burdick, A. Walencewicz, C. Glabe, C. Cotman, Neurodegeneration

induced by beta-amyloid peptides in vitro: the role of peptide assembly state, The Journal

of Neuroscience, 13 (1993) 1676-1687.

[74] J. Ghanta, C.L. Shen, L.L. Kiessling, R.M. Murphy, A strategy for designing

inhibitors of beta-amyloid toxicity, The Journal of biological chemistry, 271 (1996)

29525-29528.

[75] R.P. Nagarkar, J.P. Schneider, Synthesis and Primary Characterization of Self-

Assembled Peptide-Based Hydrogels, 474 (2008) 61-77.

[76] Y. Hong, R.L. Legge, S. Zhang, P. Chen, Effect of Amino Acid Sequence and pH on

Nanofiber Formation of Self-Assembling Peptides EAK16-II and EAK16-IV,

Biomacromolecules, 4 (2003) 1433-1442.

[77] S. Zhang, A. Rich, Direct conversion of an oligopeptide from a β-sheet to an α-helix:

A model for amyloid formation, Proceedings of the National Academy of Sciences, 94

(1997) 23-28.

[78] http://micro.magnet.fsu.edu/primer/techniques/fluorescence/fluorescenceintro.html.

29 Jun, 2018

[79] L.A. Munishkina, A.L. Fink, Fluorescence as a method to reveal structures and

membrane-interactions of amyloidogenic proteins, Biochimica et Biophysica Acta (BBA)

- Biomembranes, 1768 (2007) 1862-1885.

[80] R. Sarroukh, E. Cerf, S. Derclaye, Y. Dufrêne, E. Goormaghtigh, J.-M. Ruysschaert,

V. Raussens, Transformation of amyloid β(1–40) oligomers into fibrils is characterized

by a major change in secondary structure, Cell. Mol. Life Sci., 68 (2011) 1429-1438.

[81] M.R.H. Krebs, E.H.C. Bromley, A.M. Donald, The binding of thioflavin-T to

amyloid fibrils: localisation and implications, Journal of Structural Biology, 149 (2005)

30-37.

[82] R. Khurana, C. Coleman, C. Ionescu-Zanetti, S.A. Carter, V. Krishna, R.K. Grover,

R. Roy, S. Singh, Mechanism of thioflavin T binding to amyloid fibrils, Journal of

Structural Biology, 151 (2005) 229-238.

[83] H. LeVine III, [18] Quantification of β-sheet amyloid fibril structures with thioflavin

T, in: W. Ronald (Ed.) Methods in Enzymology, Academic Press 1999, pp. 274-284.

Page 187: Self-Assembling Bioaffinity Hydrogels for Localized

171

[84] Z. Luo, X. Zhao, S. Zhang, Self-Organization of a Chiral D-EAK16 Designer

Peptide into a 3D Nanofiber Scaffold, Macromolecular Bioscience, 8 (2008) 785-791.

[85] D.J. Müller, H. Janovjak, T. Lehto, L. Kuerschner, K. Anderson, Observing

structure, function and assembly of single proteins by AFM, Progress in Biophysics and

Molecular Biology, 79 (2002) 1-43.

[86] A. Rajendran, M. Endo, H. Sugiyama, Chapter 2 - Structural and Functional

Analysis of Proteins by High-Speed Atomic Force Microscopy, in: C. Christo, K.-C.

Tatyana (Eds.) Advances in Protein Chemistry and Structural Biology, Academic

Press2012, pp. 5-55.

[87] Z. Luo, S. Wang, S. Zhang, Fabrication of self-assembling d-form peptide nanofiber

scaffold d-EAK16 for rapid hemostasis, Biomaterials, 32 (2011) 2013-2020.

[88] M.R. Nichols, M.A. Moss, D.K. Reed, W.-L. Lin, R. Mukhopadhyay, J.H. Hoh, T.L.

Rosenberry, Growth of β-Amyloid(1−40) Protofibrils by Monomer Elongation and

Lateral Association. Characterization of Distinct Products by Light Scattering and

Atomic Force Microscopy†, Biochemistry, 41 (2002) 6115-6127.

[89] C. Goldsbury, J. Kistler, U. Aebi, T. Arvinte, G.J.S. Cooper, Watching amyloid

fibrils grow by time-lapse atomic force microscopy, Journal of Molecular Biology, 285

(1999) 33-39.

[90] D.R. Pier Carlo Braga, Atomic Force Microscopy Biomedical Methods and

Applications, Humana press2003.

[91] N.J. Greenfield, Using circular dichroism spectra to estimate protein secondary

structure, Nature Protocols, 1 (2007) 2876-2890.

[92] A.K. Nowinski, F. Sun, A.D. White, A.J. Keefe, S. Jiang, Sequence, Structure, and

Function of Peptide Self-Assembled Monolayers, Journal of the American Chemical

Society, 134 (2012) 6000-6005.

[93] R.V. Rughani, D.A. Salick, M.S. Lamm, T. Yucel, D.J. Pochan, J.P. Schneider,

Folding, Self-Assembly, and Bulk Material Properties of a De Novo Designed Three-

Stranded β-Sheet Hydrogel, Biomacromolecules, 10 (2009) 1295-1304.

[94] N. Sreerama, R.W. Woody, Estimation of Protein Secondary Structure from Circular

Dichroism Spectra: Comparison of CONTIN, SELCON, and CDSSTR Methods with an

Expanded Reference Set, Analytical Biochemistry, 287 (2000) 252-260.

[95] R.A. Dluhy, R. Mendelsohn, Emerging techniques in biophysical FT-IR, Analytical

Chemistry, 60 (1988) 269A-278A.

[96] M. Jackon, H.H. Mantsch, The Use and Misuse of FTIR Spectroscopy in the

Determination of Protein Structure, Critical Reviews in Biochemistry and Molecular

Biology, 30 (1995) 95-120.

[97] C.L. Wilder, A.D. Friedrich, R.O. Potts, G.O. Daumy, M.L. Francoeur, Secondary

structural analysis of two recombinant murine proteins, interleukins 1 alpha and 1 beta: is

infrared spectroscopy sufficient to assign structure?, Biochemistry, 31 (1992) 27-31.

[98] H.W. David Weitz, Ryan Larsen, oscillatory rheology, G.I.T. laboratory journal, 3-4

(2007) 68-70.

[99] L.L. Hyland, M.B. Taraban, Y. Feng, B. Hammouda, Y.B. Yu, Viscoelastic

properties and nanoscale structures of composite oligopeptide-polysaccharide hydrogels,

Biopolymers, 97 (2012) 177-188.

[100] O.I.d. Rı́o, A.W. Neumann, Axisymmetric Drop Shape Analysis: Computational

Methods for the Measurement of Interfacial Properties from the Shape and Dimensions of

Page 188: Self-Assembling Bioaffinity Hydrogels for Localized

172

Pendant and Sessile Drops, Journal of Colloid and Interface Science, 196 (1997) 136-

147.

[101] S.Y. Fung, C. Keyes, J. Duhamel, P. Chen, Concentration Effect on the

Aggregation of a Self-Assembling Oligopeptide, Biophysical Journal, 85 (2003) 537-548.

[102] S. Sharpe, K. Simonetti, J. Yau, P. Walsh, Solid-State NMR Characterization of

Autofluorescent Fibrils Formed by the Elastin-Derived Peptide GVGVAGVG,

Biomacromolecules, 12 (2011) 1546-1555.

[103] R. Tycko, Solid-State NMR Studies of Amyloid Fibril Structure, Annual Review of

Physical Chemistry, 62 (2011) 279-299.

[104] A.R. Cormier, C. Ruiz-Orta, R.G. Alamo, A.K. Paravastu, Solid State Self-

Assembly Mechanism of RADA16-I Designer Peptide, Biomacromolecules, 13 (2012)

1794-1804.

[105] K.J. Nagy, M.C. Giano, A. Jin, D.J. Pochan, J.P. Schneider, Enhanced Mechanical

Rigidity of Hydrogels Formed from Enantiomeric Peptide Assemblies, Journal of the

American Chemical Society, 133 (2011) 14975-14977.

[106] S. Zhang, Discovery and design of self-assembling peptides, Interface focus, 7

(2017) 20170028.

[107] S. Zhang, C. Lockshin, A. Herbert, E. Winter, A. Rich, Zuotin, a putative Z-DNA

binding protein in Saccharomyces cerevisiae, EMBO J, 11 (1992) 3787-3796.

[108] S. Jun, Y. Hong, H. Imamura, B.Y. Ha, J. Bechhoefer, P. Chen, Self-Assembly of

the Ionic Peptide EAK16: The Effect of Charge Distributions on Self-Assembly,

Biophysical Journal, 87 (2004) 1249-1259.

[109] S. Zhang, Emerging biological materials through molecular self-assembly,

Biotechnology Advances, 20 (2002) 321-339.

[110] V.P. Shastri, In vivo Engineering of Tissues: Biological Considerations,

Challenges, Strategies, and Future Directions, Advanced Materials, 21 (2009) 3246-3254.

[111] S. Zhang, C. Lockshin, R. Cook, A. Rich, Unusually stable beta-sheet formation in

an ionic self-complementary oligopeptide, Biopolymers, 34 (1994) 663-672.

[112] M.R. Caplan, E.M. Schwartzfarb, S. Zhang, R.D. Kamm, D.A. Lauffenburger,

Control of self-assembling oligopeptide matrix formation through systematic variation of

amino acid sequence, Biomaterials, 23 (2002) 219-227.

[113] T.C. Holmes, Novel peptide-based biomaterial scaffolds for tissue engineering,

Trends in Biotechnology, 20 (2002) 16-21.

[114] K. Wang, J.D. Keasling, S.J. Muller, Effects of the sequence and size of non-polar

residues on self-assembly of amphiphilic peptides, International Journal of Biological

Macromolecules, 36 (2005) 232-240.

[115] Z. Ye, H. Zhang, H. Luo, S. Wang, Q. Zhou, X. Du, C. Tang, L. Chen, J. Liu, Y.-

K. Shi, E.-Y. Zhang, R. Ellis-Behnke, X. Zhao, Temperature and pH effects on

biophysical and morphological properties of self-assembling peptide RADA16-I, Journal

of Peptide Science, 14 (2008) 152-162.

[116] M. Altman, P. Lee, A. Rich, S. Zhang, Conformational behavior of ionic self-

complementary peptides, Protein Science, 9 (2000) 1095-1105.

[117] Y. Hong, M.D. Pritzker, R.L. Legge, P. Chen, Effect of NaCl and peptide

concentration on the self-assembly of an ionic-complementary peptide EAK16-II,

Colloids and Surfaces B: Biointerfaces, 46 (2005) 152-161.

Page 189: Self-Assembling Bioaffinity Hydrogels for Localized

173

[118] M.R. Caplan, P.N. Moore, S. Zhang, R.D. Kamm, D.A. Lauffenburger, Self-

Assembly of a β-Sheet Protein Governed by Relief of Electrostatic Repulsion Relative to

van der Waals Attraction, Biomacromolecules, 1 (2000) 627-631.

[119] E.J. Leon, N. Verma, Z. Shuguang, D.A. Lauffenburger, R.D. Kamm, Mechanical

properties of a self-assembling oligopeptide matrix, Journal of Biomaterials Science,

Polymer Edition, 9 (1998) 297-312.

[120] N.A. Hammond, R.D. Kamm, Mechanical characterization of self-assembling

peptide hydrogels by microindentation, Journal of Biomedical Materials Research Part B:

Applied Biomaterials, 101B (2013) 981-990.

[121] H. Yokoi, Dynamic reassembly of peptide RADA16 nanofiber scaffold,

Proceedings of the National Academy of Sciences, 102 (2005) 8414-8419.

[122] Z. Luo, X. Zhao, S. Zhang, Structural Dynamic of a Self-Assembling Peptide d-

EAK16 Made of Only D-Amino Acids, PLoS ONE, 3 (2008) e2364.

[123] J. Kisiday, Self-assembling peptide hydrogel fosters chondrocyte extracellular

matrix production and cell division: Implications for cartilage tissue repair, Proceedings

of the National Academy of Sciences, 99 (2002) 9996-10001.

[124] C.E. Semino, J.R. Merok, G.G. Crane, G. Panagiotakos, S. Zhang, Functional

differentiation of hepatocyte-like spheroid structures from putative liver progenitor cells

in three-dimensional peptide scaffolds, Differentiation, 71 (2003) 262-270.

[125] C.E. Semino, J. Kasahara, Y. Hayashi, S. Zhang, Entrapment of Migrating

Hippocampal Neural Cells in Three-Dimensional Peptide Nanofiber Scaffold, Tissue

Engineering, 10 (2004) 643-655.

[126] D.A. Narmoneva, O. Oni, A.L. Sieminski, S. Zhang, J.P. Gertler, R.D. Kamm, R.T.

Lee, Self-assembling short oligopeptides and the promotion of angiogenesis,

Biomaterials, 26 (2005) 4837-4846.

[127] A.L. Sieminski, C.E. Semino, H. Gong, R.D. Kamm, Primary sequence of ionic

self-assembling peptide gels affects endothelial cell adhesion and capillary

morphogenesis, Journal of Biomedical Materials Research Part A, 87A (2008) 494-504.

[128] T.C. Holmes, S. de Lacalle, X. Su, G. Liu, A. Rich, S. Zhang, Extensive neurite

outgrowth and active synapse formation on self-assembling peptide scaffolds, Proc Natl

Acad Sci U S A, 97 (2000) 6728-6733.

[129] T. Wang, X. Zhong, S. Wang, F. Lv, X. Zhao, Molecular Mechanisms of

RADA16-1 Peptide on Fast Stop Bleeding in Rat Models, International Journal of

Molecular Sciences, 13 (2012) 15279-15290.

[130] Cunha, Villa, Silva, Gelain, S. Panseri, 3D culture of adult mouse neural stem cells

within functionalized self-assembling peptide scaffolds, International Journal of

Nanomedicine, (2011) 943-955.

[131] E. Genové, C. Shen, S. Zhang, C.E. Semino, The effect of functionalized self-

assembling peptide scaffolds on human aortic endothelial cell function, Biomaterials, 26

(2005) 3341-3351.

[132] Y. Chau, Y. Luo, A.C.Y. Cheung, Y. Nagai, S. Zhang, J.B. Kobler, S.M. Zeitels, R.

Langer, Incorporation of a matrix metalloproteinase-sensitive substrate into self-

assembling peptides – A model for biofunctional scaffolds, Biomaterials, 29 (2008)

1713-1719.

Page 190: Self-Assembling Bioaffinity Hydrogels for Localized

174

[133] Z. Luo, Y. Yue, Y. Zhang, X. Yuan, J. Gong, L. Wang, B. He, Z. Liu, Y. Sun, J.

Liu, M. Hu, J. Zheng, Designer D-form self-assembling peptide nanofiber scaffolds for 3-

dimensional cell cultures, Biomaterials, 34 (2013) 4902-4913.

[134] A. Nishimura, T. Hayakawa, Y. Yamamoto, M. Hamori, K. Tabata, K. Seto, N.

Shibata, Controlled release of insulin from self-assembling nanofiber hydrogel,

PuraMatrix™: Application for the subcutaneous injection in rats, European Journal of

Pharmaceutical Sciences, 45 (2012) 1-7.

[135] A. Liedmann, S. Frech, P.J. Morgan, A. Rolfs, M.J. Frech, Differentiation of

human neural progenitor cells in functionalized hydrogel matrices, BioResearch open

access, 1 (2012) 16-24.

[136] S.Y. Fung, H. Yang, P.T. Bhola, P. Sadatmousavi, E. Muzar, M. Liu, P. Chen, Self-

Assembling Peptide as a Potential Carrier for Hydrophobic Anticancer Drug Ellipticine:

Complexation, Release and In Vitro Delivery, Advanced Functional Materials, 19 (2009)

74-83.

[137] F. Gelain, L.D. Unsworth, S. Zhang, Slow and sustained release of active cytokines

from self-assembling peptide scaffolds, Journal of Controlled Release, 145 (2010) 231-

239.

[138] R. Miller, P. Kopesky, A. Grodzinsky, Growth Factor Delivery Through Self-

assembling Peptide Scaffolds, Clin Orthop Relat Res, 469 (2011) 2716-2724.

[139] S. Koutsopoulos, S. Zhang, Two-layered injectable self-assembling peptide

scaffold hydrogels for long-term sustained release of human antibodies, Journal of

Controlled Release, 160 (2012) 451-458.

[140] Y. Nagai, L.D. Unsworth, S. Koutsopoulos, S. Zhang, Slow release of molecules in

self-assembling peptide nanofiber scaffold, Journal of Controlled Release, 115 (2006) 18-

25.

[141] P. Sadatmousavi, P. Chen, Self/Co-Assembling Peptide, EAR8-II, as a Potential

Carrier for a Hydrophobic Anticancer Drug Pirarubicin (THP)-Characterization and in-

Vitro Delivery, Int J Mol Sci, 14 (2013) 23315-23329.

[142] H. Yang, M. Pritzker, S.Y. Fung, Y. Sheng, W. Wang, P. Chen, Anion Effect on

the Nanostructure of a Metal Ion Binding Self-Assembling Peptide, Langmuir, 22 (2006)

8553-8562.

[143] J.P. Schneider, D.J. Pochan, B. Ozbas, K. Rajagopal, L. Pakstis, J. Kretsinger,

Responsive Hydrogels from the Intramolecular Folding and Self-Assembly of a Designed

Peptide, Journal of the American Chemical Society, 124 (2002) 15030-15037.

[144] D.J. Pochan, J.P. Schneider, J. Kretsinger, B. Ozbas, K. Rajagopal, L. Haines,

Thermally Reversible Hydrogels via Intramolecular Folding and Consequent Self-

Assembly of a de Novo Designed Peptide, Journal of the American Chemical Society,

125 (2003) 11802-11803.

[145] B. Ozbas, J. Kretsinger, K. Rajagopal, J.P. Schneider, D.J. Pochan, Salt-Triggered

Peptide Folding and Consequent Self-Assembly into Hydrogels with Tunable Modulus,

Macromolecules, 37 (2004) 7331-7337.

[146] L.A. Haines, K. Rajagopal, B. Ozbas, D.A. Salick, D.J. Pochan, J.P. Schneider,

Light-Activated Hydrogel Formation via the Triggered Folding and Self-Assembly of a

Designed Peptide, Journal of the American Chemical Society, 127 (2005) 17025-17029.

Page 191: Self-Assembling Bioaffinity Hydrogels for Localized

175

[147] J.K. Kretsinger, L.A. Haines, B. Ozbas, D.J. Pochan, J.P. Schneider,

Cytocompatibility of self-assembled β-hairpin peptide hydrogel surfaces, Biomaterials,

26 (2005) 5177-5186.

[148] L. Haines-Butterick, K. Rajagopal, M. Branco, D. Salick, R. Rughani, M. Pilarz,

M.S. Lamm, D.J. Pochan, J.P. Schneider, Controlling hydrogelation kinetics by peptide

design for three-dimensional encapsulation and injectable delivery of cells, Proceedings

of the National Academy of Sciences, 104 (2007) 7791-7796.

[149] T. Yucel, C.M. Micklitsch, J.P. Schneider, D.J. Pochan, Direct Observation of

Early-Time Hydrogelation in β-Hairpin Peptide Self-Assembly, Macromolecules, 41

(2008) 5763-5772.

[150] T.H. Larsen, M.C. Branco, K. Rajagopal, J.P. Schneider, E.M. Furst, Sequence-

Dependent Gelation Kinetics of β-Hairpin Peptide Hydrogels, Macromolecules, 42

(2009) 8443-8450.

[151] C. Yan, A. Altunbas, T. Yucel, R.P. Nagarkar, J.P. Schneider, D.J. Pochan,

Injectable solid hydrogel: mechanism of shear-thinning and immediate recovery of

injectable [small beta]-hairpin peptide hydrogels, Soft Matter, 6 (2010) 5143-5156.

[152] C. Yan, M.E. Mackay, K. Czymmek, R.P. Nagarkar, J.P. Schneider, D.J. Pochan,

Injectable Solid Peptide Hydrogel as a Cell Carrier: Effects of Shear Flow on Hydrogels

and Cell Payload, Langmuir, 28 (2012) 6076-6087.

[153] A. Altunbas, S.J. Lee, S.A. Rajasekaran, J.P. Schneider, D.J. Pochan,

Encapsulation of curcumin in self-assembling peptide hydrogels as injectable drug

delivery vehicles, Biomaterials, 32 (2011) 5906-5914.

[154] M.C. Branco, D.J. Pochan, N.J. Wagner, J.P. Schneider, The effect of protein

structure on their controlled release from an injectable peptide hydrogel, Biomaterials, 31

(2010) 9527-9534.

[155] M.C. Branco, D.J. Pochan, N.J. Wagner, J.P. Schneider, Macromolecular diffusion

and release from self-assembled β-hairpin peptide hydrogels, Biomaterials, 30 (2009)

1339-1347.

[156] C. Sinthuvanich, L.A. Haines-Butterick, K.J. Nagy, J.P. Schneider, Iterative design

of peptide-based hydrogels and the effect of network electrostatics on primary

chondrocyte behavior, Biomaterials, 33 (2012) 7478-7488.

[157] J.D. Hartgerink, E. Beniash, S.I. Stupp, Peptide-amphiphile nanofibers: A versatile

scaffold for the preparation of self-assembling materials, Proceedings of the National

Academy of Sciences, 99 (2002) 5133-5138.

[158] H. Cui, T. Muraoka, A.G. Cheetham, S.I. Stupp, Self-Assembly of Giant Peptide

Nanobelts, Nano Letters, 9 (2009) 945-951.

[159] J.D. Hartgerink, Self-Assembly and Mineralization of Peptide-Amphiphile

Nanofibers, Science, 294 (2001) 1684-1688.

[160] H.A. Behanna, J.J.J.M. Donners, A.C. Gordon, S.I. Stupp, Coassembly of

Amphiphiles with Opposite Peptide Polarities into Nanofibers, Journal of the American

Chemical Society, 127 (2005) 1193-1200.

[161] X.R. Zhou, R. Ge, S.Z. Luo, Self-assembly of pH and calcium dual-responsive

peptide-amphiphilic hydrogel, Journal of Peptide Science, 19 (2013) 737-744.

[162] S.E. Paramonov, H.-W. Jun, J.D. Hartgerink, Self-Assembly of

Peptide−Amphiphile Nanofibers:  The Roles of Hydrogen Bonding and Amphiphilic

Packing, Journal of the American Chemical Society, 128 (2006) 7291-7298.

Page 192: Self-Assembling Bioaffinity Hydrogels for Localized

176

[163] Y.S. Velichko, S.I. Stupp, M.O. de la Cruz, Molecular Simulation Study of Peptide

Amphiphile Self-Assembly, The Journal of Physical Chemistry B, 112 (2008) 2326-

2334.

[164] T. Yu, G.C. Schatz, Free-Energy Landscape for Peptide Amphiphile Self-

Assembly: Stepwise versus Continuous Assembly Mechanisms, The Journal of Physical

Chemistry B, 117 (2013) 14059-14064.

[165] E. Beniash, J.D. Hartgerink, H. Storrie, J.C. Stendahl, S.I. Stupp, Self-assembling

peptide amphiphile nanofiber matrices for cell entrapment, Acta Biomaterialia, 1 (2005)

387-397.

[166] J.D. Hartgerink, Supramolecular Chemistry And Self-assembly Special Feature:

Peptide-amphiphile nanofibers: A versatile scaffold for the preparation of self-assembling

materials, Proceedings of the National Academy of Sciences, 99 (2002) 5133-5138.

[167] E.T. Pashuck, H. Cui, S.I. Stupp, Tuning Supramolecular Rigidity of Peptide Fibers

through Molecular Structure, Journal of the American Chemical Society, 132 (2010)

6041-6046.

[168] K.L. Niece, J.D. Hartgerink, J.J.J.M. Donners, S.I. Stupp, Self-Assembly

Combining Two Bioactive Peptide-Amphiphile Molecules into Nanofibers by

Electrostatic Attraction, Journal of the American Chemical Society, 125 (2003) 7146-

7147.

[169] T.P. Kraehenbuehl, P. Zammaretti, A.J. Van der Vlies, R.G. Schoenmakers, M.P.

Lutolf, M.E. Jaconi, J.A. Hubbell, Three-dimensional extracellular matrix-directed

cardioprogenitor differentiation: Systematic modulation of a synthetic cell-responsive

PEG-hydrogel, Biomaterials, 29 (2008) 2757-2766.

[170] A.J. Engler, S. Sen, H.L. Sweeney, D.E. Discher, Matrix Elasticity Directs Stem

Cell Lineage Specification, Cell, 126 (2006) 677-689.

[171] F. Rehfeldt, A.J. Engler, A. Eckhardt, F. Ahmed, D.E. Discher, Cell responses to

the mechanochemical microenvironment—Implications for regenerative medicine and

drug delivery, Advanced Drug Delivery Reviews, 59 (2007) 1329-1339.

[172] M.D. Pierschbacher, E. Ruoslahti, Cell attachment activity of fibronectin can be

duplicated by small synthetic fragments of the molecule, Nature, 309 (1984) 30-33.

[173] K. Tashiro, G.C. Sephel, B. Weeks, M. Sasaki, G.R. Martin, H.K. Kleinman, Y.

Yamada, A synthetic peptide containing the IKVAV sequence from the A chain of

laminin mediates cell attachment, migration, and neurite outgrowth, Journal of Biological

Chemistry, 264 (1989) 16174-16182.

[174] Y. Yamada, H.K. Kleinman, Functional domains of cell adhesion molecules,

Current Opinion in Cell Biology, 4 (1992) 819-823.

[175] G.A. Silva, C. Czeisler, K.L. Niece, E. Beniash, D.A. Harrington, J.A. Kessler, S.I.

Stupp, Selective Differentiation of Neural Progenitor Cells byHigh-Epitope Density

Nanofibers, Science, 303 (2004) 1352-1355.

[176] S. Ghanaati, M.J. Webber, R.E. Unger, C. Orth, J.F. Hulvat, S.E. Kiehna, M.

Barbeck, A. Rasic, S.I. Stupp, C.J. Kirkpatrick, Dynamic in vivo biocompatibility of

angiogenic peptide amphiphile nanofibers, Biomaterials, 30 (2009) 6202-6212.

[177] L.W. Chow, L.-j. Wang, D.B. Kaufman, S.I. Stupp, Self-assembling nanostructures

to deliver angiogenic factors to pancreatic islets, Biomaterials, 31 (2010) 6154-6161.

[178] M.J. Webber, X. Han, S.N. Prasanna Murthy, K. Rajangam, S.I. Stupp, J.W.

Lomasney, Capturing the stem cell paracrine effect using heparin-presenting nanofibres

Page 193: Self-Assembling Bioaffinity Hydrogels for Localized

177

to treat cardiovascular diseases, Journal of Tissue Engineering and Regenerative

Medicine, 4 (2010) 600-610.

[179] S.M. Standley, D.J. Toft, H. Cheng, S. Soukasene, J. Chen, S.M. Raja, V. Band, H.

Band, V.L. Cryns, S.I. Stupp, Induction of Cancer Cell Death by Self-assembling

Nanostructures Incorporating a Cytotoxic Peptide, Cancer Research, 70 (2010) 3020-

3026.

[180] R.H. Zha, S. Sur, S.I. Stupp, Self-assembly of Cytotoxic Peptide Amphiphiles into

Supramolecular Membranes for Cancer Therapy, Advanced Healthcare Materials, 2

(2013) 126-133.

[181] M.J. Webber, J. Tongers, M.-A. Renault, J.G. Roncalli, D.W. Losordo, S.I. Stupp,

Development of bioactive peptide amphiphiles for therapeutic cell delivery, Acta

Biomaterialia, 6 (2010) 3-11.

[182] M.O. Guler, R.C. Claussen, S.I. Stupp, Encapsulation of pyrene within self-

assembled peptide amphiphile nanofibers, Journal of Materials Chemistry, 15 (2005)

4507-4512.

[183] J.B. Matson, C.J. Newcomb, R. Bitton, S.I. Stupp, Nanostructure-templated control

of drug release from peptide amphiphile nanofiber gels, Soft Matter, 8 (2012) 3586-3595.

[184] S.R. Bull, M.O. Guler, R.E. Bras, T.J. Meade, S.I. Stupp, Self-Assembled Peptide

Amphiphile Nanofibers Conjugated to MRI Contrast Agents, Nano Letters, 5 (2004) 1-4.

[185] S.R. Bull, M.O. Guler, R.E. Bras, P.N. Venkatasubramanian, S.I. Stupp, T.J.

Meade, Magnetic Resonance Imaging of Self-Assembled Biomaterial Scaffolds,

Bioconjugate Chemistry, 16 (2005) 1343-1348.

[186] N. Stephanopoulos, J.H. Ortony, S.I. Stupp, Self-assembly for the synthesis of

functional biomaterials, Acta Materialia, 61 (2013) 912-930.

[187] M. Kabiri, I. Bushnak, M.T. McDermot, L.D. Unsworth, Toward a Mechanistic

Understanding of Ionic Self-Complementary Peptide Self-Assembly: Role of Water

Molecules and Ions, Biomacromolecules, 14 (2013) 3943-3950.

[188] K. Morris, L. Serpell, From natural to designer self-assembling biopolymers, the

structural characterisation of fibrous proteins & peptides using fibre diffraction, Chem

Soc Rev, 39 (2010) 3445-3453.

[189] A.R. Cormier, X. Pang, M.I. Zimmerman, H.-X. Zhou, A.K. Paravastu, Molecular

Structure of RADA16-I Designer Self-Assembling Peptide Nanofibers, ACS Nano,

(2013).

[190] E. Vacchelli, J. Pol, N. Bloy, A. Eggermont, I. Cremer, W.H. Fridman, J. Galon, A.

Marabelle, H. Kohrt, L. Zitvogel, G. Kroemer, L. Galluzzi, Trial watch: Tumor-targeting

monoclonal antibodies for oncological indications, Oncoimmunology, 4 (2015) e985940.

[191] N.A. Hotaling, L. Tang, D.J. Irvine, J.E. Babensee, Biomaterial Strategies for

Immunomodulation, Annu Rev Biomed Eng, (2015).

[192] N. Bhattarai, J. Gunn, M. Zhang, Chitosan-based hydrogels for controlled,

localized drug delivery, Advanced Drug Delivery Reviews, 62 (2010) 83-99.

[193] G.M. Cruise, D.S. Scharp, J.A. Hubbell, Characterization of permeability and

network structure of interfacially photopolymerized poly(ethylene glycol) diacrylate

hydrogels, Biomaterials, 19 (1998) 1287-1294.

[194] B. Jeong, Y.H. Bae, S.W. Kim, Drug release from biodegradable injectable

thermosensitive hydrogel of PEG-PLGA-PEG triblock copolymers, Journal of controlled

release : official journal of the Controlled Release Society, 63 (2000) 155-163.

Page 194: Self-Assembling Bioaffinity Hydrogels for Localized

178

[195] S. Zhang, T. Holmes, C. Lockshin, A. Rich, Spontaneous assembly of a self-

complementary oligopeptide to form a stable macroscopic membrane, Proc Natl Acad Sci

U S A, 90 (1993) 3334-3338.

[196] S. Zhang, Fabrication of novel biomaterials through molecular self-assembly, Nat

Biotechnol, 21 (2003) 1171-1178.

[197] C. Keyes-Baig, J. Duhamel, S.Y. Fung, J. Bezaire, P. Chen, Self-assembling

peptide as a potential carrier of hydrophobic compounds, J Am Chem Soc, 126 (2004)

7522-7532.

[198] Y. Wen, S.L. Roudebush, G.A. Buckholtz, T.R. Goehring, N. Giannoukakis, E.S.

Gawalt, W.S. Meng, Coassembly of amphiphilic peptide EAK16-II with histidinylated

analogues and implications for functionalization of β-sheet fibrils< i> in vivo</i>,

Biomaterials, 35 (2014) 5196-5205.

[199] S. Emamyari, H. Fazli, pH-dependent self-assembly of EAK16 peptides in the

presence of a hydrophobic surface: Coarse-grained molecular dynamics simulation, Soft

matter, 10 (2014) 4248-4257.

[200] M.J. Saunders, W. Liu, C. Szent-Gyorgyi, Y. Wen, Z. Drennen, A.S. Waggoner,

W.S. Meng, Engineering fluorogen activating proteins into self-assembling materials,

Bioconjug Chem, 24 (2013) 803-810.

[201] Y. Wen, H.R. Kolonich, K.M. Kruszewski, N. Giannoukakis, E.S. Gawalt, W.S.

Meng, Retaining antibodies in tumors with a self-assembling injectable system, Mol

Pharm, 10 (2013) 1035-1044.

[202] Y. Wen, W. Liu, C. Bagia, S. Zhang, M. Bai, J.M. Janjic, N. Giannoukakis, E.S.

Gawalt, W.S. Meng, Antibody-functionalized peptidic membranes for neutralization of

allogeneic skin antigen-presenting cells, Acta Biomater, 10 (2014) 4759-4767.

[203] Y. Zheng, Y. Wen, A.M. George, A.M. Steinbach, B.E. Phillips, N. Giannoukakis,

E.S. Gawalt, W.S. Meng, A peptide-based material platform for displaying antibodies to

engage T cells, Biomaterials, 32 (2011) 249-257.

[204] K. Saha, F. Bender, E. Gizeli, Comparative study of IgG binding to proteins G and

A: nonequilibrium kinetic and binding constant determination with the acoustic

waveguide device, Anal Chem, 75 (2003) 835-842.

[205] M. Eliasson, A. Olsson, E. Palmcrantz, K. Wiberg, M. Inganas, B. Guss, M.

Lindberg, M. Uhlen, Chimeric IgG-binding receptors engineered from staphylococcal

protein A and streptococcal protein G, The Journal of biological chemistry, 263 (1988)

4323-4327.

[206] C. Szent-Gyorgyi, B.F. Schmidt, Y. Creeger, G.W. Fisher, K.L. Zakel, S. Adler,

J.A. Fitzpatrick, C.A. Woolford, Q. Yan, K.V. Vasilev, P.B. Berget, M.P. Bruchez, J.W.

Jarvik, A. Waggoner, Fluorogen-activating single-chain antibodies for imaging cell

surface proteins, Nat Biotechnol, 26 (2008) 235-240.

[207] C. Szent-Gyorgyi, R.L. Stanfield, S. Andreko, A. Dempsey, M. Ahmed, S. Capek,

A.S. Waggoner, I.A. Wilson, M.P. Bruchez, Malachite green mediates homodimerization

of antibody VL domains to form a fluorescent ternary complex with singular symmetric

interfaces, J Mol Biol, 425 (2013) 4595-4613.

[208] A.S. Fu, T.R. Thatiparti, G.M. Saidel, H.A. von Recum, Experimental studies and

modeling of drug release from a tunable affinity-based drug delivery platform, Annals of

biomedical engineering, 39 (2011) 2466-2475.

Page 195: Self-Assembling Bioaffinity Hydrogels for Localized

179

[209] J.D. Hoffman, S. Frankel, Numerical methods for engineers and scientists, CRC

press2001.

[210] A.V. Hill, The possible effects of the aggregation of the molecules of haemoglobin

on its dissociation curves, J Physiol (Lond), 40 (1910) 4-7.

[211] T.L. Grove, M.B. Baker, Phase equilibrium controls on the tholeiitic versus calc‐alkaline differentiation trends, Journal of Geophysical Research: Solid Earth (1978–

2012), 89 (1984) 3253-3274.

[212] M.B. Baker, T.L. Grove, Kinetic controls on pyroxene nucleation and metastable

liquid lines of descent in a basaltic andesite, American Mineralogist, 70 (1985) 279-287.

[213] C. Engman, Y. Wen, W.S. Meng, R. Bottino, M. Trucco, N. Giannoukakis,

Generation of antigen-specific Foxp3+ regulatory T-cells in vivo following

administration of diabetes-reversing tolerogenic microspheres does not require provision

of antigen in the formulation, Clin Immunol, (2015).

[214] I.E. Zarraga, R. Taing, J. Zarzar, J. Luoma, J. Hsiung, A. Patel, F.J. Lim, High

shear rheology and anisotropy in concentrated solutions of monoclonal antibodies,

Journal of pharmaceutical sciences, 102 (2013) 2538-2549.

[215] S. Jun, Y. Hong, H. Imamura, B.Y. Ha, J. Bechhoefer, P. Chen, Self-assembly of

the ionic peptide EAK16: the effect of charge distributions on self-assembly, Biophys J,

87 (2004) 1249-1259.

[216] E.E. Friedrich, A. Azofiefa, E. Fisch, N.R. Washburn, Local Delivery of Antitumor

Necrosis Factor-α Through Conjugation to Hyaluronic Acid: Dosing Strategies and Early

Healing Effects in a Rat Burn Model, Journal of Burn Care & Research, 36 (2015) e90-

e101.

[217] E.E. Friedrich, A. Azofiefa, E. Fisch, N.R. Washburn, Local delivery of antitumor

necrosis factor-alpha through conjugation to hyaluronic acid: dosing strategies and early

healing effects in a rat burn model, Journal of burn care & research : official publication

of the American Burn Association, 36 (2015) e90-e101.

[218] H. Xu, K.T. Uysal, J.D. Becherer, P. Arner, G.S. Hotamisligil, Altered Tumor

Necrosis Factor-α (TNF-α) Processing in Adipocytes and Increased Expression of

Transmembrane TNF-α in Obesity, Diabetes, 51 (2002) 1876-1883.

[219] Z.N. Mahmoud, D.J. Grundy, K.J. Channon, D.N. Woolfson, The non-covalent

decoration of self-assembling protein fibers, Biomaterials, 31 (2010) 7468-7474.

[220] J.H. Collier, J.S. Rudra, J.Z. Gasiorowski, J.P. Jung, Multi-component extracellular

matrices based on peptide self-assembly, Chem Soc Rev, 39 (2010) 3413-3424.

[221] J.P. Jung, J.V. Moyano, J.H. Collier, Multifactorial optimization of endothelial cell

growth using modular synthetic extracellular matrices, Integr Biol (Camb), 3 (2011) 185-

196.

[222] F. Gelain, D. Silva, A. Caprini, F. Taraballi, A. Natalello, O. Villa, K.T. Nam, R.N.

Zuckermann, S.M. Doglia, A. Vescovi, BMHP1-derived self-assembling peptides:

hierarchically assembled structures with self-healing propensity and potential for tissue

engineering applications, ACS nano, 5 (2011) 1845-1859.

[223] A. Tajima, W. Liu, I. Pradhan, S. Bertera, C. Bagia, M. Trucco, W.S. Meng, Y.

Fan, Bioengineering mini functional thymic units with EAK16-II/EAKIIH6 self-

assembling hydrogel, Clin Immunol, (2015).

[224] N. Annabi, A. Tamayol, J.A. Uquillas, M. Akbari, L.E. Bertassoni, C. Cha, G.

Camci-Unal, M.R. Dokmeci, N.A. Peppas, A. Khademhosseini, 25th anniversary article:

Page 196: Self-Assembling Bioaffinity Hydrogels for Localized

180

Rational design and applications of hydrogels in regenerative medicine, Adv Mater, 26

(2014) 85-123.

[225] D.E. Leckband, The influence of protein and interfacial structure on the self-

assembly of oriented protein arrays, Adv Biophys, 34 (1997) 173-190.

[226] R.A. Vijayendran, D.E. Leckband, A quantitative assessment of heterogeneity for

surface-immobilized proteins, Anal Chem, 73 (2001) 471-480.

[227] C.P. Johnson, I.E. Jensen, A. Prakasam, R. Vijayendran, D. Leckband, Engineered

protein a for the orientational control of immobilized proteins, Bioconjug Chem, 14

(2003) 974-978.

[228] J.D. Bryers, C.M. Giachelli, B.D. Ratner, Engineering biomaterials to integrate and

heal: the biocompatibility paradigm shifts, Biotechnology and bioengineering, 109 (2012)

1898-1911.

[229] C. Salvador-Morales, L. Zhang, R. Langer, O.C. Farokhzad, Immunocompatibility

properties of lipid-polymer hybrid nanoparticles with heterogeneous surface functional

groups, Biomaterials, 30 (2009) 2231-2240.

[230] M. Zhang, S.K. Chakraborty, P. Sampath, J.J. Rojas, W. Hou, S. Saurabh, S.H.

Thorne, M.P. Bruchez, A.S. Waggoner, Fluoromodule-based reporter/probes designed for

in vivo fluorescence imaging, The Journal of clinical investigation, 125 (2015) 3915-

3927.

[231] H. Ozhalici-Unal, C.L. Pow, S.A. Marks, L.D. Jesper, G.L. Silva, N.I. Shank, E.W.

Jones, J.M. Burnette, 3rd, P.B. Berget, B.A. Armitage, A rainbow of fluoromodules: a

promiscuous scFv protein binds to and activates a diverse set of fluorogenic cyanine

dyes, J Am Chem Soc, 130 (2008) 12620-12621.

[232] C. Szent-Gyorgyi, B.F. Schmidt, J.A. Fitzpatrick, M.P. Bruchez, Fluorogenic

dendrons with multiple donor chromophores as bright genetically targeted and activated

probes, J Am Chem Soc, 132 (2010) 11103-11109.

[233] W. Liu, M.J. Saunders, C. Bagia, E.C. Freeman, Y. Fan, E.S. Gawalt, A.S.

Waggoner, W.S. Meng, Local retention of antibodies in vivo with an injectable film

embedded with a fluorogen-activating protein, Journal of Controlled Release, 230 (2016)

1-12.

[234] C. Ballow, A. Leh, K. Slentz-Kesler, J. Yan, D. Haughey, E. Bernton, Safety,

pharmacokinetic, immunogenicity, and pharmacodynamic responses in healthy

volunteers following a single intravenous injection of purified staphylococcal protein A,

Journal of clinical pharmacology, 53 (2013) 909-918.

[235] B. Akerström, T. Brodin, K. Reis, L. Björck, Protein G: a powerful tool for binding

and detection of monoclonal and polyclonal antibodies, The Journal of Immunology, 135

(1985) 2589-2592.

[236] B. Akerström, L. Björck, A physicochemical study of protein G, a molecule with

unique immunoglobulin G-binding properties, Journal of Biological Chemistry, 261

(1986) 10240-10247.

[237] B. Akerstrom, E. Nielsen, L. Bjorck, Definition of IgG- and albumin-binding

regions of streptococcal protein G, The Journal of biological chemistry, 262 (1987)

13388-13391.

[238] U. Sjöbring, C. Falkenberg, E. Nielsen, B. Akerström, L. Björck, Isolation and

characterization of a 14-kDa albumin-binding fragment of streptococcal protein G, The

Journal of Immunology, 140 (1988) 1595-1599.

Page 197: Self-Assembling Bioaffinity Hydrogels for Localized

181

[239] M.D. Swartzlander, C.A. Barnes, A.K. Blakney, J.L. Kaar, T.R. Kyriakides, S.J.

Bryant, Linking the foreign body response and protein adsorption to PEG-based

hydrogels using proteomics, Biomaterials, 41 (2015) 26-36.

[240] Q. Tang, K. Lee, Regulatory T-cell therapy for transplantation: how many cells do

we need?, Current opinion in organ transplantation, 17 (2012) 349-354.

[241] F. Issa, J. Hester, K. Milward, K.J. Wood, Homing of regulatory T cells to human

skin is important for the prevention of alloimmune-mediated pathology in an in vivo

cellular therapy model, PLoS One, 7 (2012) e53331.

[242] H. Yagi, T. Nomura, K. Nakamura, S. Yamazaki, T. Kitawaki, S. Hori, M. Maeda,

M. Onodera, T. Uchiyama, S. Fujii, S. Sakaguchi, Crucial role of FOXP3 in the

development and function of human CD25+CD4+ regulatory T cells, International

immunology, 16 (2004) 1643-1656.

[243] M. Ezzelarab, A.W. Thomson, Tolerogenic dendritic cells and their role in

transplantation, Seminars in immunology, 23 (2011) 252-263.

[244] J. Marino, J. Paster, G. Benichou, Allorecognition by T Lymphocytes and Allograft

Rejection, Frontiers in immunology, 7 (2016) 582.

[245] O.S. Qureshi, Y. Zheng, K. Nakamura, K. Attridge, C. Manzotti, E.M. Schmidt, J.

Baker, L.E. Jeffery, S. Kaur, Z. Briggs, T.Z. Hou, C.E. Futter, G. Anderson, L.S. Walker,

D.M. Sansom, Trans-endocytosis of CD80 and CD86: a molecular basis for the cell-

extrinsic function of CTLA-4, Science, 332 (2011) 600-603.

[246] A.S. Dejean, D.R. Beisner, I.L. Ch'en, Y.M. Kerdiles, A. Babour, K.C. Arden, D.H.

Castrillon, R.A. DePinho, S.M. Hedrick, Transcription factor Foxo3 controls the

magnitude of T cell immune responses by modulating the function of dendritic cells,

Nature immunology, 10 (2009) 504-513.

[247] E.A. Tivol, F. Borriello, A.N. Schweitzer, W.P. Lynch, J.A. Bluestone, A.H.

Sharpe, Loss of CTLA-4 leads to massive lymphoproliferation and fatal multiorgan tissue

destruction, revealing a critical negative regulatory role of CTLA-4, Immunity, 3 (1995)

541-547.

[248] D.K. Sojka, A. Hughson, D.J. Fowell, CTLA-4 is required by CD4+CD25+ Treg to

control CD4+ T-cell lymphopenia-induced proliferation, European journal of

immunology, 39 (2009) 1544-1551.

[249] W. Zhang, D. Zhang, M. Shen, Y. Liu, Y. Tian, A.W. Thomson, W.P. Lee, X.X.

Zheng, Combined administration of a mutant TGF-beta1/Fc and rapamycin promotes

induction of regulatory T cells and islet allograft tolerance, Journal of immunology

(Baltimore, Md. : 1950), 185 (2010) 4750-4759.

[250] X.X. Zheng, A.W. Steele, W.W. Hancock, A.C. Stevens, P.W. Nickerson, P. Roy-

Chaudhury, Y. Tian, T.B. Strom, A noncytolytic IL-10/Fc fusion protein prevents

diabetes, blocks autoimmunity, and promotes suppressor phenomena in NOD mice,

Journal of immunology (Baltimore, Md. : 1950), 158 (1997) 4507-4513.

[251] P.O. Flores-Villanueva, X.X. Zheng, T.B. Strom, M.J. Stadecker, Recombinant IL-

10 and IL-10/Fc treatment down-regulate egg antigen-specific delayed hypersensitivity

reactions and egg granuloma formation in schistosomiasis, Journal of immunology

(Baltimore, Md. : 1950), 156 (1996) 3315-3320.

[252] X.X. Zheng, A.W. Steele, P.W. Nickerson, W. Steurer, J. Steiger, T.B. Strom,

Administration of noncytolytic IL-10/Fc in murine models of lipopolysaccharide-induced

Page 198: Self-Assembling Bioaffinity Hydrogels for Localized

182

septic shock and allogeneic islet transplantation, Journal of immunology (Baltimore, Md.

: 1950), 154 (1995) 5590-5600.

[253] X. Feng, X.X. Zheng, S. Yi, A.M. Lehnert, T.B. Strom, P.J. O'Connell, IL-10/Fc

inhibits macrophage function and prolongs pancreatic islet xenograft survival,

Transplantation, 68 (1999) 1775-1783.

[254] W. Steurer, P.W. Nickerson, A.W. Steele, J. Steiger, X.X. Zheng, T.B. Strom, Ex

vivo coating of islet cell allografts with murine CTLA4/Fc promotes graft tolerance,

Journal of immunology (Baltimore, Md. : 1950), 155 (1995) 1165-1174.

[255] L.C. Kam, K. Shen, M.L. Dustin, Micro- and nanoscale engineering of cell

signaling, Annual review of biomedical engineering, 15 (2013) 305-326.

[256] C.C. Lin, A.T. Metters, Metal-chelating affinity hydrogels for sustained protein

release, Journal of biomedical materials research. Part A, 83 (2007) 954-964.

[257] V. Pucino, M. Bombardieri, C. Pitzalis, C. Mauro, Lactate at the crossroads of

metabolism, inflammation, and autoimmunity, European journal of immunology, 47

(2017) 14-21.

[258] L. Kagan, D.E. Mager, Mechanisms of subcutaneous absorption of rituximab in

rats, Drug metabolism and disposition: the biological fate of chemicals, 41 (2013) 248-

255.

[259] S.F. Banani, H.O. Lee, A.A. Hyman, M.K. Rosen, Biomolecular condensates:

organizers of cellular biochemistry, Nature reviews. Molecular cell biology, 18 (2017)

285-298.

[260] Y. Loo, S. Zhang, C.A.E. Hauser, From short peptides to nanofibers to

macromolecular assemblies in biomedicine, Biotechnology Advances, 30 (2012) 593-

603.

[261] H. Gouda, M. Shiraishi, H. Takahashi, K. Kato, H. Torigoe, Y. Arata, I. Shimada,

NMR study of the interaction between the B domain of staphylococcal protein A and the

Fc portion of immunoglobulin G, Biochemistry, 37 (1998) 129-136.

[262] R. Edmondson, J.J. Broglie, A.F. Adcock, L. Yang, Three-Dimensional Cell

Culture Systems and Their Applications in Drug Discovery and Cell-Based Biosensors,

Assay and Drug Development Technologies, 12 (2014) 207-218.

[263] K.L. Anderson, N.C. Moore, D.E. McLoughlin, E.J. Jenkinson, J.J. Owen, Studies

on thymic epithelial cells in vitro, Developmental and comparative immunology, 22

(1998) 367-377.

[264] A. Tajima, W. Liu, I. Pradhan, S. Bertera, C. Bagia, M. Trucco, W.S. Meng, Y.

Fan, Bioengineering mini functional thymic units with EAK16-II/EAKIIH6 self-

assembling hydrogel, Clinical immunology (Orlando, Fla.), 160 (2015) 82-89.