<<

PLATELET-INSPIRED FOR THE HEMOSTATIC

MANAGEMENT OF COMPLICATIONS IN

THROMBOCYTOPENIA AND TRAUMA

by

DASHAWN A. HICKMAN

Submitted in partial fulfillment of the requirement for the degree of

Doctor of Philosophy

Department of

CASE WESTERN RESERVE UNIVERSITY

January, 2019

CASE WESTERN RESERVE UNIVERSITY

SCHOOL OF GRADUATE STUDIES

We hereby approve the thesis/dissertation of

DaShawn A. Hickman

candidate for the degree of Doctor of Philosophy*.

Committee Chair Nicholas Ziats, PhD

Committee Member Anirban Sen Gupta, PhD

Committee Member James Anderson, MD, PhD

Committee Member Agata Exner, PhD

Committee Member Howard Meyerson, MD

Committee Member Clive Hamlin, PhD

Date of Defense August 16, 2018

*We also certify that written approval has been obtained for any proprietary material contained therein.

Dedication

I would like to dedicate my dissertation to my family who have always supported me and encouraged me to be great, especially my parents and siblings; my friends for setting the bar high and keeping my spirits higher, the MSTP entering class of 201l and my labmates.

i Table of Contents Dedication ...... i

List of Tables ...... ix

List of Figures ...... x

Acknowledgements ...... xvi

List of Abbreviations ...... xvii

Abstract ...... xxvii

Chapter 1 : Bleeding Complications in and Trauma ...... 1

1.1 Introduction ...... 1

1.2 Trauma ...... 2

1.3 Thrombocytopenia...... 4

1.4 Complex Mechanism of Hemostasis ...... 7

1.5 Management of Traumatic Bleeding ...... 10

1.6 Acknowledgments ...... 12

1.7 References ...... 12

Chapter 2 : Externally Administered (Topical and Intracavitary) Biomaterials and

Advanced Technologies for Management of Traumatic Bleeding ...... 28

2.1 Introduction ...... 28

2.2 Tourniquets ...... 28

2.3 Naturally derived biomaterials for hemostatic applications ...... 30

ii 2.3.1 Absorptive and passively interactive materials in hemostatic technologies ...... 31

2.3.2 Bioactive materials in hemostatic technologies ...... 35

2.4 Synthetically derived hemostatic materials ...... 52

2.5 Compression bandage technologies ...... 57

2.6 Combination systems and advanced technologies ...... 61

2.7 Hemostatic materials and technologies for intracavitary applications...... 62

2.8 Discussion ...... 63

2.9 Acknowledgments ...... 66

2.10 References ...... 66

Chapter 3 : Intravenously Administered Biomaterials and Advanced Technologies for

Management of Traumatic Non-compressible Hemorrhage ...... 110

3.1 Introduction ...... 110

3.2 Naturally derived intravascular pro-coagulant materials ...... 111

3.3 Synthetically derived pro-coagulant systems ...... 114

3.4 Materials and technologies to enhance clot strength and stability ...... 116

3.5 Materials and technologies for vascular embolization ...... 122

3.6 Natural platelet-derived systems ...... 124

3.7 Synthetic biomaterials-based platelet mimics and substitutes ...... 134

3.8 Discussion ...... 139

3.9 Acknowledgments ...... 143

iii 3.10 References ...... 143

Chapter 4 : Designing an Ideal Platelet-inspired System for the Intravenous

Management of Traumatic Bleeding...... 166

4.1 Introduction ...... 166

4.2 Current State of Traumatic Bleeding Management ...... 167

4.3 Designing an Ideal System ...... 169

4.4 Design of SynthoPlateTM System ...... 171

4.5 SynthoPlateTM In Vitro Evaluation ...... 173

4.6 In Vivo Evaluation of SynthoPlateTM In Thrombocytopenia Model ...... 174

4.7 Discussion ...... 178

4.8 Acknowledgments ...... 180

4.9 References ...... 181

Chapter 5 : Intravenous administration of synthetic platelets (SynthoPlate) in a mouse liver model of uncontrolled hemorrhage improves hemostasis ...... 189

5.1 Introduction ...... 189

5.2 Materials and Methods ...... 190

5.2.1 Animals ...... 190

5.2.2 SynthoPlate Manufacture ...... 190

5.2.3 Liver Laceration Model of Uncontrolled Hemorrhage and SynthoPlate Evaluation ...... 191

5.2.4 In Vivo Biodistribution of SynthoPlate Particles ...... 193

5.2.5 Statistics...... 194

iv 5.3 Results ...... 194

5.3.1 SynthoPlate Transfusion as a Pretreatment Results in Decreased Loss and Improved

Hemodynamics During Acute, Severe Hemorrhagic Shock ...... 194

5.3.2 Systemic Biodistribution SynthoPlate Particles Following Liver Laceration ...... 196

5.3.3 SynthoPlate Particles Reduce Blood Loss and Improve Hemodynamics in Mice in

Hemorrhagic Shock ...... 198

5.4 Discussion ...... 198

5.5 Authorship...... 203

5.6 Acknowledgement ...... 203

5.7 Disclosure ...... 203

5.8 References ...... 204

Chapter 6 : Intravenous synthetic platelet (SynthoPlateTM) nanoconstructs reduce bleeding and improve ‘golden hour’ survival in a porcine model of traumatic arterial hemorrhage ...... 207

6.1 Introduction ...... 207

6.2 Results ...... 212

6.2.1 SynthoPlateTM stability in storage ...... 212

6.2.2 SynthoPlateTM sterilization and its effects ...... 213

6.2.3 SynthoPlateTM cytotoxicity analysis ...... 217

6.2.4 In vivo Safety and Biodistribution Studies in Pigs ...... 217

6.2.5 Hemostatic Efficacy Evaluation in Femoral Artery Bleeding Model ...... 221

6.3 Discussion ...... 226

6.4 Materials and Methods ...... 230

v 6.4.1 Materials ...... 230

6.4.2 Manufacture of SynthoPlateTM ...... 232

6.4.3 Long term Storage-in-Suspension Studies ...... 233

6.4.4 Studies on SynthoPlateTM sterilization and its biofunctional effects ...... 233

6.4.5 In vitro cytotoxicity assay with SynthoPlateTM ...... 235

6.4.6 In vivo Safety and Biodistribution Studies in Pigs ...... 236

6.4.7 Femoral Artery Bleeding Model in Pigs ...... 239

6.4.8 Hemostatic Efficacy Evaluation in Femoral Artery Bleeding Model ...... 240

6.4.9 Statistical Analysis ...... 241

6.5 Acknowledgements...... 241

6.6 Author Contributions ...... 242

6.7 Competing Financial Interests ...... 242

6.8 References ...... 243

Chapter 7 : Trauma-targeted Delivery of for Augmenting Hemostasis

...... 255

7.1 Abstract ...... 255

7.2 Introduction ...... 257

7.3 Materials and Methods ...... 261

7.3.1 Materials ...... 261

7.3.2 TXA-loaded targeted nanovesicle (TTNV) development and characterization ...... 262

7.3.3 In vitro analysis of the effect of free TXA and TTNVs on tPA-induced fibrinolysis ...... 264

7.3.4 In vivo safety of FMP-decorated empty nanovesicles (TNVs) and TXA-loaded TTNVs ...... 266

7.3.5 Development of rat liver hemorrhage traumatic injury model ...... 267

vi 7.3.6 Evaluation of TTNVs in rat liver hemorrhage model ...... 268

7.3.7 Statistical analysis ...... 269

7.4 Results ...... 270

7.4.1 Characterization of TTNVs and TXA release kinetics ...... 270

7.4.2 ROTEM analysis of the effect of TXA-loaded in human blood ...... 271

7.4.3 In vivo safety evaluation of various doses of FMP-decorated liposomes ...... 274

7.4.4 Effect of clot-targeted delivery of TXA-loaded nanovesicles in rat liver injury model ...... 276

7.5 Discussion ...... 280

7.6 Conclusion ...... 284

7.7 Acknowledgement ...... 285

7.8 References ...... 286

Chapter 8 : Design and Administration of to Prevent Pseudoallergic

Responses in Intravenous Hemostatic Management of Bleeding ...... 296

8.1 Introduction ...... 296

8.2 Materials and Methods ...... 298

8.2.1 Materials ...... 298

8.2.2 Manufacture of SynthoPlateTM 1.0 ...... 299

8.2.3 Manufacture of SynthoPlateTM 1.5 ...... 300

8.2.4 Manufacture of SynthoPlateTM 2.0 ...... 300

8.2.5 In vivo Safety Studies in Pigs...... 300

8.2.6 Femoral Artery Bleeding Model in Pigs ...... 303

8.2.7 Hemostatic Efficacy Evaluation in Femoral Artery Bleeding Model ...... 304

8.2.8 Statistical Analysis ...... 304

vii 8.3 Results ...... 305

8.3.1 Synthesis of SynthoPlate PEG variants ...... 305

8.3.2 Evaluation of SynthoPlate PEG variants ...... 305

8.3.3 Evaluation of SynthoPlate Administration Protocol Safety ...... 309

8.3.4 Evaluation of SynthoPlate Administration Protocol Efficacy ...... 310

8.4 Discussion ...... 312

8.5 References ...... 314

Chapter 9 : Conclusion and Future Directions ...... 319

9.1 Conclusion ...... 319

9.2 Future Directions ...... 322

9.2.1 Nanovesicle Design ...... 322

9.2.2 Nanoparticle Evaluation ...... 323

9.2.3 Nanoparticle Technology Translation ...... 324

9.3 References ...... 324

Bibliography ...... 327

viii List of Tables

Table 1-1. Classification of hemorrhagic shock. Reproduced with permission from

Cannon, JW10, Copyright Massachusetts Medical Society...... 4

Table 1-2. Principles of damage control resuscitation. Reproduced with permission from Cannon, JW10, Copyright Massachusetts Medical Society...... 5

Table 2-1 Topical and Externally Administered Hemostatic Biomaterials from

Natural Sources ...... 50

Table 2-2 Topical and Externally Administered Hemostatic Biomaterials from

Synthetic Sources ...... 59

Table 8-1. Characteristics of SynthoPlate PEG variants ...... 299

ix List of Figures

Figure 1-1. Pathobiology of Hemorrhagic Shock...... 2

Figure 1-2 Schematic of the complex mechanism of blood vessel hemostasis ...... 8

Figure 2-1 Representative chemical structures of cotton (cellulose) biopolymers

and its derivatives ...... 34

Figure 2-2 Schematic of concomitant roles of and fibrin(ogen) in

propagating the formation of hemostatic clots via activation of platelets and

formation of fibrin mesh...... 36

Figure 2-3 Multiscale schematic representation of fibrillar collagen structure ...... 40

Figure 2-4 Chemical structures of some polysaccharide polymers ...... 48

Figure 3-1. Selected results from hemostasis-relevant studies carried out with

PolyP-coated silica particle systems (A and B) and thrombin-loaded CaCO3-based

particles mixed in TXA-NH3+ (C, D and E) ...... 117

Figure 3-2 Schematic of conversion to fibrin and assembly of cross-

linked fibrin biopolymeric mesh catalyzed by thrombin (FIIa) and FXIIIa, and

hemostatic materials and technologies inspired by these mechanisms...... 120

Figure 3-3 Selected results from hemostatic studies carried out with platelet-like

microgel particles (PLPs, A-D) and with polySTAT injectable polymer systems (E-

G) ...... 122

Figure 3-4 Schematic of platelet’s injury site-selective adhesion and aggregation

mechanism, and various hemostatic technologies inspired by these mechanisms...... 125

Figure 3-5 Selected results from hemostatic studies with infusible platelet

membrane (IPM) technology ...... 128

x Figure 3-6 Selected results from studies involving technologies that leverage platelet’s bleeding site-selective adhesion mechanisms ...... 131

Figure 3-7 Selected results from studies carried out with various fibrinogen-coated particle designs...... 133

Figure 3-8 Selected results from studies carried out with various particle platforms surface-decorated with fibrinogen-relevant RGD or H12 peptides to mimic platelet aggregation mechanism ...... 138

Figure 4-1. Platelet aggregometry studies where undecorated latex beads, or H12 peptide-decorated latex beads, or latex beads co-decorated with H12 peptides and rGPIbα fragments were added to ADP-activated platelets ...... 170

Figure 4-2. Manufacture and characterization of SynthoPlate constructs...... 171

Figure 4-3. Ex vivo aggregometry studies for SynthoPlate with mouse platelets...... 172

Figure 4-4. Effect of SynthoPlate on secondary hemostasis ...... 174

Figure 4-5. Development of thrombocytopenia model in mouse and evaluation of

SynthoPlate capability in correcting tail transection bleeding time in thrombocytopenic mouse...... 175

Figure 4-6. Characterization of SynthoPlate‐mediated hemostatic clot in transected tail of thrombocytopenic mouse...... 176

Figure 4-7. Evaluation of systemic pro‐coagulant risk and biodistribution of

SynthoPlate in mice...... 177

Figure 5-1 Experimental timeline, liver laceration model, and SynthoPlate mechanism ...... 192

xi Figure 5-2 Analysis of SynthoPlate pretreatment on blood loss and hemodynamics

during hemorrhagic shock...... 195

Figure 5-3 In vivo biodistribution of SynthoPlate particles ...... 197

Figure 5-4 Analysis of SynthoPlate transfusion on blood loss and hemodynamics

as a rescue strategy in acute, hemorrhagic shock ...... 199

Figure 6-1 Schematic representation of SynthoPlateTM design and mechanism ...... 209

Figure 6-2 Lipid-to-peptide bioconjugation schemes (along with corresponding

chemical structures) for synthesizing DSPE-PEG-VBP, DSPE-PEG-CBP and

DSPE-PEG-CMP molecules, that were used to manufacture the SynthoPlateTM

nanoconstruct...... 210

Figure 6-3 [A] SynthoPlateTM diameter analysis over a 6-month period and [B]

Analysis after sterilization with filtration or E-beam. [C] Representative fluorescent images and quantitative analysis of surface-averaged fluorescence intensity of adhered SynthoPlateTM. [D] Aggregometry analysis ...... 212

Figure 6-4 Representative Dynamic Light Scattering (DLS) data of SynthoPlateTM

size distribution characterization over a 6-month period ...... 213

Figure 6-5 Bacteriostatic and fungistatic analysis of SynthoPlateTM suspension

exposed to E-beam sterilization and challenged with 6 different organisms ...... 214

Figure 6-6 Representative Lumi-Aggregometry raw data from studies on

SynthoPlateTM (fresh unsterilized versus sterilized) interaction with platelets

(resting or agonist-activated) in PRP ...... 215

Figure 6-7 MTT-based metabolic activity analysis data of human 3T3 fibroblasts

in culture incubated with SynthoPlateTM ...... 216

xii Figure 6-8 [A] Average vitals and [B] average blood lab values over the course of the experiment of uninjured pigs upon administration of unmodified particles or

SynthoPlateTM [C] Analysis of C3:C3a plasma concentration ratios in blood [D]

Biodistribution analysis from harvested organs...... 218

Figure 6-9 Representative hematoxylin and eosin (H&E) stained histology images

(32x magnification) of organ samples from pigs treated with saline, unmodified particles or SynthoPlateTM particles ...... 220

Figure 6-10 Schematic representation of pig femoral artery hemorrhage model setup ...... 221

Figure 6-11 Hemostatic efficacy analysis in injured pigs ...... 222

Figure 6-12 Schematic representation and representative images of hemostasized injury site in the femoral artery of pigs treated with SynthoPlateTM or saline or control particles: [A] Representative hematoxylin and eosin (H&E), [B] bright field, and [C-E] fluorescent images of the site of injury (transected artery) with the injured vessel components and hemostatic clot in view for injured pig treated with

SynthoPlateTM ...... 224

Figure 6-13[A] Complement (C3:C3a) analysis data on drawn blood from pigs and

[B] biodistribution data from pigs subjected to femoral artery injury and administered intravenously with unmodified particles or SynthoPlate nanoconstructs...... 225

Figure 6-14 [A] Representative ex vivo Lumi-aggregometry analysis data and [B]

ROTEM analysis data (CT, MCF and A10 parameters) of blood samples drawn

xiii from pigs after being subjected to femoral artery injury and administered

intravenously with saline or unmodified particles or SynthoPlate nanoconstructs...... 227

Figure 7-1. Cellular and molecular components of hemostasis regulation ...... 258

Figure 7-2. Ascorbic Acid-based reaction for assaying TXA, and representative

calibration curve using this assay, based on which TXA release kinetics from

nanovesicles was estimated...... 264

Figure 7-3. Representative instrument set-up, characteristic profile, actual

instrument and relevant measured parameters in Rotational Thromboelastometry

(ROTEM) analysis of whole blood ...... 265

Figure 7-4. Representative confocal images showing red fluorescent FMP- decorated nanovesicles can bind and localize with activated platelets within

crosslinked fibrin-rich clots formed from PRP in presence of thrombin ...... 269

Figure 7-5. [A] Schematic of the manufacturing process of TTNVs [B] Envisioned targeted delivery of TXA at the trauma site from TTNVs [C] Representative DLS characterization results [D] Representative encapsulation efficiency of TXA in

TTNVs [E] Representative release profile of TXA from TTNVs ...... 271

Figure 7-6. ROTEM analysis of clot characteristics of whole blood spiked with tPA and treated with saline (no TXA), free TXA or TXA loaded within nanovesicles ...... 272

Figure 7-7. In vivo safety evaluation with TNVs as well as TTNVs ...... 274

Figure 7-8. [A] Schematic of rat liver injury hemorrhage model set-up [B] Post- injury vitals of animals [C] Blood loss analysis post-injury and treatment

xiv administration [D&E] 1-hr and 72-hr survival analysis post-injury and treatment administration ...... 275

Figure 7-9. Representative images from histopathologic analysis (H & E staining) of excised tissue sections from liver injured rats administered with various treatments ...... 277

Figure 7-10. Representative confocal microscopy images of immunostained tissue sections from injured liver site ...... 280

Figure 8-1 Schematic representation of pig femoral artery hemorrhage model setup

...... 303

Figure 8-2 Vital analysis in safety pigs administered SynthoPlate PEG variants ...... 307

Figure 8-3 Images of pig skin responses after SynthoPlate PEG variant administration ...... 308

Figure 8-4 Analysis of C3a plasma concentration in blood drawn from pigs ...... 309

Figure 8-5 Vital analysis in safety pigs administered SynthoPlate 2.0 at 50ml

(bolus) or 500ml (infusion) ...... 310

Figure 8-6 Vital analysis in injury pigs administered SynthoPlate 2.0 at 50ml

(bolus) or 500ml (infusion) ...... 311

Figure 8-7 Hemostatic efficacy analysis in injured pigs ...... 312

xv Acknowledgements

I would like to thank Dr. Anirban Sen Gupta for allowing me to join his lab, for supporting me in carrying out this great research during my PhD and for being a better mentor than I could have asked for. I would like to acknowledge my committee members past and present

(Drs. Nicholas Ziats, James Anderson, Agata Exner, Howard Meyerson, Clive Hamlin,

Nicole Steinmetz and Jeffrey Capadona) for their advice and guidance throughout this entire process. I am grateful to the Medical Scientist Training Program, and American

Heart Association for supporting me during my PhD.

xvi List of Abbreviations

Abbreviation Full Name

A10 amplitude 10 minutes after CT

AAAS American Association for the Advancement of Science

ADP adenosine di-phosphate

Ala alanine (A)

Alb albumin

ANOVA analysis of variance

aPC activated C

APTT activated partial thromboplastin time

Arg arginine (R)

Asp aspartic acid (D)

ATLS Advanced Trauma Life Support

BCOP blood colloidal osmotic pressure

BEH ethylene bridged hybrid

BL black

bpm beats per minute

bpm breaths per minute

BSA bovine serum albumin

C-A-T® Combat Application Tourniquet

Ca++ Calcium

CARPA complement activation-related pseudoallergy

xvii CBP collagen binding peptide

CD cluster of differentiation

cDNA complementary deoxyribonucleic acid

CFT clot formation time

CO2 carbon dioxide

CP control particles

CRASH-2 Clinical Randomization of an in Significant Hemorrhage 2 trial CT clotting time

Cys cystine (C)

DAPI 4′,6-diamidino-2-phenylindole

DFSD Dry Fibrin Sealant Dressing

DHPE dihexadecanoyl-sn-glycero-3-phosphoethanolamine

DIC disseminated intravascular

dL deciliter

DLS dynamic light scattering

DMEM Dulbecco’s Modified Eagle Medium

DMSO dimethyl sulfoxide

DSPC 1,2-Distearoyl-sn-glycero-3-phosphocholine

DSPE 1,2-Distearoyl-sn-glycero-3-phosphoethanolamine

EACA e-

e-beam electron beam

EC endothelial cell

xviii EDTA ethylenediamine tetraacetic acid

EKG electrocardiogram

ELAD elastic adhesive

ELISA enzyme-linked immunosorbent assay

EMT Emergency Medical Tourniquet

EU endotoxin units

FBS fetal bovine serum

FDA Federal Drug Administration

FDPs fibrin degradation products

FFP fresh frozen plasma

Fg fibrinogen

FII coagulation factor II/ prothrombin

FIIa activated coagulation factor IIa/ thrombin

FIII coagulation factor III/ tissue factor/thrombokinase/ thromboplastin

FITC fluorescein isothiocyanate

FIX coagulation factor IX

FIXa activated coagulation factor IX

FMP fibrinogen mimetic peptide

FV coagulation factor V

FVa activated coagulation factor V

FVII coagulation factor VII

FVIIa activated coagulation factor VII

xix FVIII coagulation factor VIII

FX coagulation

FXa activated coagulation factor X

FXI coagulation factor XI

FXIa activated coagulation factor XI

FXIII coagulation factor XIII

FXIIIa activated coagulation factor XIII

Gln glutamine (Q)

Glu glutamic acid (E)

Gly glycine (G)

GMP good manufacturing practice

GP glycoprotein

GRF gelatin-resorcin-formalin

H&E hematoxylin and eosin

Hct hematocrit

Hg hemoglobin

His histidine (H)

HLA human leukocyte antigen

HPMA N-(2-Hydroxypropyl) methacrylamide

HR heart rate

HUS hemolytic urine syndrome

I.V. Intravenous

xx IACUC Institutional Animal Care and Use Committee

ICU Intensive Care Unit

IgE immunoglobulin E

Ile isoleucine (I)

INR international normalized ratio

IPM infusible platelet membrane

IRB Institutional Review Board

ITP Immune Thrombocytopenia

kg kilograms

L Liter

LAL limulous ameobocyte lysate

Leu leucine (L)

LI30 lysis index after 30 minutes

Lys lysine (K)

mal maleimide

MALDI-TOF matrix assisted laser desorption ionization-time of flight

MAP mean arterial pressure

MATTERs Military Application of Tranexamic Acid in Trauma Emergency Resuscitation study MCF maximum clot firmness

MCHC mean corpuscular hemoglobin concentration

MCV fL mean corpuscular volume in femtoliters

mg miligram

xxi min minutes

ml mililiters

ML maximum lysis

mmHg milimeters of mercury

mPEG methoxy(polyethylene glycol)

mRNA messenger ribonucleic acid

MW molecular weight

MWCO molecular weight cutoff

NHLBI National Heart Lung and Blood Institute

NHSMA N-hydroxysuccinimide methacrylate

NIGMS National Institute of General Medical Sciences

NIH National Institutes of Health

nm nanometer

NS normal saline

OC

oC degrees Celsius

ORC oxidized regenerated cellulose

PAI-1 plasminogen activator inhibitors-1

PAI-2 plasminogen activator inhibitors-2

PAMPer Prehospital Air Medical Plasma clinical trial

PaO2 partial pressure of oxygen

PATCH Pre-hospital Anti-fibrinolytic for Traumatic and Haemorrhage

xxii PBS phosphate buffered saline

PC platelet concentrates

PEG poly (ethylene glycol)

PEO poly (ethylene oxide)

PF4 platelet factor 4

PGA poly (glycolic acid)

Phe phenylalanine (F)

Plg plasminogen

PLGA poly (lactic co-glycolic acid)

PLL poly (L-lysine)

PLPs platelet-like particles

Plt platelet

pNIPAm- poly (N-isopropyacrylamide-co-acrylic acid) AAc poly-HEMA poly (2-hydroxyethyl methacrylate)

PolyP polyphosphate

PoyP-SNP PolyP loaded in silica nanoparticles

PPO poly (propylene oxide)

PPP platelet poor plasma

Pro proline (P)

PROPPR Pragmatic, Randomized, Optimal, Platelet and Plasma Ratios clinical trial

PRP platelet rich plasma

PS penicillin-streptomycin

xxiii PS phosphatidylserine

PT prothrombin time

QC-ACS Quikclot-Advanced Clotting Sponge

QCG Quikclot Combat Gauze

R&D research and development

RBC red blood cells

RDH Rapid Deployment Gauze

RDW-CV red blood cell distribution width- coefficient of variation

RES reticulo-endothelial system

RGDS arginine-glycine-aspartic acid-serine

rGP recombinant glycoprotein

RhB rhodamine B

RMT Ratcheting Medical Tourniquet

ROTEM rotational thromboelastometry

rpm revolutions per minute

SaO2 oxygen saturation

SC Synthocyte particles

SD standard deviation

SDS sodium dodecyl sulfate

Ser serine (S)

serine protease inhibitor

SK streptokinase

xxiv SMC smooth muscle cell

SNP silica nanoparticles

SOF-TT Special Operations Forces Tactical Tourniquet

sPLA2 secreted phospholipase A2

SpO2 peripheral capillary oxygen saturation

STAAMP Study of Tranexamic Acid during Air Medical Prehospital Transport

TACTIC Trans-agency Collaboration for Trauma Induced Coagulopathy consortium

TAFI thrombin-activated fibrinolysis inhibitor

TBS tris-buffered saline

TCCC Tactical Combat Casualty Care

TCNVs TXA-loaded control (untargeted) nanovesicles

TCP thrombocytopenia

TEMPO 2,2,6,6-tetramethylpiperidine-1-oxyl radical

TF tissue factor

THPTA tris(3-hydroxypropyltriazolylmethyl)amine

Thr threonine (T)

TNV targeted nanovesicles

tPA tissue plasminogen activator

Trp tryptophan (W)

TTNV TXA-loaded targeted nanovesicles

TTP thrombotic thrombocytopenic purpura

TXA tranexamic acid

xxv Tyr tyrosine (Y)

U.S. United States

g microgram

UHCMC𝜇𝜇 University Hospitals Cleveland Medical Center

uPA urokinase

UPLC ultra high performance liquid chromatography

USA United States of America

USAISR US Army Institute of Surgical Research

UV ultraviolet

Val valine (V)

VBP von Willebrand Factor binding peptide

vWF von Willebrand Factor vWF-A1

WBC white blood cells

WHO World Health Organization

WT wild type

WWII World War II

xxvi Platelet-inspired Nanomedicine for the Hemostatic Management of Bleeding

Complications in Thrombocytopenia and Trauma

Abstract

by

DASHAWN A. HICKMAN

Traumatic non-compressible hemorrhage and coagulopathy remain leading causes of civilian and military mortalities, especially in pre-hospital and limited resource scenarios.

Patients with thrombocytopenia are at an even higher risk due to their low number of platelets, a major blood component involved in clotting. Current clinical strategy to treat this involves massive transfusion of whole blood or blood components (e,g. RBC: platelet: plasma at 1:1:1 ratio), but such blood products present issues of limited availability and portability, high risk or contamination and short shelf-life, that substantially limit their widespread use in resource-limited scenarios. Therefore, there remains a significant need for intravenous hemostatic products that allow hemorrhage control and coagulopathy mitigation, while avoiding the above issues associated with blood-derived products. To address this need, my research has focused on developing an I.V.-administrable platelet- inspired nanomedicine system for the hemostatic management of bleeding complications in thrombocytopenia and trauma. To this end, the primary focus of this dissertation is on evaluating the mechanism, safety and efficacy of platelet-inspired nanomedicine systems

xxvii in the hemostatic management of clinically relevant bleeding complications. My overall

hypothesis is that the design of nanoscale delivery systems with platelet-inspired

heteromultivalent surface-modifications, can enable injury site-specific biointeractions and

to enhance platelet-mediated hemostatic processes, leading to decreased

blood loss and increased survival. I have tested this hypothesis in two platelet-inspired

nanomedicine designs: (i) in a synthetic nanoparticle based functional mimic of platelets

(termed SynthoPlateTM) that can leverage and amplify the body’s physiological clotting

mechanisms specifically at the bleeding site and (ii) in a platelet-targeted nanovehicle for

injury-site selective delivery of hemostasis augmenting drugs. The SynthoPlateTM

technology was designed and evaluated in (i) tail-bleeding model in thrombocytopenic mice, (ii) femoral artery injury model in pigs, and (iii) liver resection hemorrhage model in mice, while the injury site-targeted hemostatic drug delivery strategy was evaluated in

liver resection hemorrhage model in rats. Beyond this, I have also investigated ways to

refine nanomedicine design to improve systemic safety, by altering surface charge and

hydrophilicity, as well as modulating the administration protocol. In summary, my research

has established the potential for platelet-inspired for the hemostatic

management of bleeding complications. I believe that these nanomedicine technologies

have significant potential for clinical translation in the future phases of this research.

xxviii Chapter 1: Bleeding Complications in Thrombocytopenia and Trauma

Some content based on: Hickman DA, Pawlowski CL, Sekhon UDS, Marks J, Gupta AS. Biomaterials and Advanced Technologies for Hemostatic Management of Bleeding. Adv Mater. 2017 Nov 22. doi: 10.1002/adma.201700859.

1.1 Introduction

Blood, a fluid connective tissue composed of red blood cells (RBCs), white blood cells

(WBCs), platelets, and non-cellular liquid (plasma) containing salts, nutrients and ,

is present at a volume of approximately 5 liters in the average (70 kg body

weight). Blood is responsible for transport of gases and nutrients to tissues as well as

providing immune surveillance and hemostatic responses as needed. Hence, loss of blood

can result in a variety of pathologic scenarios that can lead to tissue morbidities and

mortalities. For example, in traumatic in both battlefield and civilian conditions,

significant blood loss from truncal, junctional and internal non-compressible injuries can

result in significant pre-hospital (and potentially preventable) mortalities stemming from

hypothermia, coagulopathy, , acidosis and multiple organ failure1–6. Also, certain

congenital or disease-associated conditions (e.g. coagulation factor deficiencies and

platelet dysfunctions) as well as drug-induced effects (e.g. bone marrow suppression due to chemotherapy and radiotherapy in cancer patients) can put certain patient populations at high bleeding risks7–9. Therefore, management of various bleeding complications remains

a highly significant clinical area, and substantial interdisciplinary research efforts are

focused on the development of materials and technologies for efficient hemostatic

management of bleeding.

1

Figure 1-1. Pathobiology of Hemorrhagic Shock. Reproduced with permission from Cannon, JW10, Copyright Massachusetts Medical Society.

1.2 Trauma

Trauma is an injury that can lead to severe disability or death and these injuries can vary

in severity and location. As a leading cause of mortality worldwide with a U.S. morbidity

of over 45 million each year, trauma represents 30% of ICU admissions11,12. Trauma is

expected to rise to the third leading cause of disability worldwide by 203013. Mechanisms

2 of trauma include motor vehicle collisions, motorcycle accidents, jumps/falls greater than

20 feet, stabbings, gunshot , plane crashes, blast injuries and being crushed. An

analysis of over 500 prehospital deaths from trauma revealed that ~30% of the injuries

were potentially survivable given current in-hospital treatment options14. Of those

potentially survivable deaths, more than half of the patients died from hemorrhage which

lead to shock14,15. Hemorrhagic shock is a severe loss of blood that leads to poor tissue perfusion and oxygen delivery. On a cellular level, the lack of oxygen forces cells to switch to anaerobic metabolism which leads to the accumulation of lactic acid, inorganic phosphates and free radicals16. If reperfusion does occur, this could lead to ischemia- reperfusion injury by flushing this build-up of anaerobic metabolic by-products to healthy tissue17. If normal blood flow is not re-established, the cells will eventually undergo

necrosis and apoptosis. On a tissue level, this hypoperfusion can lead to end-organ damage and multi-organ failure. Figure 1-1 depicts how massive loss of blood causes simultaneous activation of the hemostasis and fibrinolytic systems, and compensatory mechanisms that lead to the phenotype of coagulopathy, hypothermia, and acidosis which result in even more dysregulation and eventual death.10

Four classes of shock are described by the Advanced Trauma Life Support (ATLS) manual

produced by the American College of Surgeons (Table 1-1). Class 1 hemorrhage is defined

as having blood volume loss up to 15% with a normal blood pressure, pulse pressure and

respiratory rate. Class II hemorrhage is a 15-30% blood loss with tachycardia (100-

120bpm), tachypnea (20-30bpm), normal blood pressure and narrowed pulse pressure.

Class III involves 30-40% blood loss, tachycardia (120-140bpm), tachypnea (30-40bpm),

3

Table 1-1. Classification of hemorrhagic shock. Reproduced with permission from Cannon, JW10, Copyright Massachusetts Medical Society.

confusion. Class IV hemorrhage patients have lost more than 40% of their blood volume

leading to a significant drop in blood pressure and mental status with a compensatory

increase in heart rate (greater than 140bpm), a narrowed pulse pressure, and tachypnea

(greater than 35bpm). They will also present with cold and pale skin and a delayed capillary

refill18. Management of trauma patients is done on a case by case basis, but major priorities

for all patients include volume resuscitation and hemorrhage control (Table 1-2). Current methods emphasize the use of red blood cells, plasma and platelets10.

1.3 Thrombocytopenia

Risk of hemorrhage in trauma is increased in patients with thrombocytopenia.

Thrombocytopenia is a diagnosis given with a platelet count below the lower limit of

normal (<150,000/µL [150 x 109/L] for adults). There are three levels of thrombocytopenia; mild (platelet count 100,000 to 150,000/µL), moderate (50,000 to 99,000/µL), and severe

(<50,000/µL)19. While a normal platelet count range is 150,000 to 450,000/µL with means

of 266,000 and 237,000/µL for men and

4 Table 1-2. Principles of damage control resuscitation. Reproduced with permission from Cannon, JW10, Copyright Massachusetts Medical Society.

women respectively20, most individuals have little variation in their platelet count

(differences in platelet count greater than 98,000/µL occurred in less than 0.1% of

people)20. This wide population range and narrow individual range is important to note

because a small proportion of the population (2.5%) will have a baseline platelet count

lower than normal range, but will have no clotting issues. At the same time, an individual

can have a clinically significant decrease in platelet count and still be in the normal

population normal range.

Platelets are derived from megakaryocytes in the bone. Major causes of thrombocytopenia

include decreased platelet production in the bone marrow; peripheral platelet destruction

5 or consumption; dilution; and sequestration. Bone disorders that can be caused by nutrient deficiencies, myelodysplastic syndromes, and sepsis that lead to dysregulation

of megakaryocytes can lead to decreases in platelet production21. Platelet destruction can

be caused by antibodies in immune thrombocytopenia (ITP). ITP is an autoimmune disease

where the body creates antibodies targeting its own platelets leading to their destruction by

the body’s immune cells. This mechanism can also be drug induced by drugs such as

heparin and quinine22. Drugs and treatments such as chemotherapy or radiation can

suppress the bone marrow and lead to low platelet counts. Platelets can be consumed by

thrombus in disorders such as disseminated intravascular coagulation (DIC), thrombotic

thrombocytopenic purpura (TTP), and hemolytic urine syndrome (HUS) where

dysregulated coagulation leads to depletion of circulating platelets. Massive transfusions

can also lead to the dilutions of platelets and a low platelet count. About 1/3 of the blood

platelets reside in the spleen and in cases of splenomegaly, the number of platelets in the

spleen can increase, decreasing the total amount of circulating platelets, without decreased

the total platelet count leading to thrombocytopenia23.

There are a number of principles used to guide the management of patients with thrombocytopenia depending on the etiology of their low platelet count, but one consistent principle is in the emergency management of bleeding. Critical bleeding in a patient with severe thrombocytopenia requires immediate platelet transfusion24.

6 1.4 Complex Mechanism of Hemostasis

The word ‘hemostasis’ was coined by ancient Greeks from the terms haíma (meaning blood) and stasis (meaning stoppage) to describe the phenomenon of blood stagnation when alum came in contact with a wound25. Although scattered evidences of utilizing blood products and application of coagulation/hemostasis concepts are found in ancient Greek and Roman history, as well as in reports of transfusion and organ transplants around late

18th and early 19th centuries, the modern concept of hemostatic mechanisms is credited to the seminal report written by Paul Morawitz in 190526,27. In his report, Morawitz emphasized the role of platelets, ‘thrombokinase’ (now known as tissue factor), calcium, prothrombin and fibrinogen in promoting blood coagulation. Later in the 20th century, additional coagulation factors were identified and characterized, and the concepts of coagulation being guided by a cascade of enzymes and co-factors (i.e. the intrinsic and extrinsic coagulation cascades) leading to the final output of fibrin formation (common pathway of cascades) were put forward28,29. According to the current (and still evolving) understanding of the process, the body’s natural mechanisms of hemostasis are rendered by a complex, spatio-temporally regulated sequence of responses involving a combination of cellular (e.g. platelets and tissue factor-bearing cells) and plasma (e.g. coagulation factors) components, as depicted in the schematic of Figure 1-2.

In the absence of injury, healthy endothelial cells lining the luminal wall of blood vessels avoid blood clotting by secretion of heparin-like molecules, thrombomodulin, nitric oxide and prostacyclin, as well as, sterically hindering of clot-relevant proteins on the vessel wall due to presence of endothelial glycocalyx. Tissue injury and bleeding result in

7

Figure 1-2 Schematic of the complex mechanism of blood vessel hemostasis. Vessel injury can lead to endothelial activation and denudation resulting in secretion and deposition of von Willebrand Factor (vWF) and exposure of collagen at the injury site, as well as, exposure of tissue factor (TF) bearing cells at the site; vWF and collagen exposure allows platelet adhesion and activation, while TF exposure allows extrinsic pathway of coagulation to propagate and produce moderate amounts of thrombin (FIIa) that activates other coagulation factors in the intrinsic pathway; activated platelets aggregate via fibrinogen (Fg) mediated interaction with platelet surface integrin GPIIb-IIIa to form a platelet plug (primary hemostasis) that staunches bleeding; the surface of aggregated active platelets exposes negatively charged phospholipids that allow co-localization and further activation of coagulation factors to form the prothrombinase (FVa + FXa + FII) complex in presence of calcium (Ca++), leading to amplified generation of thrombin (FIIa) that breaks down fibrinogen (Fg) to fibrin; fibrin self-assembles and undergoes further crosslinking by action of FXIIIa to form a dense biopolymeric mesh that forms the hemostatic clot and arrests flow of blood components (secondary hemostasis).

endothelial damage, leading to vasoconstriction to reduce blood flow out of the injury site and secretion/exposure of pro-coagulant proteins and factors. Platelets rapidly respond to the bleeding site by undergoing adhesion (primarily to von Willebrand Factor and sub- endothelial collagen), activation and inter-platelet aggregation at the site to form the platelet plug30–33, a process commonly known as ‘primary hemostasis’. In tandem, exposed

8 sub-endothelial collagen, von Willebrand factor (vWF) secreted from injured endothelium and activated platelets, and tissue factor (Coagulation Factor III, formerly known as

thrombokinase or thromboplastin) on sub-endothelial matrix and localized leukocytes lead

to initiation, amplification and propagation of the coagulation cascade, culminating in

thrombin-catalyzed formation of fibrin from fibrinogen, a process commonly known as

‘secondary hemostasis’34–36. Critical upstream steps of this process (e.g. conversion of

prothrombin to thrombin) are greatly amplified by anionic lipids (phosphatidylserines) on

the membrane of active platelets as well as by anionic polymers (polyphosphates or PolyP)

secreted by active platelets37–44. Pro-hemostatic active platelets also secrete molecules

adenosine di-phosphate (ADP) and platelet factor 4 (PF4) that can modulate hemostatic

mechanisms. The fibrin, formed as the final product of the coagulation cascade, associates

into a cross-linked biopolymeric mesh facilitated by activated coagulation factor XIII

(FXIIIa) and platelet-secreted polyphosphate (PolyP) to secure the platelet plug and other

blood components at the bleeding site and form the final clot45,46. The clot-incorporated

active platelets also facilitate clot retraction and healing mechanisms47–50. Post-healing of

the injury site, the mature fibrin clot is lysed through the action of plasmin, which is

generated from the zymogen plasminogen on the surface of the fibrin, as well as, on

neighboring cell surfaces by the action of tissue plasminogen activator (tPA) or urokinase

(uPA)51. The tPA is produced from endothelial cells while uPA is produced from

monocytes, macrophages and urinary epithelium. Plasmin-induced proteolysis of fibrin

results in formation of soluble fibrin degradation products (FDPs), which can have some immunomodulatory and chemotactic functions relevant to healing phases. In healthy individuals the clot formation and fibrinolytic systems are highly regulated to ensure

9 hemostatic balance, and any dysregulation can lead impaired or weak clot formation (poor

hemostasis and re-bleeding) or overly strong occlusive clot growth (thrombosis). For

example, the plasminogen activators (tPA and uPA) and plasmin action are regulated by

local high concentration of serine protease inhibitor (serpin) molecules like plasminogen

activator inhibitors-1 (PAI-1), plasminogen activator inhibitor-2 (PAI-2) and α-2

antiplasmin52, as well as non-serpin molecules like α-2 macroglubulin and thrombin- activated fibrinolysis inhibitor (TAFI)53,54. Fibrin-bound tPA shows greatly enhanced catalytic efficiency of plasminogen activation compared to solution phase tPA55. In

parallel, plasmin is protected from inhibition by α-2 antiplasmin upon binding to fibrin56.

Thus, a combination of feedback mechanisms of fibrin formation and fibrin destruction

maintains the precise spatiotemporal regulation of hemostasis.

1.5 Management of Traumatic Bleeding

Based on this complex concert of mechanisms in clotting, biomaterials-based approaches

to render and augment hemostasis have focused on mimicking and leveraging the various

mechanistic aspects, including constriction (pressure), platelet (primary hemostasis)-

relevant components and coagulation (secondary hemostasis)-relevant components57. The optimum requirements for a hemostatic material (and technology) is that it (i) should be applicable and adaptable to a variety of actively bleeding wounds, (ii) should act rapidly to reduce blood loss and maintain this hemostatic condition for long durations if needed, (iii) should be easily manufacturable, sterilizable and portable, (iv) should be stable under a variety of atmospheric and ambient conditions, (v) should be easily usable by non- specialized personnel if needed and (vi) should be reasonably biocompatible so as to not

10 induce any short-term or long-term adverse effect in the body57. To this end, for externally

visible and accessible (often compressible) injuries, a variety of biomaterials-based

technologies in the form of powders, bandages, sprays, foams, gels, tourniquets and

tamponades have been developed58–64. In contrast, for management of internal (often non-

compressible) bleeding, the clinical gold standard is the transfusion of whole blood or

blood components (RBC, plasma and platelets)65–70, and use of fibrinogen concentrate or

recombinant coagulation factors in selected groups of patients (e.g. recombinant Factor

VIIa used in hemophilia patients)71–75. Patients with thrombocytopenia in the intensive care

unit can range from 8%-67%. Up to 30% of these patients may require platelet transfusions,

primarily for purposes of hemorrhage prophylaxis70. The importance and early

administration of these clinical gold standards has been validated by recent clinical trials.

The Prehospital Air Medical Plasma (PAMPer) clinical trial evaluated a total of 501 patients and concluded that prehospital administration of thawed plasma was safe and

resulted in lower 30-day mortality and a lower median prothrombin time ratio than

standard-care resuscitation76. The PROPPR (Pragmatic, Randomized Optimal Platelet and

Plasma Ratios) trial demonstrated the importance of platelets in particular for the

management of traumatic bleeding. This study determined that early platelet administration

is associated with improved hemostasis and reduced mortality in traumatic bleeding

patients compared to patients who received blood products without platelets77. Despite these data, donor-derived blood and its components often have limited availability, require meticulous type matching, pose issues of high pathologic contamination or immunogenic risks, and have limited portability and short shelf-life. These issues are particularly challenging for transfusion of platelets, which present challenges of alloimmunization and

11 refractoriness, have a high bacterial contamination risk and have a shelf-life of only 3-7

days78–85. Therefore, significant research efforts are currently being focused on minimizing

contamination and improving storage life of platelets86–89 as well as on developing in vitro bioreactor technologies for production of ‘donor-independent platelets’ from cultures of precursor cells90–92. In parallel, robust interdisciplinary research efforts are being directed

on developing biomaterials-based technologies that can be administered intravenously and

can mimic, amplify and leverage various mechanistic components of hemostasis to rapidly

staunch bleeding. The next two chapters will comprehensively and critically review the various externally used and internally applicable hemostatic technologies in terms of materials, clotting mechanisms and current state-of-art, and discuss the challenges and opportunities associated with these technologies to help interdisciplinary advancement of the field.

1.6 Acknowledgments

This work was supported by the National Heart Lung and Blood Institute (NHLBI) of the

National Institutes of Health under award numbers R01 HL121212 (PI: Sen Gupta). The content of this publication is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

1.7 References

1. Kauvar DS, Lefering R, Wade CE. Impact of Hemorrhage on Trauma Outcome:

An Overview of Epidemiology, Clinical Presentations, and Therapeutic

Considerations. 2006 Jun;60(Supplement):S3–S11. Available from:

12 https://insights.ovid.com/crossref?an=00005373-200606001-00002

2. Krug EG, Sharma GK, Lozano R. The global burden of injuries. American Public

Health Association; 2000 Apr;90(4):523–6. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/10754963 PMID: 10754963

3. Champion HR, Bellamy RF, Roberts CP, Leppaniemi A. A profile of combat

injury. 2003;54(5 Suppl):S13–S19. PMID: 12768096

4. Butler FK, Blackbourne LH. Battlefield trauma care then and now. 2012

Dec;73:S395–S402. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201212005-00005

5. Holcomb JB. Optimal use of blood products in severely injured trauma patients.

American Society of ; 2010 Dec 4;2010(1):465–9. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/21239837 PMID: 21239837

6. Simmons JW, Pittet J-F, Pierce B. Trauma-Induced Coagulopathy. 2014 Sep

1;4(3):189–199. Available from: http://link.springer.com/10.1007/s40140-014-

0063-8 PMID: 25587242

7. Furie MB, Atkins M, Mayer R. Clinical Hematology and : Presentation,

Diagnosis and Treatment, 1e. 2003; Available from:

https://mrz3jdqtk01.storage.googleapis.com/MDQ0MzA2NTU2WA==01.pdf

8. Blajchman MA. Substitutes and alternatives to platelet transfusions in

thrombocytopenic patients. Wiley/Blackwell (10.1111); 2003 Jul 1;1(7):1637–

1641. Available from: http://doi.wiley.com/10.1046/j.1538-7836.2003.00332.x

9. Demetri GD. Pharmacologic treatment options in patients with thrombocytopenia.

13 W.B. Saunders; 2000 Apr 1;37:11–18. Available from:

https://www.sciencedirect.com/science/article/pii/S0037196300900489

10. Cannon JW. Hemorrhagic Shock. Massachusetts Medical Society; 2018 Jan

25;378(4):370–379. Available from:

http://www.nejm.org/doi/10.1056/NEJMra1705649

11. WHO | Global Health Estimates. World Health Organization; 2018; Available

from: http://www.who.int/healthinfo/global_burden_disease/en/

12. Mackenzie EJ, Rivara FP, Jurkovich GJ, Nathens AB, Frey KP, Egleston BL,

Salkever DS, Weir S, Scharfstein DO. The National Study on Costs and Outcomes

of Trauma. 2007;63:54–67. Available from:

https://insights.ovid.com/pubmed?pmid=18091213

13. Initial management of trauma in adults - UpToDate [Internet]. Available from:

https://www.uptodate.com/contents/initial-management-of-trauma-in-

adults?search=trauma&source=search_result&selectedTitle=1~150&usage_type=d

efault&display_rank=1

14. Davis JS, Satahoo SS, Butler FK, Dermer H, Naranjo D, Julien K, Van Haren RM,

Namias N, Blackbourne LH, Schulman CI. An analysis of prehospital deaths: Who

can we save? Accreditation Statement AMA PRA Category 1 Creditsi. 2014;

Available from: http://www.aast.org/

15. Teixeira PGR, Inaba K, Hadjizacharia P, Brown C, Salim A, Rhee P, Browder T,

Noguchi TT, Demetriades D. Preventable or Potentially Preventable Mortality at a

Mature Trauma Center. 2007; Available from:

https://insights.ovid.com/pubmed?pmid=18212658

14 16. Chaudry IH. Cellular mechanisms in shock and ischemia and their correction.

American Physiological Society Bethesda, MD ; 1983 Aug;245(2):R117-34.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/6309022 PMID: 6309022

17. Eltzschig HK, Eckle T. Ischemia and reperfusion--from mechanism to translation.

NIH Public Access; 2011 Nov 7;17(11):1391–401. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/22064429 PMID: 22064429

18. American College of Surgeons. Committee on Trauma. ATLS, advanced trauma

life support for doctors. American College of Surgeons; 2008.

19. Williamson DR, Albert M, Heels-Ansdell D, Arnold DM, Lauzier F, Zarychanski

R, Crowther M, Warkentin TE, Dodek P, Cade J, Lesur O, Lim W, Fowler R,

Lamontagne F, Langevin S, Freitag A, Muscedere J, Friedrich JO, Geerts W,

Burry L, Alhashemi J, Cook D. Thrombocytopenia in Critically Ill Patients

Receiving Thromboprophylaxis: Frequency, Risk Factors, and Outcomes. Elsevier;

2013 Oct 1;144(4):1207–1215. Available from:

https://www.sciencedirect.com/science/article/pii/S0012369213606649?via%3Dih

ub

20. Buckley MF, James JW, Brown DE, Whyte GS, Dean MG, Chesterman CN,

Donald JA. A novel approach to the assessment of variations in the human platelet

count. 2000 Mar;83(3):480–4. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/10744157 PMID: 10744157

21. Harker LA, Finch CA. Thrombokinetics in man. American Society for Clinical

Investigation; 1969 Jun;48(6):963–74. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/5814231 PMID: 5814231

15 22. Greinacher A, Selleng K, Warkentin TE. Autoimmune heparin-induced

thrombocytopenia. Wiley/Blackwell (10.1111); 2017 Nov 1;15(11):2099–2114.

Available from: http://doi.wiley.com/10.1111/jth.13813

23. Greenberg EM. Thrombocytopenia: A Destruction of Platelets. 2017;40(1):41–50.

Available from: https://insights.ovid.com/pubmed?pmid=28030481 PMID:

28030481

24. Approach to the adult with unexplained thrombocytopenia - UpToDate [Internet].

Available from: https://www.uptodate.com/contents/approach-to-the-adult-with-

unexplained-

thrombocytopenia?search=thrombocytopenia&source=search_result&selectedTitle

=1~150&usage_type=default&display_rank=1#H20171687

25. Clark H. The history of hemostasis. 1929; Available from:

https://scholar.google.com/scholar_lookup?hl=en&volume=37&publication_year=

1929&journal=Semin+Hematol.&author=H.+S.+Clark&title=The+History+of+He

mostasis

26. Morawitz P. Die Chemie der Blutgerinnung. Springer-Verlag; 1905 Dec;4(1):307–

422. Available from: http://link.springer.com/10.1007/BF02321003

27. Boulton F. A hundred years of cascading - started by Paul Morawitz (1879-1936),

a pioneer of haemostasis and of transfusion. Wiley/Blackwell (10.1111); 2006 Feb

1;16(1):1–10. Available from: http://doi.wiley.com/10.1111/j.1365-

3148.2006.00643.x

28. MACFARLANE RG. An Enzyme Cascade in the Blood Clotting Mechanism and

its Function as a Biochemical Amplifier. Nature Publishing Group; 1964 May

16 2;202(4931):498–499. Available from:

http://www.nature.com/doifinder/10.1038/202498a0

29. Davie EW, Ratnoff OD. Waterfall Sequence for Intrinsic Blood Clotting. 1964 Sep

18;145(3638):1310–1312. Available from:

http://www.sciencemag.org/cgi/doi/10.1126/science.145.3638.1310

30. Ruggeri ZM, Mendolicchio GL. Adhesion Mechanisms in Platelet Function. 2007

Jun 22;100(12):1673–1685. Available from:

http://circres.ahajournals.org/cgi/doi/10.1161/01.RES.0000267878.97021.ab

31. Varga-Szabo D, Pleines I, Nieswandt B. Cell Adhesion Mechanisms in Platelets.

2008 Mar 1;28(3):403–412. Available from:

http://atvb.ahajournals.org/cgi/doi/10.1161/ATVBAHA.107.150474

32. Versteeg HH, Heemskerk JWM, Levi M, Reitsma PH. New Fundamentals in

Hemostasis. American Physiological Society Bethesda, MD; 2013 Jan;93(1):327–

358. Available from: http://www.physiology.org/doi/10.1152/physrev.00016.2011

33. Eyre L, Gamlin F. Haemostasis, blood platelets and coagulation. Elsevier; 2010

Jun 1;11(6):244–246. Available from:

https://www.sciencedirect.com/science/article/pii/S147202991000055X

34. Roberts H, Hoffman M, Monroe D. A Cell-Based Model of Thrombin Generation.

2006 Feb;32(S 1):032–038. Available from: http://www.thieme-

connect.de/DOI/DOI?10.1055/s-2006-939552

35. Gale AJ. Continuing Education Course #2: Current Understanding of Hemostasis.

2011 Jan 30;39(1):273–280. Available from:

http://journals.sagepub.com/doi/10.1177/0192623310389474

17 36. Hoffman M. Remodeling the Blood Coagulation Cascade. 2003 Aug;16(1/2):17–

20. Available from:

http://link.springer.com/10.1023/B:THRO.0000014588.95061.28

37. Heemskerk J, Bevers E, Lindhout T. Platelet Activation and Blood Coagulation.

Schattauer GmbH; 2002 Dec 7;88(08):186–193. Available from:

http://www.thieme-connect.de/DOI/DOI?10.1055/s-0037-1613209

38. RC Becker. Cell-based models of coagulation: a paradigm in evolution. 2005;

Available from: https://link.springer.com/content/pdf/10.1007/s11239-005-3118-

3.pdf

39. Yamaji-Hasegawa A, Tsujimoto M. Asymmetric distribution of phospholipids in

biomembranes. 2006; Available from:

https://www.jstage.jst.go.jp/article/bpb/29/8/29_8_1547/_article/-char/ja/

40. Zwaal RFA, Comfurius P, Bevers EM. Surface exposure of phosphatidylserine in

pathological cells. 2005 May;62(9):971–988. Available from:

http://link.springer.com/10.1007/s00018-005-4527-3

41. Lentz B. Exposure of platelet membrane phosphatidylserine regulates blood

coagulation. 2003; Available from:

https://www.sciencedirect.com/science/article/pii/S0163782703000250

42. Smith S, Mutch N, Baskar D-. Polyphosphate modulates blood coagulation and

fibrinolysis. 2006; Available from: http://www.pnas.org/content/103/4/903.short

43. Ruiz F, Lea C, Oldfield E, Docampo R. Human platelet dense granules contain

polyphosphate and are similar to acidocalcisomes of bacteria and unicellular

eukaryotes. 2004; Available from: http://www.jbc.org/content/279/43/44250.short

18 44. Morrissey J, Choi S, Smith S. Polyphosphate: an ancient molecule that links

platelets, coagulation and inflammation. 2012; Available from:

http://www.bloodjournal.org/content/early/2012/04/18/blood-2012-03-

306605.short

45. Morrissey JH. Polyphosphate: a link between platelets, coagulation and

inflammation. Springer Japan; 2012 Apr 5;95(4):346–352. Available from:

http://link.springer.com/10.1007/s12185-012-1054-5

46. Muszbek L, Bereczky Z, Bagoly Z, Komáromi I, Katona É. Factor XIII: A

Coagulation Factor With Multiple Plasmatic and Cellular Functions. 2011

Jul;91(3):931–972. Available from:

http://www.physiology.org/doi/10.1152/physrev.00016.2010

47. Ariëns R, Lai T, Weisel J, Greenberg C. Role of factor XIII in fibrin clot formation

and effects of genetic polymorphisms. 2002; Available from:

http://www.bloodjournal.org/content/100/3/743.short

48. Jen CJ, McIntire L V. The structural properties and contractile force of a clot.

1982;2(5):445–455. Available from: http://doi.wiley.com/10.1002/cm.970020504

49. Weisel J. The mechanical properties of fibrin for basic scientists and clinicians.

2004; Available from:

https://www.sciencedirect.com/science/article/pii/S030146220400198X

50. Lam WA, Chaudhuri O, Crow A, Webster KD, Li T-D, Kita A, Huang J, Fletcher

DA. Mechanics and contraction dynamics of single platelets and implications for

clot stiffening. Nature Publishing Group; 2011 Jan 5;10(1):61–66. Available from:

http://www.nature.com/articles/nmat2903

19 51. Chapin J, Hajjar K. Fibrinolysis and the control of blood coagulation. 2015;

Available from:

https://www.sciencedirect.com/science/article/pii/S0268960X14000721

52. Travis J, Salvesen GS. Human plasma proteinase inhibitors. 1983 Jun;52(1):655–

709. Available from:

http://www.annualreviews.org/doi/10.1146/annurev.bi.52.070183.003255 PMID:

6193754

53. Broze GJ, Higuchi D. Coagulation-dependent inhibition of fibrinolysis: role of

carboxypeptidase-U and the premature lysis of clots from hemophilic plasma.

1996;88(10). Available from:

http://www.bloodjournal.org/content/88/10/3815.short?sso-checked=true

54. Mosnier L, Lisman T, Berg M van den, Nieuwenhuis K, Meijers J, Bouma B. The

Defective Down Regulation of Fibrinolysis in A Can Be Restored by

Increasing the TAFI Plasma Concentration. Schattauer GmbH; 2001 Dec

9;86(10):1035–1039. Available from: http://www.thieme-

connect.de/DOI/DOI?10.1055/s-0037-1616530

55. Hoylaerts M, Rijken DC, Lijnen HR, Collen D. Kinetics of the activation of

plasminogen by human tissue plasminogen activator. Role of fibrin. 1982 Mar

25;257(6):2912–9. Available from: http://www.ncbi.nlm.nih.gov/pubmed/7199524

PMID: 7199524

56. Plow E, Collen D. The presence and release of alpha 2-antiplasmin from human

platelets. 1981;58(6). Available from:

http://www.bloodjournal.org/content/58/6/1069.short

20 57. Modery-Pawlowski CL, Tian LL, Pan V, McCrae KR, Mitragotri S, Sen Gupta A.

Approaches to synthetic platelet analogs. Elsevier Ltd; 2013;34(2):526–541.

Available from: http://linkinghub.elsevier.com/retrieve/pii/S0142961212010976

58. Peng T. Biomaterials for hemorrhage control. 2010;24(1):27–68.

59. di Lena F. Hemostatic polymers: the concept, state of the art and perspectives.

Royal Society of Chemistry; 2014 May 21;2(23):3567–3577. Available from:

http://xlink.rsc.org/?DOI=C3TB21739F

60. Alam HB, Burris D, DaCorta J a, Rhee P. Hemorrhage control in the battlefield:

role of new hemostatic agents. 2005;170(1):63–69. PMID: 15724857

61. Pusateri AE, Holcomb JB, Kheirabadi BS, Alam HB, Wade CE, Ryan KL. Making

Sense of the Preclinical Literature on Advanced Hemostatic Products. 2006

Mar;60(3):674–682. Available from:

https://insights.ovid.com/crossref?an=00005373-200603000-00034

62. Annabi N, Tamayol A, Shin SR, Ghaemmaghami AM, Peppas NA,

Khademhosseini A. Surgical materials: Current challenges and nano-enabled

solutions. Elsevier; 2014 Oct 1;9(5):574–589. Available from:

https://www.sciencedirect.com/science/article/pii/S1748013214001315

63. Gabay M, Boucher BA. An Essential Primer for Understanding the Role of

Topical Hemostats, Surgical Sealants, and Adhesives for Maintaining Hemostasis.

Wiley-Blackwell; 2013 Sep 1;33(9):935–955. Available from:

http://doi.wiley.com/10.1002/phar.1291

64. Blackbourne LH, Baer DG, Eastridge BJ, Kheirabadi B, Kragh JF, Cap AP,

Dubick M a., Morrison JJ, Midwinter MJ, Butler FK, Kotwal RS, Holcomb JB.

21 Military medical revolution. 2012;73:S372–S377. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201212005-00002 PMID: 23192060

65. Chandler MH, Roberts M, Sawyer M, Myers G. The US military experience with

fresh whole blood during the conflicts in Iraq and Afghanistan. 2012 Sep

27;16(3):153–9. Available from:

http://journals.sagepub.com/doi/10.1177/1089253212452344 PMID: 22927704

66. Kauvar DS, Holcomb JB, Norris GC, Hess JR. Fresh Whole Blood Transfusion: A

Controversial Military Practice. 2006 Jul;61(1):181–184. Available from:

https://insights.ovid.com/crossref?an=00005373-200607000-00027

67. Spinella PC, Perkins JG, Grathwohl KW, Repine T, Beekley AC, Sebesta J,

Jenkins D, Azarow K, Holcomb JB. Risks associated with fresh whole blood and

red blood cell transfusions in a combat support hospital. 2007 Nov;35(11):2576–

2581. Available from: https://insights.ovid.com/crossref?an=00003246-

200711000-00015

68. Spitalnik SL, Triulzi D, Devine D V., Dzik WH, Eder AF, Gernsheimer T,

Josephson CD, Kor DJ, Luban NLC, Roubinian NH, Mondoro T, Welniak LA,

Zou S, Glynn S. 2015 proceedings of the National Heart, Lung, and Blood

Institute’s State of the Science in Transfusion symposium.

2015;55(9):2282–2290. PMID: 26260861

69. Kaufman RM, Djulbegovic B, Gernsheimer T, Kleinman S, Tinmouth AT,

Capocelli KE, Cipolle MD, Cohn CS, Fung MK, Grossman BJ, Mintz PD,

O’Malley BA, Sesok-Pizzini DA, Shander A, Stack GE, Webert KE, Weinstein R,

22 Welch BG, Whitman GJ, Wong EC, Tobian AAR. Platelet Transfusion: A Clinical

Practice Guideline From the AABB. American College of ; 2015 Feb

3;162(3):205. Available from: http://annals.org/article.aspx?doi=10.7326/M14-

1589

70. Etchill EW, Myers SP, Raval JS, Hassoune A, SenGupta A, Neal MD. Platelet

Transfusion in Critical Care and . 2017 May;47(5):537–549. Available

from: http://insights.ovid.com/crossref?an=00024382-201705000-00002

71. Roberts HR. Recombinant factor VIIa (NovoSeven®) and the safety of treatment.

W.B. Saunders; 2001 Oct 1;38:48–50. Available from:

https://www.sciencedirect.com/science/article/pii/S0037196301901489

72. Berrettini M, Mariani G, Schiavoni M, Rocino A, Di Paolantonio T, Longo G,

Morfini M. Pharmacokinetic evaluation of recombinant, activated factor VII in

patients with inherited factor VII deficiency. Haematologica; 2001 Jun

1;86(6):640–5. Available from: http://www.ncbi.nlm.nih.gov/pubmed/11418374

PMID: 11418374

73. Fries D, Martini WZ. Role of fibrinogen in trauma-induced coagulopathy. Oxford

University Press; 2010 Aug 1;105(2):116–121. Available from:

http://linkinghub.elsevier.com/retrieve/pii/S0007091217335298

74. RAHE-MEYER N, SØRENSEN B. Fibrinogen concentrate for management of

bleeding. Wiley/Blackwell (10.1111); 2011 Jan 1;9(1):1–5. Available from:

http://doi.wiley.com/10.1111/j.1538-7836.2010.04099.x

75. Levy JH, Goodnough LT. How I use fibrinogen replacement in acquired

bleeding. American Society of Hematology; 2015 Feb 26;125(9):1387–93.

23 Available from: http://www.ncbi.nlm.nih.gov/pubmed/25519751 PMID: 25519751

76. Sperry JL, Guyette FX, Brown JB, Yazer MH, Triulzi DJ, Early-Young BJ, Adams

PW, Daley BJ, Miller RS, Harbrecht BG, Claridge JA, Phelan HA, Witham WR,

Putnam AT, Duane TM, Alarcon LH, Callaway CW, Zuckerbraun BS, Neal MD,

Rosengart MR, Forsythe RM, Billiar TR, Yealy DM, Peitzman AB, Zenati MS.

Prehospital Plasma during Air Medical Transport in Trauma Patients at Risk for

Hemorrhagic Shock. 2018;379(4):315–326. Available from:

http://www.nejm.org/doi/10.1056/NEJMoa1802345 PMID: 30044935

77. Cardenas JC, Zhang X, Fox EE, Cotton BA, Hess JR, Schreiber MA, Wade CE,

Holcomb JB, PROPPR Study Group. Platelet transfusions improve hemostasis and

survival in a substudy of the prospective, randomized PROPPR trial. American

Society of Hematology; 2018 Jul 24;2(14):1696–1704. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/30030268 PMID: 30030268

78. Corash L. Inactivation of viruses, bacteria, protozoa, and leukocytes in platelet

concentrates: Current research perspectives. W.B. Saunders; 1999 Jan 1;13(1):18–

30. Available from:

https://www.sciencedirect.com/science/article/pii/S0887796399800856

79. Heal JM, Blumberg N. Optimizing platelet transfusion therapy. Churchill

Livingstone; 2004 Sep 1;18(3):149–165. Available from:

https://www.sciencedirect.com/science/article/pii/S0268960X03000572

80. Mohanty D. Current concepts in platelet transfusion. Wolters Kluwer -- Medknow

Publications; 2009 Jan;3(1):18–21. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/20041092 PMID: 20041092

24 81. Lee DH, Blajchman MA. Novel treatment modalities: New platelet preparations

and subsititutes. 2001;114(3):496–505. PMID: 11552973

82. Blajchman MA. Novel platelet products, substitutes and alternatives.

2001;8(3):267–271. PMID: 11499975

83. Epstein JS, Vostal JG. FDA approach to evaluation of pathogen reduction

technology. Wiley/Blackwell (10.1111); 2003 Oct 1;43(10):1347–1350. Available

from: http://doi.wiley.com/10.1046/j.1537-2995.2003.00584.x

84. Galan AM, Lozano M, Molina P, Navalon F, Marschner S, Goodrich R, Escolar G.

Impact of pathogen reduction technology and storage in platelet additive solutions

on platelet function. Wiley/Blackwell (10.1111); 2011 Apr 1;51(4):808–815.

Available from: http://doi.wiley.com/10.1111/j.1537-2995.2010.02914.x

85. Lambert MP, Sullivan SK, Fuentes R, French DL, Poncz M, Dc W. Challenges

and promises for the development of donor-independent platelet transfusions

Review Article Challenges and promises for the development of donor-

independent platelet transfusions. 2013;121(17):3319–3324.

86. Solheim BG. Pathogen reduction of blood components. Pergamon; 2008 Aug

1;39(1):75–82. Available from:

https://www.sciencedirect.com/science/article/pii/S1473050208000888

87. Seghatchian J, de Sousa G. Pathogen-reduction systems for blood components:

The current position and future trends. Pergamon; 2006 Dec 1;35(3):189–196.

Available from:

https://www.sciencedirect.com/science/article/pii/S1473050206001364

88. Janetzko K, Hinz K, Marschner S, Goodrich R, Klüter H. Pathogen reduction

25 technology (Mirasol®) treated single-donor platelets resuspended in a mixture of

autologous plasma and PAS. Wiley/Blackwell (10.1111); 2009 Oct 1;97(3):234–

239. Available from: http://doi.wiley.com/10.1111/j.1423-0410.2009.01193.x

89. Pidcoke HF, McFaul SJ, Ramasubramanian AK, Parida BK, Mora AG, Fedyk CG,

Valdez-Delgado KK, Montgomery RK, Reddoch KM, Rodriguez AC, Aden JK,

Jones JA, Bryant RS, Scherer MR, Reddy HL, Goodrich RP, Cap AP. Primary

hemostatic capacity of whole blood: a comprehensive analysis of pathogen

reduction and refrigeration effects over time. Wiley/Blackwell (10.1111); 2013

Jan;53:137S–149S. Available from: http://doi.wiley.com/10.1111/trf.12048

90. Avanzi MP, Mitchell WB. Ex Vivo production of platelets from stem cells.

Wiley/Blackwell (10.1111); 2014 Apr 1;165(2):237–247. Available from:

http://doi.wiley.com/10.1111/bjh.12764

91. Nakamura S, Takayama N, Hirata S, Seo H, Endo H, Ochi K, Fujita K, Koike T,

Harimoto K, Dohda T, Watanabe A, Okita K, Takahashi N, Sawaguchi A,

Yamanaka S, Nakauchi H, Nishimura S, Eto K. Expandable Megakaryocyte Cell

Lines Enable Clinically Applicable Generation of Platelets from Human Induced

Pluripotent Stem Cells. Cell Press; 2014 Apr 3;14(4):535–548. Available from:

https://www.sciencedirect.com/science/article/pii/S1934590914000125?via%3Dih

ub

92. Thon JN, Mazutis L, Wu S, Sylman JL, Ehrlicher A, Machlus KR, Feng Q, Lu S,

Lanza R, Neeves KB, Weitz DA, Italiano JE. Platelet bioreactor-on-a-chip.

American Society of Hematology; 2014 Jul 21;blood-2014-05-574913. Available

from: http://www.bloodjournal.org/content/early/2014/07/18/blood-2014-05-

26 574913.short?sso-checked=true PMID: 25049277

27 Chapter 2: Externally Administered (Topical and Intracavitary) Biomaterials and

Advanced Technologies for Management of Traumatic Bleeding

Content based on: Hickman DA, Pawlowski CL, Sekhon UDS, Marks J, Gupta AS. Biomaterials and Advanced Technologies for Hemostatic Management of Bleeding. Adv Mater. 2017 Nov 22. doi: 10.1002/adma.201700859.

2.1 Introduction

Externally administered hemostatic treatment of an injury or lesion is applicable to simple

cuts and , exposed and puncture wounds, surgical lacerations, and heterogenous

. These scenarios result in various degrees of bleeding and tissue damage, and

hence the development of appropriate hemostatic materials and technologies is driven by

the need to quickly staunch bleeding, absorb shed blood, cover and protect the injury,

prevent contamination and provide a suitable environment for healing. These materials and

strategies include topical use of tourniquets, dressings, bandages, foams, powders and gels,

as well as, administering materials in an intracavitary fashion. A wide variety of materials

and technologies have been developed, pre-clinically evaluated and clinically translated in this area as discussed in the following sections.

2.2 Tourniquets

As described previously, a blood vessel’s first natural response to injury is spasm and construction, leading to reduction in blood flow (stypsis) and facilitating the subsequent cellular and biomolecular clotting mechanisms. Therefore, a number of hemostatic materials and technologies have focused on augmenting this process by collapsing blood vessels with application of pressure to allow platelet activity, coagulation cascade and

28 formation of a fibrin clot. Beyond applying manual pressure with fingers and palm, a widely used technology in this category is that of tourniquets. A tourniquet is essentially a circumferentially constrictive bandage that can restrict blood supply to extremities. In ancient Roman history, Galen of Pergamon (129 - 200 AD) was known to use this method for stopping blood flow and was criticized for its use by those who feared that tourniquets would increase blood loss from a . In the early 16th century, the famous Prussian military Hans von Gersdorff had described tourniquet use in surgery1. Towards the end of the 16th century, Wilhelm Fabry of German surgical science

fame, reportedly also used tourniquets in amputation surgery, utilizing mechanical

tightening2. Historically in such uses, tourniquets have been reported to present certain

negative consequences, especially in the context of denying blood supply to extremities

hereby causing ischemia and infarction as well as causing mechanical nerve damage.

Despite this criticism, over time tourniquets have become an adapted and accepted

hemostatic technology in pre-hospital management of bleeding in the military (i.e. in combat wounds), as emphasized by documents of the Tactical Combat Casualty Care

(TCCC) of the US military and reviewed recently by Lakstein et al and Kragh et al3,4.

Before being adapted as a tactical mainstay, most tourniquets were made of an improvised

stick with a cloth or a silicone elastic strip. However, due to material inferiority and risks

of slippage (defective fastening), in the mid 2000s significant R&D efforts were dedicated

to tourniquet design and standard-of-usage. The first result was a technology named

Combat Application Tourniquet (C-A-T®, North American Rescue, USA), which is essentially a composite improvisation of the ‘cloth with turning stick for tightening’ approach (aka Spanish windlass design) that can be used with minimal effort (e.g. with one

29 hand)5. Continued studies on tourniquet design, ease of usage in complex time-limiting

environment, patient comfort, hemostatic efficacy and reduction of tissue morbidity and

mortality have led to other tourniquet technologies such as Emergency Medical Tourniquet

(EMT, Delfi Medical Innovations, Canada), Special Operations Forces Tactical Tourniquet

(SOF-TT), Ratcheting Medical Tourniquet (RMT, M2 Inc, USA) etc. The EMT tourniquet

is formed of circumferentially usable bladder that can go around the limb, a clamp that

limits the inflated portion while holding the bladder close to the limb, and an inflator bulb

equipped with a connector tube and twist cap. Hence, the design is similar to a blood

pressure cuff except for the clamp component. The SOF-TT tourniquet is similar in design

principle to the C-A-T system (i.e. the Spanish windlass design) with an aluminum-based

stick. The RMT system requires use of a ratcheting lever instead of a windlass stick for

tightening the tourniquet. The usefulness of tourniquets to reduce blood loss (and hence

morbidity/mortality) by simple yet efficient application of styptic pressure is responsible

for its current consideration as a standard-of-care component for emergency responders in

civilian trauma6,7.

2.3 Naturally derived biomaterials for hemostatic applications

The environment has been a dependable source for a wide variety of materials like cotton, collagen, gelatin, silk, elastin, fibrin etc., that have found extensive applications in biomedical areas of device coatings, tissue adhesives and sutures, tissue repair and regeneration, cell encapsulation, drug and gene delivery etc.8. A variety of naturally

derived materials have also found significant applications in the area of hemostasis and

bleeding management. Some of these materials have only absorption and passive

30 interaction properties, while others have active biointeractions to promote hemostatic

mechanisms. Absorptive and passively interactive materials impart partial hemostasis merely by wound site coverage, absorption of blood and exudates, and subsequent protection, but do not contain any specific component that promote or augment hemostasis or biologically protect from bacterial infection. Bioactive materials and dressings are systems that adhere to the bleeding tissue to either render hemostasis-stimulating properties by themselves or by virtue of components embedded in them that facilitate hemostatic mechanisms and prevent infection. The following sections will provide a descriptive review of absorptive, passively interactive and bioactive materials derived from natural sources that have various hemostatic applications, and at the end Table 2-1 will provide an

‘at-a-glance’ summary of naturally derived hemostatic materials along with representative technology names, characteristic evaluations and current application status.

2.3.1 Absorptive and passively interactive materials in hemostatic technologies

While the function of a tourniquet is to augment the vasoconstriction and stypsis at the bleeding site, the function of a bandage or dressing is to directly cover and protect the site from further contamination and tissue damage, allow exchange of gas and fluids (blood and exudate absorption and removal), potentially prevent infection and allow healing in the long term. To this end, sterile wound dressings and absorbent pads have been made of various materials including cotton gauze or cotton pad, oxidized cellulose material, nylon/rayon/polyester variants, tulle gauze, semi-permeable porous polymer membranes and foams, hydrocolloidal and hydrofiber materials, and amorphous hydrogels. Cotton is a cellulosic polymer that differs from wood cellulose by the fact that it has a much higher

31 degree of polymerization and crystallinity than wood cellulose. Cellulose is a

homopolysaccharide of glucopyranose, polymerized through β-glucosidic bonds9,10.

Cotton contains about 90% of this cellulosic polymer along with a small amount of

hydrophobic waxes and pectin, while wood contains about 40-50% cellulosic polymer. The

US military standard field dressing consists of two layers of gauze wrapped over densely

packed cotton, such that it can absorb a large volume of blood while the cotton strands are

thought to trigger platelet activation and aggregation due to high hydrophilicity, negative

surface charge and surface energy11. The cotton gauze or pad, when directly placed over a

bleeding injury, may provide high blood absorption, partial hemostatic reaction (thrombin

generation via contact pathway) and additional tamponade effect with adjunctive

compression components, but they can adhere to the wound tissue which may be an issue

during their subsequent removal. Nonetheless, in recent years interesting research is being

focused on modulating the hemostatic characteristics of cotton dressings and gauze by

modifying them with other hemostatic or strengthening materials like kaolin mineral,

chitosan, viscose, rayon etc.11,12. Oxidized cellulose (OC) derived from cotton and

oxidized regenerated cellulose (ORC) derived usually from wood pulp, refer to

manipulation of the cellulose structure where primary and secondary alcohol moieties are

oxidatively converted to aldehyde, ketone or carboxyl groups, which significantly changes

the physico-chemical and mechanical properties of OC and ORC compared to native

cellulose13–17. The use of OC and ORC as hemostatic dressing materials in surgery was reported around WWII18–20, and since then its widespread application as a surgical wound

dressing material has established its biocompatibility, bactericidal and hemostatic

properties10,21–26. The acidic pH of these oxidized cellulose materials and the negative

32 charge are thought to impart platelet activation, aggregation and intrinsic pathway of

coagulation mechanisms. These materials are also reported to be biodegradable via

enzymatic (glycosidase-based) and macrophagic processes27,28. Recent advancement of

these materials involves modification of the material matrix with other hemostatic agents

like fibrin29. Like cotton gauze and pads, the OC and ORC materials can present the issue of adherence to the bleeding tissue, that may pose some logistical difficulties during their subsequent removal.

The adherence issue can be resolved by making materials non-adherent yet absorbent in dressings, e.g., paraffin-soaked tulle gauze and various polymer-based, hydrogel-based, hydrofiber-based and hydrocolloid-based dressings. Tulle dressings (Tulle Gras in French, meaning ‘oily tulle’) are essentially open weave cloth gauze soaked in paraffin (petroleum jelly), balsam and olive oil to impart hydrophobicity and non-adherence to bleeding wound site30. These materials are usually used for low exudate wounds and can be further modified

by impregnating with anti-septic agents like chlorhexidine gluconate to impart infection-

protection. The cotton pad, gauze and tulle materials are traditionally considered as

‘passive wound dressing’ materials since their primary function is to fully cover the wound

for physical and mechanical protection, while allowing exudate absorption and fluid exchange. For high exudate situation, cellulose-based hydrogels have also been developed.

For example, amorphous hydrogel materials consisting of mostly water with about 2-3%

33

Figure 2-1 Representative chemical structures of cotton (cellulose) biopolymers and its derivatives that have undergone extensive research in the development of hemostatic technologies like gauze and wound dressings.

of a gel-forming polymer such as sodium carboxymethylcellulose, modified starch or

sodium alginate, along with about 15-20% propylene glycol (a humectant and

preservative), have been developed for wound dressing and hemostatic applications31,32.

These dressings, along with synthetic polymer-based hydrocolloid and hydrofiber systems

(discussed later in this review), are highly suitable for wounds and necrotic or sloughy wound beds, since the water content can cool down the injury site to aid in comfort and

34 healing. The gels are often used with a secondary dressing material like perforated or gas-

permeable plastic film that prevents the water content of the hydrogel from evaporating

outward but rather donated towards the wound. Similar hydrated water-donating property

can be achieved with corboxymethylcellulose-based hydrogelic particles dispersed within

polyurethane film or foam (for hydrocolloids) or carboxymethylcellulose-based fibers

manufactured into non-woven pads or ribbons33. Both of these systems are useful for

hemostatic action, protection and healing maintenance of heavily exudating deep wounds.

Figure 2-1 shows representative chemical structures of common cellulose-based materials

that are used in hemostatic wound-dressing applications. It is important to note here that

although cotton and oxidized cellulose system are traditionally included in the ‘passive

dressing’ group, they can be argued to possess a degree of bioactivity due to their ability

to potentially stimulate primary (platelet activation and/or aggregation) and secondary

(contact activation of intrinsic coagulation pathway) hemostatic mechanism components.

However, a systematic study of the hemostatic mechanisms triggered by these materials is

yet to be reported, and most studies report performance output in terms of extent of

hemostasis, blood loss and tissue morbidity evaluation.

2.3.2 Bioactive materials in hemostatic technologies

Thrombin, Fibrinogen and Fibrin: Thrombin, fibrinogen and fibrin (along with active platelets), are the critical components for forming the hemostatic clot, as shown in Figure

2-2. Consequently, fibrin which is the protein formed as the end product of the coagulation cascade via reaction of thrombin on fibrinogen, has become an important bioactive material to be used in hemostatic applications. Reports from the early 20th century suggest fibrin to

35 be superior to cotton in terms of hemostatic capacity34–36, and since then fibrin dressings

and sealants have become one of the most-studied hemostatic material, especially in the

surgery field37–41. In early reports by Grey and Harvey, fibrin was used as pre-polymerized

material processed into tamponade and plaque-like devices for treatment of bleeding in

parenchymal organs34,35. These products demonstrated hemostatic capacity but had

reduced capacities of tissue-integration since the fibrin was already pre-polymerized.

Figure 2-2 Schematic of concomitant roles of thrombin and fibrin(ogen) in propagating the formation of hemostatic clots via activation of platelets and formation of fibrin mesh.

Around WWII, Cronkite et al and Tedrick et al reported the usage of fibrinogen with

thrombin to produce fibrin in situ in relevance to skin transplant procedures42,43, which has

led to the modern-day usage of this precursor mixture for fibrin-based hemostatic products.

Fibrin can be used in dry condition where animal- or human-sourced thrombin and

fibrinogen are freeze-dried, processed into powder, foam, fleece, etc. and impregnated into a secondary bandage or carrier dispersant system to be applied directly onto the bleeding site41,44–49. The secondary dressing can be a passive polymeric strip like silicone with an

absorbent vicryl mesh backing or another bioactive but mechanically more robust material

like collagen sheet50–52. The materials design rationale for all these systems is that upon

contact with an actively bleeding site, the fibrinogen and thrombin in the dressing will

36 interact to form fibrin in situ, leading to a hemostatic effect. In all evaluations so far, these

fibrin(-generating) dressings have shown superior hemostatic performance compared to

passive dressings, possibly because of their pro-coagulant bioactivity49,53–58. Fibrin

adhesives can also be used topically in a liquid form, where the freeze-dried fibrinogen and thrombin components are reconstituted in sterile saline immediately before administration, often through a specially manufactured dual-barrel injection device37,38,59,60. These liquid sealants can have high degree of tissue adherence due to physico-chemical (electrostatic,

hydrogen and covalent bonding) interactions as well as mechanical integration into the

tissue and can be applied to heterogenous injury sites due to their form-filling nature. The dry fibrin dressings and the liquid fibrin adhesive sealants have been reported to contain varying degrees of fibrinogen and thrombin, depending upon products from various companies and laboratories61,62. These compositionally different products have been tested for assessing variations in hemostatic performance, tensile strength, tissue adhesiveness etc. In some cases, no significant difference was found in the output performances, while in other cases changing the fibrinogen concentration showed some effect in mechanical

strength and tissue adherence63–66. The fibrin-forming precursor systems (dry or liquid) can

also contain other pro-coagulant materials like Factor XIII (a fibrin cross-linking

transglutaminase enzyme) as well as anti-fibrinolytic agents like and tranexamic

acid (TXA, discussed in detail later in the next chapter) to modulate clot formation speed

and clot strength/stability67–69. Although such modifications may provide marginal benefits, the significance of such benefits have not been systematically evaluated and statistically established.

37 The lingering issue for a long time with such fibrinogen and thrombin-based products has

been the risk of immunogenicity and viral contamination, since the components are sourced

from animal (bovine, porcine) or human pooled blood. Bovine thrombin preparations have

been implicated in immunogenic reactions and increased risk of adverse clinical outcomes

following its use in surgical procedures70. Fibrin based products that were mass-produced around 1944-45 to meet the needs during WWII were withdrawn in 1946 because of reports of hepatitis transmission71. As a progression of that, in the late 1970s many plasma-sourced fibrinogen-based products, originally clinically approved for hemostatic and surgical procedures, were recalled by the FDA. In the last decade, due to the emergence and establishment of rigorous blood screening, serological testing and pathogen (bacteria, virus) reduction/inactivation technologies, plasma-sourced products including fibrin systems have undergone a revival36,41,71. Through rigorous research conducted at several

laboratories, including studies led by the US Army Institute of Surgical Research

(USAISR), products like TachoComb and TachoSil (Nycomed, Austria) and Dry Fibrin

Sealant Dressing (DFSD, developed by American Red Cross with USAISR) have become

an important component in the current state-of-art toolbox for hemostatic management of traumatic and surgical bleeding36,41,49. Fibrin’s natural spatio-temporally regulated

characteristics of biodegradation and are also responsible for its continued

popularity in the milieu of hemostatic materials72. In recent years, there have also been

reports on sourcing the fibrinogen and thrombin from salmon73–77 as well as developing recombinant versions of such coagulation proteins78–81. For example, RecothromTM

(ZymoGenetic Inc, USA) is a fully recombinant human thrombin that has been clinically

approved as a topical hemostatic agent to treat oozing blood and capillary bleeding and can

38 be used in conjunction with other wound dressings. A transgenic approach has also been

utilized to develop proteins such as fibrinogen in the milk of other mammals for potential pharmacotherapeutic use82. Recombinant and transgenic versions of thrombin and

fibrinogen are under pre-clinical and clinical evaluation, and can potentially resolve the

availability and immunogenicity issues that are otherwise associated with human or animal

plasma-resourced products. As for liquid fibrin sealants and adhesive products, although

they remain a relevant material in surgery, their widespread pre-hospital use (e.g. in the

battlefield) has been somewhat restricted due to time-consuming rehydration and mixing process of the lyophilized powders for in situ delivery. Furthermore, the resultant fibrin is capable of hemostasis in low volume and pressure bleeding scenarios but not heavy traumatic bleeding. In recent years, autologous fibrin generation technologies like

CryoSeal (Asahi Kasei Medical Co., Tokyo, Japan) and Vivostat (Vivostat A/S, Denmark)

have been reported that utilize small volumes of patient’s own plasma to generate fibrin

sealant in the operating room83. These sealants have shown significant clinical promise as

hemostatic materials during spine and sternum surgeries84–87. Fibrin has also been

developed into ‘foam’ technologies for spray-based hemostatic use as sealants for heavy parenchymal bleeding88–90. Fibrin sealants have also been combined with other hemostatic

materials like oxidized cellulose-based Surgicel® (Ethicon, USA) to cumulatively enhance

hemostatic capability in surgical applications91. Due to their potential for externally

injectable and space-filling properties, the liquid versions of fibrin sealant products may

find use in intracavitary hemostatic applications.

39 Collagen and Gelatin: Another important bioactive material in hemostatic applications is collagen and its denatured variant, gelatin (multiscale schematic structure shown in Figure

2-3). Collagen is the most abundant structural protein found in the extracellular matrices

of many connective tissues in mammals, making up about 25-35% of the whole-body

protein content92,93. Fibrous collagen is present in the sub-endothelial matrix (Type III

collagen mainly, with Type IV in the basement membrane) and upon injury (including

endothelial disruption and denudation), collagen is exposed to circulating blood

components. Von Willebrand Factor secreted from injured endothelium and activated

platelets can bind and self-associate on exposed collagen. Platelets can bind to self-

associated vWF via platelet surface GPIbα receptor protein, and can also bind to collagen

directly via platelet surface GPIa/IIa and GPVI receptor proteins94–96. These adhesion

mechanisms of platelets at the injury site lead to further activation of localized platelets,

ultimately leading to active platelet aggregation at the site (primary hemostasis) as well as

Figure 2-3 Multiscale schematic representation of fibrillar collagen structure where triple-helical microfibrils formed of Gly-X-Y amino acid repeat units assemble in staggered orientation to form collagen fibrils, which in turn further assemble to form high molecular weight collagen fibers and bundles; denaturation of collagen disrupts this assembled structures to form gelatin, which can be partly reassembled into helical components to form lower molecular weight gels.

40 augmentation of coagulation cascade events on the active platelet membrane (secondary

hemostasis)97,98. The direct effect of collagen on coagulation factor localization and activation has also been reported, e.g. for FXII and FIX99. These physiological mechanisms have inspired the utilization of bio-derived collagen as a topical hemostatic material. In

1960s and 1970s, Hait et al had reported on the capability of isolated bovine collagen to adhere effectively to bleeding surfaces and promote hemostasis100–102. Such properties have

eventually led to development of a wide variety of collagen-based products in sheet,

powder, foam and fiber forms to be utilized as a topical hemostatic material, especially in

surgical applications103–108. Representative commercially available technologies that utilize

such collagen forms are AviteneTM (Davol Inc, USA), HelistatTM (Integra LifeSciences,

USA), InstatTM (Ethicon, J & J, USA) etc. Collagen-based materials and technologies have also been studied for liquid form sealants. For example, a liquid composite spray consisting of microfibrillar bovine collagen, bovine thrombin and autologous plasma named

CoStasis® (Cohesion Technologies, USA) was reported to render efficient hemostasis

when externally administered in pre-clinical animal bleeding models (e.g. liver, spleen and

kidney bleeding) and was subsequently clinically approved for surgical applications109–112.

As with animal-sourced fibrin, animal-sourced collagen can pose immunogenic risks, and

therefore research into reducing immunogenicity and infectivity has led to the popularity

of a low-immunogenic collagen-derived material called gelatin113–117. Gelatin can be

formed by thermal denaturation or irreversible hydrolysis of collagen and is extensively

used in the food industry118. Gelatin was also been found to retain hemostatic properties

like collagen119,120. Gelatin has been mixed into fibrin dressings to enhance the mechanical stability of the material72. Gelatin-based solid materials in spongy and powder form (e.g.

41 GelFoam®, Pfizer, USA) have been evaluated for hemostatic dressings in surgical procedures but have shown limited efficacy in controlling severe bleeding121–123. Gelatin

has also been evaluated as a material component in liquid hemostatic sealants, e.g. in

products like Floseal® (Baxter, USA) where it was combined with thrombin to form a composite hemostatic sealant matrix124. In these composite gelatin-based sealants, the

gelatin component is made up of collagen-derived gelatin cross-linked by glutaraldehyde

and ground into macroscopic particles, which is mixed with bovine thrombin in a special

syringe for intra-operative administration125–128. Another gelatin-based liquid sealant

material has been reported as gelatin-resorcin-formalin glue (GRF glue), where gelatin, resorcinol, formaldehyde and glutaraldehyde are mixed in an aqueous medium for application to a bleeding site such that the aldehydes promote reaction and integration with tissue, gelatin promotes hemostatic mechanisms and resorcinol provides bacteriostatic action129–131. This material has been reported in extensive use in acute aortic dissection procedures but has also raised issues of formaldehyde-associated toxicity. Similar to liquid fibrin-based sealants, collagen-based and gelatin-based liquid sealants have also been evaluated for heavy hemorrhage treatment and these may also find use in intracavitary hemostasis applications. In recent years, there is also growing interest in recombinant collagen and gelatin materials that may potentially find application in future hemostatic technologies that pose reduced immunogenic risks106,132,133.

Polysaccharide derivatives: Aside from materials and technologies based on fibrin

(fibrinogen + thrombin), collagen and gelatin, a third prominent category of bioactive

materials in hemostatic applications is that of natural polysaccharides in native or modified

42 forms134. One prominent material in this category is alginate, which is a linear

polysaccharide derived from brown algae and is comprised of blocks of (1,4)-linked β-D- mannuronate (M) and α-L-guluronate (G) residues where the blocks may be consecutive

G (i.e. GGGGGG), consecutive M (i.e. MMMMMM) or alternating G and M residues (i.e.

GMGMGM)135,136. The negatively charged uronic acid chains of alginate can gel in

presence of a cation such as calcium (Ca++), and this sequestration of Ca++ is thought to be

responsible for the hemostatic property of these gels since Ca++ is a co-factor for platelet

activation as well as several coagulation cascade reactions137,138. In pre-clinical animal

model studies as well as in clinical investigation, alginate-based dressings have shown

superior hemostatic performance compared to traditional gauze in low-to-moderate bleeding scenarios139–141. The hydrogel state of alginate dressings is suitable to keep the

wound-bed moist for healing and provide comfort during dressing changes. Besides the

traditional hydrogel form, the alginate material can also be made into micro/nano particles

as well as micro/nano fiber using suitable processing techniques, and these solid forms

have been recently evaluated in hemostatic applications. Alginate microspheres loaded

with the anti-fibrinolytic agent tranexamic acid or the pro-coagulant agent thrombin, have

shown promising hemostatic capabilities in vitro and in vivo in pre-clinical models142,143.

Currently a large number of alginate-based wound dressings are clinically approved for

surgical and hemostatic applications, e.g. Algosteril® (Johnson & Johnson, USA) and

KALTOSTAT® (Conva Tec, UK). Alginate material has also been combined with other

hemostatic materials (collagen, gelatin, oxidized cellulose, other polysachharides etc.) for

the development of composite hemostat technologies that may provide enhanced treatment

capabilities compared to mono-component systems. For example, highly absorbent

43 collagen-alginate and gelatin-alginate dressings have shown superior wound treatment

capabilities144–146. Alginate fibers have also been combined or co-spun with other polysachharides (e.g. chitosan) and natural polymers (e.g. gelatin), to enhance hemostatic and anti-bacterial effects136,147.

Along with alginate, chitosan itself has attained prominence in the area of hemostatic dressings and technologies. Chitin is a hard nitrogenous polysaccharide composed of β

(1,4)-linked 2-acetamido-2-deoxy-β-D-glucose (N-acetylglucosamine), and it is the second most ubiquitous natural polysaccharide on earth (after cellulose), usually found in the exoskeleton as well as in the internal structure of invertebrates (crustaceans, shellfish etc.)148. Enzymatic or alkaline de-acetylation of chitin results in the linear polysaccharide

chitosan, which can essentially remain as a co-polymer of N-acetylglucosamine and glucosamine depending on the degree of de-acetylation. The ability of chitosan to facilitate coagulation was reported first in the early 1980s by Malette et al149, and this has led to

extensive investigation of chitosan in various forms for hemostatic materials and technologies. The hemostatic ability of positively charged chitosan is thought to stem from

its electrostatic interaction with negatively charged cell membranes of RBCs, resulting in

RBC agglutination as a ‘physical’ mechanism of hemostatic plug formation150,151. Chitosan has also been reported to enhance adhesion, activation and aggregation of platelets (i.e. enhancement of primary hemostatic mechanisms) as well as to be able to adsorb fibrinogen from plasma, as well as, to trigger complement activation152–156. The platelet stimulating

effect of chitosan has been attributed to Ca++ mobilization. Chitosan is an acidic polyelectrolyte where about 50% de-acetylation of chitin results in a chitosan system with

44 pKa of ~7.5 (i.e. soluble in water). Therefore, modulating the de-acetylation degree

provides a control to modulate chitosan physico-mechanical as well as chemical properties.

Chitosan can be made into films, fibers, hydrogels, lyophilized particulates and solutions, and all of these forms have demonstrated hemostatic capabilities in various studies. For example, chitosan-based layered materials with interconnected open porous structures and high specific surface area were used to develop a highly effective hemostatic dressing named HemCon® (HemCon Medical Technologies, Oregon, USA) that was extensively evaluated by USAISR investigators for management of traumatic hemorrhage, resulting in the technology being added to the metric of hemostatic strategies for US military71,157,158.

This material however is rigid and inflexible, which may make it difficult to be applied

over complex injuries. In further progress of this technology, the material has shown great

clinical success in both military and civilian populations and a flexible version

(Chitoflex®, HemCon Medical Technologies) has been developed for further

evaluation159–161. Particulate (granulated, powder) forms of chitosan have also been utilized

in developing hemostatic technologies. A prominent example is a technology named

Celox® (MedTrade Products Ltd, UK), which utilizes chitosan granules and flakes to

provide high contact surface area for interact with blood162. Upon direct administration at

the wound site, blood interaction with this material is known to result in swelling of the

granules for hydrogelic absorptive effect, and the chitosan contact promotes multiple

mechanistic components of hemostasis. This material has been extensively evaluated for

hemostatic management of heavy bleeding wounds (liver blunt trauma, arterial puncture

bleeding, groin laceration etc.), showing highly effective hemostasis and reduced

occurrence of re-bleeding163–166. Along with HemCon®, Celox® is now a prominent

45 component of hemostatic management in pre-hospital and hospital scenarios. Chitosan has also been utilized to make hydrogel systems in situ by reacting thiol-modified chitosan with maleimide-modified ε-polylysine, and these materials have shown promising hemostatic capabilities added with tissue-adhesive properties of polylysine167. There are

also several recent reports on developing foams with chitosan or mixture of chitosan with

other materials (e.g. gelatin etc.), for potential application in hemostatic technologies168–

170. Chitosan gauze has also been recently reported to be coated with N-(2-

Hydroxypropyl)methacrylamide (HPMA) based synthetic fibrin-strengthening polymer

(e.g. PolySTAT) for enhanced hemostatic action151. The PolySTAT material itself will be

discussed in the context of fibrin-strengthening materials, later in Chapter 3. A chitosan

analog, poly(-N-acetyl glucosamine), produced by a fermentation process and isolated

from controlled, aseptic, microalgal cultures grown on a defined culture medium, has also

gained reputation as a hemostatic material in gel form, fiber slurry and membrane form171–

173. Thromboelastographic evaluation of this material, along with other analogous materials

like chitin, chitosan etc. demonstrated reduced clot induction time when mixed with blood.

The membrane form prepared by lyophilization and sheet-pressing of this material was

added to a backing of standard hemostatic gauze, resulting in a hemostatic technology named Rapid Deployment Gauze (RDH, Marine Polymer Technologies, USA) that was

evaluated extensively for its hemostatic capability in severe traumatic bleeding174,175. The

hemostatic efficacy shown by this technology in pre-clinical and clinical studies has resulted in its clinical approval and its adaptation as an important component in the bleeding injury management strategies in pre-hospital scenarios. Porous polysaccharide microparticles developed from processed potato starch (dextrin) have also been studied for

46 potential hemostatic applications, especially as a technology named TraumaDEX®

(Medafor Inc., USA) that can be directly applied to bleeding wounds. When evaluated in

pre-clinical bleeding models, this material and technology has demonstrated hemostatic efficacy comparable to standard gauze dressing. In similar studies, the hemostatic efficacy of TraumaDEX® was also found to have no statistical difference compared to

Celox®176,177. Altogether, bio-derived polysaccharides, like alginate, chitosan and dextrin,

form an important materials category for hemostatic applications, and a large variety of

technologies comprising of films, sheets, membranes, powders, microparticulates and

liquid forms have been developed from these materials, many of which have become

clinically approved systems for bleeding management in pre-hospital as well as hospital scenarios in military and civilian settings. Figure 2-4 shows representative chemical structures of relevant polysaccharide materials used in hemostatic dressings described above.

Minerals and zeolites: As mentioned previously, the ability of alum to facilitate hemostasis is by augmenting vasoconstriction and stypsis, which essentially is a mechanical augmentation of hemostasis. Certain microporous aluminosilicate minerals

(also called zeolites) have shown the ability for bioactive augmentation of hemostatic mechanisms. The most famous material of this category is Quikclot® (Z-Medica Inc.,

USA), a granular zeolite technology, that has been shown to render efficient hemostasis in severe hemorrhage from arteries, liver injuries and groin injuries in multiple animal models178–181. The high hemostatic efficacy of this material was also demonstrated in

clinical studies and this has resulted in its approval for use in combat casualties with severe

47

Figure 2-4 Chemical structures of some polysaccharide polymers, namely alginate, chitosan and dextrin that have been extensively used in development of hemostatic bandages and dressings.

bleeding182. The material is thought to promote rapid hemostasis via a combination of

super-absorbent property (which removes aqueous volume from blood and concentrates

clotting factors), platelet activation capacity and coagulation factor activation capability

(contact activation of intrinsic hemostatic pathway)183–185. However, the interaction of this

zeolite material with blood (aqueous medium) is substantially exothermic, resulting in

48 sudden rise in local temperature, which was found to cause tissue damage and debilitation186,187. In this aspect some research is being directed towards modulating the composition and structure of these zeolite minerals to retain its hemostatic property while reducing the exothermic side-effects188–191. In recent years, Quikclot® has been used to modify the matrix of standard gauze and sponge used in US military, resulting in a technology named Quikclot Combat Gauze (QCG) or Quikclot-Advanced Clotting Sponge

(QC-ACS). The exothermic side effect in this technology has been reduced by replacing the active ingredient zeolite with kaolin192–194. This new mineral-impregnated dressing material has shown similar hemostatic effect as the mineral itself and is currently an important component of tactical hemostatic strategies for the US military195. Another mineral based technology is WoundStat® (TraumaCare Inc., USA) that uses a smectite mineral and super-absorbent polymers196,197. This technology has shown highly efficient hemostasis in topical treatment of arterial bleeding and had gained much popularity as a field hemostat in US military198,199. However, further safety studies on the material revealed several systemic , including vascular endothelial injury, transmural damage, systemic thrombotic and embolic risks, due to microscopic residues of WoundStat remaining in the wound and blood vessels161,199. This has resulted in the military avoiding the usage of this material and replacing it with the QCG technology mentioned previously.

49

Table 2-1 Topical and Externally Administered Hemostatic Biomaterials from Natural Sources

50 51 2.4 Synthetically derived hemostatic materials

While a large volume of research has been conducted on naturally derived materials for

developing bleeding management technologies, a small number of synthetically derived

systems, especially certain polymers and peptides, have also been developed and evaluated

for hemostatic applications. One such polymer category is that of basic poly(amino acids)

like polylysine. During the 1950s DeVries et al reported that basic poly(amino acids) like

polylysine, polyarginine etc. augment generation of thrombin and fibrin and also retard

fibrinolysis166,200,201. It was postulated that the cationic nature of polylysine facilitates

complexation and activation of certain coagulation factors202. This kind of report has

possibly guided the use of polylysine as a cationic polyeletrolyte to be mixed with chitosan

for hemostatic applications, but stand-alone polylysine based technologies have not

undergone much development possibly because of biocompatibility and cytotoxicity issues

of cationic polymers203. Poly(alkylene oxides), e.g. poly (ethylene oxide) (PEO) and

poly(propylene oxide) (PPO), have been investigated as hemostatic materials since they

are already well-established biocompatible synthetic polymers in biomaterials

applications. For example, Wang et al and Wellisz et al have reported on the hemostatic

capability of a PEO-PPO-PEO block copolymer based waxy material (now marketed as

Ostene®, Baxter USA) to render hemostasis in orthopedic surgeries204,205. This material

acts much like natural bone wax (a mixture of beeswax with paraffin or petroleum jelly) in

the context of facilitating hemostasis by a tamponade mechanical effect rather than a

biochemical augmentation of coagulation mechanisms. A similar block copolymer made of poly(ethylene glycol)-b-poly(dihydroxyacetone) has also been reported recently by

Spector et al. to have promising mechanical hemostatic properties206. A mixture of tetra-

52 succinimidyl-derivatized and tetra-thiol-derivatized poly(ethylene glycol) has been recently evaluated as a hydrogel material for hemostatic liquid sealant207. In these studies,

it was shown that this material can crosslink with tissues and is capable of rendering a

mechanical hemostat effect by virtue of sealing a puncture hole in a rabbit artery bleed

model. Several successive evaluations in pre-clinical models and in clinical trails have led to approval of this material as a surgical sealant named CoSeal® (Baxter, USA)208,209. A

similar material has been developed by co-polymerizing poly(ethylene glycol) with poly(α- hydroxy acid) diacrylate, and after evaluation in vitro and in vivo, this material is currently marketed as AdvaSeal® (Ethicon Inc., USA). Thus, poly(ethyelene oxide) based synthetic external hemostatic sealants have undergone considerable clinical translation, especially in surgical bleeding applications.

Another class of synthetic polymers that has gained clinical significance in tissue sealant applications is poly(cyanoacrylates). This class of polymers was reported to have excellent tissue-adhesive properties via polar interactions with the tissue and was therefore used as a sutureless tissue sealant as early as the 1940s. Synthetic cyanoacrylates (e.g. 2-octyl

cyanoacrylate) under the product name Dermabond® (Ethicon Inc., New Jersey, USA),

was approved in the 1990s by the FDA for skin closure. Since then, this class of tissue

sealant has been widely used for surgical wound closure and hemostat applications116,210–

213. Long chain poly(cyanoacrylates) sealants have shown reduced tissue toxicity and have

progressed into clinically approved technologies like Histoacryl® (TissueSeal LLC, USA)

and GLUture® (World Precision Instruments) for topical applications in surgeries213,214.

Cyanoacrylate based tissue sealants have also been combined with tourniquet-based

53 procedures to evaluate their ability to staunch bleeding from larger hemorrhagic injuries215.

It should be noted that cyanoacrylates technically do not possess inherent ‘hemostatic’

property in the classical sense, but rather their mode of action is through physical sealing,

mechanical barrier and wound closure. Several other well-known polymeric biomaterials used in biomedical devices like sutures, contact lenses, coatings, drug delivery systems etc. have been investigated for hemostatic applications. For example, poly(glycolic acid) or

PGA is an established biomaterial that is used in a variety of clinically approved bioerodible devices like sutures and implants, and this polymer has been used to develop a hemostatic felt dressing. This felt (Soft PGA Felt, Aventis Behring, Germany) was evaluated in surgical hemostatic treatment of liver resection in patients, showing promising results as a hemostatic technology option combined with fibrin glue216. Poly(2- hydroxyethyl methacrylate) or poly-HEMA is an established biomaterial which had originally become known for its use in contact lens devices. Porous particles made from this polymer has been reported to have the capability of reducing blood loss in endovascular occlusion surgeries217. Poly(acrylic acid) is another synthetic polymer

biomaterial which was used to develop a hemostatic technology named Feracryl® by

Russian scientists in the 1980s218. This material was developed by reaction of acrylic acid

with Mohr’s salt in an aqueous medium and had a small percentage of iron(III) coordinated

with the polymer. In evaluation in vitro and in vivo, this material was found to be non-toxic

with promising hemostatic ability and this has led to development of several acrylic acid

based hemostatic technologies in Russia but not much has been reported elsewhere

globally. Carr et al have reported on a microporous poly(acrylamide) gel in the

development of a super-absorbent hemostatic technology named BioHemostat®

54 (Hemodyne Inc., USA), which showed great promise in the treatment of heavy bleeding

injuries219. The hemostatic ability of this material was attributed to its capacity of absorbing

high amount of fluids from the wound and expanding to exert a mechanical tamponade

effect on the wound. Casey et al. have recently reported on the development of a series of

cationic acrylamide hydrogels that demonstrated the capability of activating coagulation

factors that may aid in clot formation and hemostasis220. Another tissue sealant technology based on albumin-glutaraldehyde (e.g. BioGlue®, CryoLife Inc., Georgia, USA), is available in the United States for surgical adhesive applications in open surgical repair of large vessels (such as aorta, femoral and carotid arteries)221. In recent years, mussel-

inspired biomimetic materials design approaches have led to several interesting novel

classes of tissue-adhesive synthetic polymers, including citric-acid based, catechol based and hyperbranched poly(aminoester) based systems, which may find potential use as hemostats and sealants222–224. These studies and reports establish the promise of tissue- interactive synthetic polymeric systems for the engineering of hemostatic technologies, that can staunch bleeding through mechanical as well as biochemical mechanisms.

Recently, several research groups have reported on a synthetic peptide termed RADA16-I

(essentially a 4-mer repeat of Arginine-Alaninine-Aspartate-Alanine) that can self- assemble into supramolecular structures and can gel in presence of blood to arrest blood cells and promote a coagulatory effect225–228. The molecular mechanism of such hemostatic

action of this synthetic peptide material is thought to be more due to gelation and tissue

adhesion, and the presence of any additional biochemical involvement in the coagulation

cascade is currently unclear. Besides being used directly as absorbent dressing, sealant or

tamponade systems, synthetic polymers are also used as scaffold or backing components

55 in many hemostat technologies. Examples of these were previously described in

technologies where polyurethane films and foams were used as scaffold material for

dressings. In other composite applications, polymers like polypropylene and polyurethane

have been used as scaffold materials for impregnating chitosan material229,230. Dressings

made of polymer films, foams and amorphous hydrogels are considered to be ‘interactive

wound dressings’ as they are mostly transparent to allow of wound status, while

maintaining non-adherent and absorbent properties. The polymer film wound dressings are

usually made of semi-permeable polyurethane membrane with an acrylic adhesive backing,

while he polymer foam wound dressings are formed of soft, open cell, hydrophobic polyurethane foam sheet, approximately 6–8mm thick, with a high capacity of exudate absorption231. The film systems are usually used for low exudate wounds while the foam systems can be used for heavy bleeding and high exudate wounds. Polymer-based semi-

permeable dressings and foams also have the advantage of preventing bacteria and fluid

transfer at the injury site while allowing regulated transfer of air and moisture, such that

the wound bed can be kept optimally moist to aid comfort and healing. However, if the

wound site produces high amount of exudate, its build-up may negatively affect the

adjoining tissue (e.g. maceration). Altogether, a variety of synthetic polymeric biomaterials

have become stand-alone or integrative components of many hemostatic materials and

technologies, as they provide advantages of customizing chemistry for stimulating pro-

coagulant mechanisms while reducing immunogenicity risks otherwise associated with

some of the bio-derived materials.

56 2.5 Compression bandage technologies

The aspect of applying pressure to cause ‘stypsis’ and the aspect of applying bandages and

dressings (made from natural or synthetic biomaterials) to allow absorptive and interactive

mechanisms at the bleeding site, have been integrated to result in compression bandage

technologies. The simplest versions of such technologies consist of cotton gauze pads

placed with manual pressure over the bleeding wound, but this material as well as

procedure is unreliable due to variations in severity of injuries and external manual pressure

needed for efficient hemostatic compression232,233. A marginal improvement of this

technology is found in the Army Field Bandage where a thick pad of cotton is contained

within layers of gauze and tying straps are attached to this material to help fastening and

tightening of pressure234. A further improvement is found in the emergency field bandage

(aka ‘the Israeli bandage’, First Care Products) originally designed and developed by Israeli

military serviceman Ben Bar-Natan, which consists of an elastic bandage sewn over a

sterile non-adherent absorbent pad material. The bandage is equipped with a pressure bar

through which the wrapping material may be inserted, reversed in direction and tightened

over the bleeding site to apply tourniquet-like compression235. Unlike the direct placement

of gauze or gauze-wrapped cotton pad on the bleeding injury, which can make removal of

the bandage material cumbersome and risky, the utilization of non-adherent material along

with the compression bar made this bandage design a welcome improvement in

management of traumatic wounds in the military. Another military-tested compression

bandage is the CinchTight bandage system (H&H Medical Corporation, USA), which is

available in a variety of sizes and consists of long elastic wrapping sewn over a sterile non-

adherent absorbent pad and equipped with a metal hook (instead of the pressure bar found

57 in the Israeli bandage) through which the elastic wrap can be manipulated to wrap-around

and tighten for compressive force on the wound site. Another relatively recent technology

in the mix of compression bandages is the Elastic Adhesive (ELAD) bandage developed

by Dr. Sody Naimer in Israel236. In this design, the non-adherent sterile absorbent pad is

laminated over by a long self-adherent polyethylene strip that allows for flexible wrapping

around various tissue/organ morphologies while the transparency of the strip allows for

direct observation of the dressing-covered wound site to ensure that bleeding has been

reduced. In all the above designs, the main purpose is to provide a combination of

compression plus sterile coverage of the wound that can reduce bleeding to some extent,

and this reduction may be sufficient to reduce the tissue morbidity or mortality risk until

the patient can be attended to in a medical facility. The elastic wrap-around materials in

many cases are proprietary in composition, but it is apparent that most of these materials

are possibly latex or nylon or CobanTM (3M, USA) type non-woven elastic composites or

polyethylene-based stretchable polymers. Table 2-2 provides an ‘at-a-glance’ summary of synthetically derived hemostatic biomaterials along with representative technology names, characteristic evaluations and current application status.

58 Table 2-2 Topical and Externally Administered Hemostatic Biomaterials from Synthetic Sources

59 60

2.6 Combination systems and advanced technologies

The biggest challenge in external management of bleeding is the treatment of severely

bleeding wounds where the injury is heterogenous and the hemorrhage is often non-

compressible. This is significant in both civilian and military trauma cases where severe

exsanguination results in high mortality237. Therefore, while a staggering volume of

research endeavors, technology development and clinical translation have been dedicated

to development of hemostatic systems over the past century, the current focus is on efficient utilization of these systems as well as development of new systems to specifically treat non-compressible (often intracavitary) hemorrhage. This has not only prompted research in new materials and technologies, but also establishment of suitable pre-clinical animal models where the efficacies of these technologies can be evaluated and compared in a standardized way in terms of clotting mechanisms, time-to-clot, clot strength and stability, and survival (damage control resuscitation)232,238,239. The recent comparison studies using

such established animal models have provided interesting insight in the pros and cons of

various hemostatic materials. For example, in a comparison of WoundStat® (mineral

powder) verus Celox® (chitosan powder) using a heavy-bleeding arteriotomy model in

pigs, it was found that WoundStat® conferred better hemostasis, but Celox® conferred

better tissue-compatibility161. This can have potential impact on using these materials to

treat heavy-bleeding intracavitary wounds where the material is supposed to interact with

61 heterogenous tissue beds and would either biodegrade (e.g. for chitosan) or need to be

removed (e.g. for mineral) over time. In another study comparing QCG versus Celox®

versus chitosan-impregnated gauze versus HemCon® in porcine bleeding model, the QCG gauze was shown to have a superior performance than the rest, and it was also evident that such technologies used by a trained medical personnel (e.g. military trained in tactical combat casualty care or TCCC course) results in improved hemostasis compared to non- medical personnel240. Combining two different hemostatic materials/mechanisms may

provide a cumulative advantage over using each material by itself. This is evident in the

superior performance of DFSD (fibrinogen and thrombin impregnated dressing on polymer

mesh backing) and QCG (Quikclot zeolite impregnated gauze) compared to fibrin only or

Quikclot® zeolite only49,183. Another combination material example is a technology named

TraumaStat® (OreMedix, USA), where chitosan was combined with silica and polyethylene to form a flexible gauze that integrated the tissue adhesiveness of chitosan with the coagulation factor activating capacity (contact activation) of silica, to render superior hemostatic capability compared to just chitosan-based systems241,242.

2.7 Hemostatic materials and technologies for intracavitary applications

Hemostatic materials in gel, particulate, pelleted and other flexible forms may be more

useful in treating certain highly lethal hemorrhagic injuries like deep ballistic puncture

wounds and intracavitary blunt injuries where the standard-of-care is to rapidly administer

‘packing and absorbent’ material into the injury volume to staunch bleeding. The benefit of such ‘space-filling’ hemostatic materials was demonstrated by Kheirabadi et al using fibrin sealant foam and FloSeal® in a liver hemorrhage model (bleeding within abdominal

62 cavity) model243. To this end, in recent years several interesting biomaterials and advanced

technologies for intracavitary (or complex wound) hemostatic applications have been

reported. One such technology is made of plant-derived cellulose sponge pellets coated

with chitosan that are compressed as a starting form and upon absorption of blood (aqueous

medium), expand axially within a very short period of time (~ 20 seconds)244. As a result,

this material is capable of combining pro-coagulant (because of chitosan), absorptive

(because of cellulose) and tamponade (because of expanding) effects, to render efficient

hemostasis. After rigorous evaluation, this technology has recently undergone FDA-

approval under the name XSTAT® (RevMedX, Orgeon, USA)245. Self-expanding foams

made of in situ forming polyurethane mixtures have also been reported in the context of

evaluating hemostatic capability in heavy bleeding wounds246,247. In another recent

approach, an algae-based gel product developed by Landolina et al from alginate and other

polysachharides was shown to rapidly staunch bleeding in open wounds in soft tissue and

is being currently marketed in the veterinary as well as prospective human trauma hemostat

uses under the names Vetigel® and Traumagel® (Cresilon Inc. formerly Suneris Inc.,

USA)248. Translation of this material in hemostatic treatment of bleeding injuries in civilian

and military population would require further testing in established animal models, along

with determination of the biocompatibility and spatio-temporal fate of the material at the injury site.

2.8 Discussion

Bleeding due to injuries, trauma, surgical procedures and congenital or drug-induced coagulation disorders, can lead to a variety of health risks. Therefore, bleeding risks and

63 incidents need to be rapidly treated249,250. In many bleeding injuries, the body’s inherent

mechanisms become insufficient to rapidly staunch bleeding, and this necessitates the

administration of certain materials and technologies to facilitate, augment, compensate or

mimic the natural mechanisms of hemostasis. In the previous sections, I have

comprehensively reviewed the various materials and technologies currently approved, as

well as, under research for hemostatic applications. The bulk of hemostatic materials

research has been directed towards technologies for management externally accessible

bleeding scenarios, ranging from fibrin and thrombin glues, to zeolite powders, to bioactive

dressings, foams, gels, tourniquets and self-expanding tamponades. In recent years, there

has been a concerted effort to evaluate this large volume of technologies in standardized

bleeding models, especially for heavy-bleeding injuries. These efforts have been mostly led by the US military’s medical research divisions, since improving military survival through rapid hemostasis on the battlefield has become of paramount importance due to continued wars and conflicts over the last decade. In this context, Kheirabadi has recently described the ideal characteristics of externally administered hemostatic dressings251.

Arnaud et al reported on the comparative evaluation of ten clinically approved hemostatic

dressings in a groin transection model in pigs, by monitoring mean arterial blood pressure,

rate and time of survival, blood loss and post-treatment re-bleeding194,195. In these studies,

it was found that Quikclot® ACS (zeolite impregnated flexible absorbent sponge), Celox®

(chitosan granule), WoundStat® (smectic mineral on flexible polymer) and X-sponge®

(kaolin impregnated flexible gauze) all have comparable superior hemostatic capabilities

relative to standard gauze or rigid hemostatic technologies like HemCon®. Rall et al carried

out similar comparative evaluation of several such dressings in a pig arterial bleeding

64 model, and found that Quikclot®, Celox® and HemCon® do not have statistically significant differences in their hemostatic capabilities, and they all meet the current established standards put forward by the Committee on Tactical Combat Casualty Care for point-of-

care bleeding management252. Rall et al essentially concluded that the lack of clear

superiority of any of these hemostatic technologies is an indication that current clinically

approved hemostatic dressing technology has potentially reached a plateau for efficacy. A

potential improvement of this scenario may be achieved by the latest volume-filling technologies like XSTAT®, Traumagel® etc., especially to manage complex junctional,

intracavitary or open wound bleeding. Many of the current clinically approved or newer pre-clinically evaluated technologies contain bioactive materials for augmenting coagulation (minerals, collagen, chitosan, alginate, fibrinogen/fibrin etc.). Some of these may be biodegradable and resorbable, while some may require subsequent removal before surgical procedures. In this aspect, it is extremely important to study the biodegradation/resorption pathways, risk of residual materials post-removal and short and long term thrombo-inflammatory responses of resorbed or residual material. Almost all evaluations of the hemostatic materials and technologies so far have been in short term hemostatic capability and resultant survival, with limited focus on biodegradation

/resorption /removal pathways and long-term in vivo effects. Multi-disciplinary efforts involving biomedical and materials engineers, emergency medical and surgical personnel and physiology/pathology experts should be dedicated to such evaluations in the long term.

65 2.9 Acknowledgments

This work was supported by the National Heart Lung and Blood Institute (NHLBI) of the

National Institutes of Health under award numbers R01 HL121212 (PI: Sen Gupta). The content of this publication is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

2.10 References

1. Kragh JF, Swan KG, Smith DC, Mabry RL, Blackbourne LH. Historical review of

emergency tourniquet use to stop bleeding. Elsevier; 2012 Feb 1;203(2):242–252.

Available from:

https://www.sciencedirect.com/science/article/pii/S0002961011002546

2. Pilcher L. The Treatment of Wounds: Its Principles and Practice, General and

Special [Internet]. 1883. Available from:

https://books.google.com/books?hl=en&lr=&id=FzK6AAAAIAAJ&oi=fnd&pg=P

A37&ots=4ZYop1aJ1m&sig=eGq6CrJ6y4dyNLTUL6cJbbgqwsY

3. Lakstein D, Blumenfeld A, Sokolov T, Lin G, Bssorai R, Lynn M, Ben-Abraham

R. Tourniquets for hemorrhage control on the battlefield: a 4-year accumulated

experience. 2003;54(5 Suppl):S221–S225. PMID: 12768129

4. Kragh JF, Walters TJ, Baer DG, Fox CJ, Wade CE, Salinas J, Holcomb JB.

Survival With Emergency Tourniquet Use to Stop Bleeding in Major Limb

Trauma. 2009 Jan;249(1):1–7. Available from:

https://insights.ovid.com/crossref?an=00000658-200901000-00001

5. Walters TJ, Wenke JC, Kauvar DS, McManus JG, Holcomb JB, Baer DG.

66 Effectiveness of Self-Applied Tourniquets in Human Volunteers. Taylor &

Francis; 2005 Jan 2;9(4):416–422. Available from:

http://www.tandfonline.com/doi/full/10.1080/10903120500255123

6. Moore FA. Tourniquets. 2009 Jan;249(1):8–9. Available from:

https://insights.ovid.com/crossref?an=00000658-200901000-00002

7. Cain J. No Title. 2008;3(16).

8. Ivanova EP, Bazaka K, Crawford RJ. New functional biomaterials for medicine

and healthcare.

9. Pierce AM, Wiebkin OW, Wilson DF. SurgicelR: its fate following implantation.

Wiley/Blackwell (10.1111); 1984 Dec 1;13(6):661–670. Available from:

http://doi.wiley.com/10.1111/j.1600-0714.1984.tb01468.x

10. Lewis KM, Spazierer D, Urban MD, Lin L, Redl H, Goppelt A. Comparison of

regenerated and non-regenerated oxidized cellulose hemostatic agents. Springer

Vienna; 2013 Aug 4;45(4):213–220. Available from:

http://link.springer.com/10.1007/s10353-013-0222-z

11. Edwards JV, Prevost N. Thrombin production and human neutrophil elastase

sequestration by modified cellulosic dressings and their electrokinetic analysis.

2011 Dec 15;2(4):391–413. Available from: http://www.mdpi.com/2079-

4983/2/4/391 PMID: 24956451

12. Edwards J, Graves E, Bopp A, Prevost N, Santiago M, Condon B. Electrokinetic

and Hemostatic Profiles of Nonwoven Cellulosic/Synthetic Fiber Blends with

Unbleached Cotton. 2014;5(4):273–287. Available from:

http://www.mdpi.com/2079-4983/5/4/273/

67 13. Calvini P, Conio G, Lorenzoni M, Pedemonte E. Viscometric determination of

dialdehyde content in periodate oxycellulose. Part I. Methodology. Kluwer

Academic Publishers; 2004 Mar;11(1):99–107. Available from:

http://link.springer.com/10.1023/B:CELL.0000014766.31671.8d

14. Calvini P, Conio G, Princi E, Vicini S, Pedemonte E. Viscometric determination of

dialdehyde content in periodate oxycellulose Part II. Topochemistry of oxidation.

Springer Netherlands; 2006 Oct 28;13(5):571–579. Available from:

http://link.springer.com/10.1007/s10570-005-9035-y

15. Zimnitsky DS, Yurkshtovich TL, Bychkovsky PM. Synthesis and characterization

of oxidized cellulose. Wiley-Blackwell; 2004 Oct 1;42(19):4785–4791. Available

from: http://doi.wiley.com/10.1002/pola.20302

16. MILLER JM, JACKSON DA, COLLIER CS. An investigation of the chemical

reactions of oxidized regenerated cellulose. 1961;19:196–201. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/14474048 PMID: 14474048

17. Bajerová M, Krejčová K, Rabišková M, Gajdziok J, Masteiková R. Oxycellulose:

Significant characteristics in relation to its pharmaceutical and medical

applications. Wiley-Blackwell; 2009 Jun 1;28(3):199–208. Available from:

http://doi.wiley.com/10.1002/adv.20161

18. FRANTZ VK, LATTES R. OXIDIZED CELLULOSE—ABSORBABLE

GAUZE. American Medical Association; 1945 Nov 17;129(12):798. Available

from:

http://jama.jamanetwork.com/article.aspx?doi=10.1001/jama.1945.028604600220

06

68 19. Yackel EC, Kenyon WO. The Oxidation of Cellulose by Nitrogen Dioxide. 1942

Jan;64(1):121–127. Available from:

http://pubs.acs.org/doi/abs/10.1021/ja01253a032

20. Frantz VK. ABSORBABLE COTTON, PAPER AND GAUZE : (OXIDIZED

CELLULOSE). Lippincott, Williams, and Wilkins; 1943 Jul;118(1):116–26.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/17858245 PMID: 17858245

21. Wu Y, He J, Cheng W, Gu H, Guo Z, Gao S, Huang Y. Oxidized regenerated

cellulose-based hemostat with microscopically gradient structure. Elsevier; 2012

Apr 15;88(3):1023–1032. Available from:

https://www.sciencedirect.com/science/article/pii/S0144861712000781

22. Sassan K, Macdougall M, Lulic D, Kasasbeh A, Levy M. Clinical Experience with

the Surgicel Family of Absorbable Hemostats (Oxidized Regenerated Cellulose) in

Neurosurgical Applications: A Review. 2013;25(6):160–167. PMID: 25866982

23. Christenson LJ, Otley CC, Roenigk RK. Oxidized regenerated cellulose gauze for

hemostasis of a two-stage interpolation flap pedicle. 2004;30(12 Pt 2):1593–4.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/15606850 PMID: 15606850

24. Mair H, Kaczmarek I, Oberhoffer M, Groetzner J, Daebritz S, Reichart B. Surgicel

Nu-Knit hemostat for bleeding control of fragile sternum. Elsevier; 2005 Aug

1;130(2):605–606. Available from:

http://linkinghub.elsevier.com/retrieve/pii/S0022522305006847

25. Arat YO, Dorotheo EU, Tang RA, Boniuk M, Schiffman JS. Compressive Optic

Neuropathy after Use of Oxidized Regenerated Cellulose in Orbital Surgery:

Review of Complications, Prophylaxis, and Treatment. Elsevier; 2006 Feb

69 1;113(2):333–337. Available from:

https://www.sciencedirect.com/science/article/pii/S0161642005012996

26. Laurence S, Bareille R, Baquey C, Fricain JC. Development of a resorbable

macroporous cellulosic material used as hemostatic in an osseous environment.

2005 Jun 15;73A(4):422–429. Available from:

http://doi.wiley.com/10.1002/jbm.a.30280

27. Dan Dimitrijevich S, Tatarko M, Gracy RW, Linsky CB, Olsen C. Biodegradation

of oxidized regenerated cellulose. Elsevier; 1990 Jan 15;195(2):247–256.

Available from:

https://www.sciencedirect.com/science/article/pii/000862159084169U

28. Dan Dimitrijevich S, Tatarko M, Gracy RW, Wise GE, Oakford LX, Linsky CB,

Kamp L. In vivo degradation of oxidized, regenerated cellulose. Elsevier; 1990

May 1;198(2):331–341. Available from:

https://www.sciencedirect.com/science/article/pii/000862159084303C

29. Muench TR, Kong W, Harmon AM. The performance of a hemostatic agent based

on oxidized regenerated cellulose – polyglactin 910 composite in a liver defect

model in immunocompetent and athymic rats. Elsevier; 2010 May 1;31(13):3649–

3656. Available from:

https://www.sciencedirect.com/science/article/pii/S014296121000102X

30. Hoekstra MJ, Hermans MHE, Richters CD, Dutrieux RP. A histological

comparison of acute inflammatory responses with a hydrofibre or tulle gauze

dressing. MA Healthcare London ; 2002 Mar 29;11(3):113–117. Available from:

http://www.magonlinelibrary.com/doi/10.12968/jowc.2002.11.3.26384

70 31. Morgan D. Wound management products in the drug tariff. 1999; Available from:

https://scholar.google.com/scholar_lookup?hl=en&volume=263&publication_year

=1999&pages=820&journal=Pharm.+J.&author=D.+A.+Morgan&title=Pharm.+J.

32. Moody A. Use of a hydrogel dressing for management of a painful leg ulcer. MA

Healthcare London ; 2006 Jun 27;11(Sup3):S12–S17. Available from:

http://www.magonlinelibrary.com/doi/10.12968/bjcn.2006.11.Sup3.21212

33. Bennett-Marsden M. How to select a wound dressing [Internet]. 2010. Available

from: https://www.pharmaceutical-journal.com/career/career-feature/how-to-

select-a-wound-dressing/11039846.article?firstPass=false

34. Grey E. Fibrin as a hemostatic in cerebral surgery. 1915;(21):452.

35. Harvey S. The use of fibrin paper and forms in surgery. 1916;(174):658.

36. Spotnitz WD. Fibrin Sealant: Past, Present, and Future: A Brief Review. Springer-

Verlag; 2010 Apr 10;34(4):632–634. Available from:

http://link.springer.com/10.1007/s00268-009-0252-7

37. Albala DM. Fibrin sealants in clinical practice. No longer published by Elsevier;

2003 Aug 1;11:5–11. Available from:

https://www.sciencedirect.com/science/article/pii/S0967210903000656

38. Sierra DH. Fibrin Sealant Adhesive Systems: A Review of Their Chemistry,

Material Properties and Clinical Applications. Sage PublicationsSage CA:

Thousand Oaks, CA; 1993 Apr 1;7(4):309–352. Available from:

http://journals.sagepub.com/doi/10.1177/088532829300700402

39. Borst HG, Haverich A, Walterbusch G, Maatz W. Fibrin adhesive: an important

hemostatic adjunct in cardiovascular operations. 1982 Oct;84(4):548–53.

71 Available from: http://www.ncbi.nlm.nih.gov/pubmed/6981733 PMID: 6981733

40. Sawamura Y, Asaoka K, Terasaka S, Tada M, Uchida T. Evaluation of

Application Techniques of Fibrin Sealant to Prevent Cerebrospinal Fluid Leakage:

A New Device for the Application of Aerosolized . Oxford University

Press; 1999 Feb 1;44(2):332–337. Available from:

https://academic.oup.com/neurosurgery/article/44/2/332/2858127

41. Mandell SP, Gibran NS. Fibrin sealants: surgical hemostat, sealant and adhesive.

Taylor & Francis; 2014 Jun 13;14(6):821–830. Available from:

http://www.tandfonline.com/doi/full/10.1517/14712598.2014.897323

42. CRONKITE EP, LOZNER EL, DEAVER JM. USE OF THROMBIN AND

FIBRINOGEN IN SKIN GRAFTING. American Medical Association; 1944 Apr

1;124(14):976. Available from:

http://jama.jamanetwork.com/article.aspx?doi=10.1001/jama.1944.028501400220

06

43. Tidrick RT, Warner ED. Fibrin fixation of skin transplants. Elsevier; 1944 Jan

1;15(1):90–95. Available from: https://www.surgjournal.com/article/S0039-

6060(44)90052-8/abstract

44. Erdogan D, van Gulik TM. Evolution of fibrinogen-coated collagen patch for use

as a topical hemostatic agent. 2008 Apr;85B(1):272–278. Available from:

http://doi.wiley.com/10.1002/jbm.b.30916

45. Scheyer M, Zimmermann G. Tachocomb used in endoscopic surgery. Springer-

Verlag; 1996 May;10(5):501–503. Available from:

http://link.springer.com/10.1007/BF00188394

72 46. Schiele U, Kuntz G, Riegler A. Haemostyptic preparations on the basis of collagen

alone and as fixed combination with fibrin glue. Elsevier; 1992 Jan 1;9(3–4):169–

177. Available from:

https://www.sciencedirect.com/science/article/pii/026766059290097D

47. Larson MJ, Bowersox JC, Lim RC, Hess JR. Efficacy of a Fibrin Hemostatic

Bandage in Controlling Hemorrhage From Experimental Arterial Injuries.

American Medical Association; 1995 Apr 1;130(4):420. Available from:

http://archsurg.jamanetwork.com/article.aspx?doi=10.1001/archsurg.1995.014300

40082018

48. Jackson MR, Friedman SA, Carter AJ, Bayer V, Burge JR, MacPhee MJ, Drohan

WN, Alving BM. Hemostatic efficacy of a fibrin sealant–based topical agent in a

femoral artery injury model: A randomized, blinded, placebo-controlled study.

Mosby; 1997 Aug 1;26(2):274–280. Available from:

https://www.sciencedirect.com/science/article/pii/S0741521497701897

49. Pusateri AE, Kheirabadi BS, Delgado A V, Doyle JW, Kanellos J, Uscilowicz JM,

Martinez RS, Holcomb JB, Modrow HE. Structural design of the dry fibrin sealant

dressing and its impact on the hemostatic efficacy of the product. 2004;70(1):114–

21. Available from: http://www.ncbi.nlm.nih.gov/pubmed/15199591 PMID:

15199591

50. Holcomb J, MacPhee M, Hetz S, Harris R, Pusateri A, Hess J. Efficacy of a Dry

Fibrin Sealant Dressing for Hemorrhage Control After Ballistic Injury. American

Medical Association; 1998 Jan 1;133(1):32. Available from:

http://archsurg.jamanetwork.com/article.aspx?doi=10.1001/archsurg.133.1.32

73 51. Holcomb JB, Pusateri AE, Harris RA, Charles NC, Gomez RR, Cole JP, Beall LD,

Bayer V, MacPhee MJ, Hess JR. Effect of Dry Fibrin Sealant Dressings versus

Gauze Packing on Blood Loss in Grade V Liver Injuries in Resuscitated Swine.

1999;46(1). Available from:

https://journals.lww.com/jtrauma/Fulltext/1999/01000/Effect_of_Dry_Fibrin_Seal

ant_Dressings_versus.9.aspx

52. Jackson MR, Taher MM, Burge JR, Krishnamurti C, Reid TJ, Alving BM.

Hemostatic Efficacy of a Fibrin Sealant Dressing in an Animal Model of Kidney

Injury. 1998;45(4). Available from:

https://journals.lww.com/jtrauma/Fulltext/1998/10000/Hemostatic_Efficacy_of_a_

Fibrin_Sealant_Dressing.3.aspx

53. Kheirabadi BS, Acheson EM, Deguzman R, Crissey JM, Delgado A V., Estep SJ,

Holcomb JB. The Potential Utility of Fibrin Sealant Dressing in Repair of

Vascular Injury in Swine. 2007 Jan;62(1):94–103. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17215739 PMID: 17215739

54. Pusateri AE, Holcomb JB, Harris RA, MacPhee MJ, Charles NC, Beall LD, Hess

JR. Effect of Fibrin Bandage Fibrinogen Concentration on Blood Loss after Grade

V Liver Injury in Swine. Oxford University Press; 2001 Mar 1;166(3):217–222.

Available from: https://academic.oup.com/milmed/article/166/3/217/4819512

55. MOREY AF, ANEMA JG, HARRIS R, GRESHAM V, DANIELS R, KNIGHT

RW, BEALL D, MACPHEE M, CORNUM RL. TREATMENT OF GRADE 4

RENAL STAB WOUNDS WITH ABSORBABLE FIBRIN ADHESIVE

BANDAGE IN A PORCINE MODEL. Elsevier; 2001 Mar 1;165(3):955–958.

74 Available from:

https://www.sciencedirect.com/science/article/pii/S0022534705665834

56. Anema JG, Morey AF, Harris R, MacPhee M, Cornum RL. Potential Uses of

Absorbable Fibrin Adhesive Bandage for Genitourinary Trauma. Springer-Verlag;

2001 Dec 22;25(12):1573–1577. Available from:

http://link.springer.com/10.1007/s00268-001-0152-y

57. CORNUM R, BELL J, GRESHAM V, BRINKLEY W, BEALL D, MACPHEE

M. INTRAOPERATIVE USE OF THE ABSORBABLE FIBRIN ADHESIVE

BANDAGE: LONG TERM EFFECTS. Elsevier; 1999 Nov 1;162(5):1817–1820.

Available from:

https://www.sciencedirect.com/science/article/pii/S0022534705682444

58. Delgado A V., Kheirabadi BS, Fruchterman TM, Scherer M, Cortez D, Wade CE,

Dubick MA, Holcomb JB. A Novel Biologic Hemostatic Dressing (Fibrin Patch)

Reduces Blood Loss and Resuscitation Volume and Improves Survival in

Hypothermic, Coagulopathic Swine With Grade V Liver Injury. 2008

Jan;64(1):75–80. Available from: https://insights.ovid.com/crossref?an=00005373-

200801000-00012

59. Lundblad RL, Bradshaw RA, Gabriel D, Ortel T, Lawson J, Mann KG. A review

of the therapeutic uses of thrombin. Schattauer GmbH; 2004 Feb 27;91(05):851–

860. Available from:

http://www.schattauer.de/index.php?id=1214&doi=10.1160/TH03-12-

0792&no_cache=1

60. Jackson MR. Fibrin sealants in surgical practice: An overview. Elsevier; 2001 Aug

75 1;182(2):S1–S7. Available from:

https://www.sciencedirect.com/science/article/pii/S000296100100770X

61. Kjaergard HK, Weis-Fogh US. Important factors influencing the strength of

autologous fibrin glue; the fibrin concentration and reaction time--comparison of

strength with commercial fibrin glue. Karger Publishers; 1994;26(5):273–6.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/7835384 PMID: 7835384

62. Saltz R, Sierra D, Feldman D, Saltz MB, Dimick A, Vasconez LO. Experimental

and clinical applications of fibrin glue. 1991 Dec;88(6):1005-15; discussion 1016–

7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/1946751 PMID: 1946751

63. Kheirabadi BS, Pearson R, Rudnicka K, Somwaru L, MacPhee M, Drohan W,

Tuthill D. Development of an Animal Model for Assessment of the Hemostatic

Efficacy of Fibrin Sealant in . Academic Press; 2001 Sep

1;100(1):84–92. Available from:

https://www.sciencedirect.com/science/article/pii/S0022480401962262

64. Joch C. The safety of fibrin sealants. No longer published by Elsevier; 2003 Aug

1;11:23–28. Available from:

https://www.sciencedirect.com/science/article/pii/S0967210903000681

65. Izumi Y, Gika M, Shinya N, Miyabashira S, Imamura T, Nozaki C, Kawamura M,

Kobayashi K. Hemostatic Efficacy of a Recombinant Thrombin-Coated

Polyglycolic Acid Sheet Coupled With Liquid Fibrinogen, Evaluated in a Canine

Model of Pulmonary Arterial Hemorrhage. 2007 Oct;63(4):783–787. Available

from: https://insights.ovid.com/crossref?an=00005373-200710000-00010

66. Rothwell SW, Fudge JM, Reid TJ, Krishnamurti C. ε-Amino caproic acid additive

76 decreases fibrin bandage performance in a swine arterial bleeding model.

Pergamon; 2002 Dec 15;108(5–6):341–345. Available from:

https://www.sciencedirect.com/science/article/pii/S0049384803001117

67. Phillips M, Dickneite G, Metzner H. Fibrin sealants in supporting surgical

techniques: strength in factor XIII. No longer published by Elsevier; 2003 Aug

1;11:13–16. Available from:

https://www.sciencedirect.com/science/article/pii/S0967210903000668

68. Busuttil RW. A comparison of antifibrinolytic agents used in hemostatic fibrin

sealants. Elsevier; 2003 Dec 1;197(6):1021–1028. Available from:

http://linkinghub.elsevier.com/retrieve/pii/S1072751503008147

69. Kheirabadi BS, Pearson R, Tuthill D, Rudnicka K, Holcomb JB, Drohan W,

MacPhee MJ. Comparative Study of the Hemostatic Efficacy of a New Human

Fibrin Sealant: Is an Antifibrinolytic Agent Necessary? 2002;52(6). Available

from:

https://journals.lww.com/jtrauma/Fulltext/2002/06000/Comparative_Study_of_the

_Hemostatic_Efficacy_of_a.14.aspx

70. Ortel TL, Mercer MC, Thames EH, Moore KD, Lawson JH. Immunologic impact

and clinical outcomes after surgical exposure to bovine thrombin. Lippincott,

Williams, and Wilkins; 2001 Jan;233(1):88–96. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/11141230 PMID: 11141230

71. Kheirabadi BS, Pusateri AE, Sondeen JL, Delgado A V., Modrow HE, Hess JR,

Holcomb JB. Development of Hemostatic Dressings for Use in Military

Operations. 2004;(August):P34-1-P34-12.

77 72. Krishnan LK, Mohanty M, Umashankar PR, Vijayan Lal A. Comparative

evaluation of absorbable hemostats: advantages of fibrin-based sheets. Elsevier;

2004 Nov 1;25(24):5557–5563. Available from:

https://www.sciencedirect.com/science/article/pii/S0142961204000146

73. Wang LZ, Gorlin J, Michaud SE, Janmey PA, Goddeau RP, Kuuse R, Uibo R,

Adams D, Sawyer ES. Purification of Salmon Clotting Factors and Their Use as

Tissue Sealants. Pergamon; 2000 Dec 15;100(6):537–548. Available from:

https://www.sciencedirect.com/science/article/pii/S0049384800003625

74. Michaud SE, Wang LZ, Korde N, Bucki R, Randhawa PK, Pastore JJ, Falet H,

Hoffmeister K, Kuuse R, Uibo R, Herod J, Sawyer E, Janmey PA. Purification of

salmon thrombin and its potential as an alternative to mammalian in

fibrin sealants. Pergamon; 2002 Sep 1;107(5):245–254. Available from:

https://www.sciencedirect.com/science/article/pii/S004938480200333X

75. Rothwell SW, Reid TJ, Dorsey J, Flournoy WS, Bodo M, Janmey PA, Sawyer E.

A Salmon Thrombin-Fibrin Bandage Controls Arterial Bleeding in a Swine

Aortotomy Model. 2005 Jul;59(1):143–149. Available from:

https://insights.ovid.com/crossref?an=00005373-200507000-00023

76. Laidmäe I, McCormick ME, Herod JL, Pastore JJ, Salum T, Sawyer ES, Janmey

PA, Uibo R. Stability, sterility, coagulation, and immunologic studies of salmon

coagulation proteins with potential use for mammalian wound healing and cell

engineering. Elsevier; 2006 Dec 1;27(34):5771–5779. Available from:

https://www.sciencedirect.com/science/article/pii/S0142961206006600

77. Rothwell SW, Settle T, Wallace S, Dorsey J, Simpson D, Bowman JR, Janmey P,

78 Sawyer E. The long term immunological response of swine after two exposures to

a salmon thrombin and fibrinogen hemostatic bandage. Academic Press; 2010 Nov

1;38(6):619–628. Available from:

https://www.sciencedirect.com/science/article/pii/S1045105610001326

78. Mullin JL, Gorkun O V, Binnie CG, Lord ST. Recombinant fibrinogen studies

reveal that thrombin specificity dictates order of fibrinopeptide release. American

Society for Biochemistry and Molecular Biology; 2000 Aug 18;275(33):25239–46.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/10837485 PMID: 10837485

79. Ham SW, Lew WK, Weaver FA. Thrombin use in surgery: an evidence-based

review of its clinical use. Dove Press; 2010;1:135–42. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/22282693 PMID: 22282693

80. Cheng CM, Meyer-Massetti C, Kayser SR. A review of three stand-alone topical

thrombins for surgical hemostasis. Elsevier; 2009 Jan 1;31(1):32–41. Available

from: https://www.sciencedirect.com/science/article/pii/S0149291809000198

81. Bishop P, Lawson J. Recombinant biologics for treatment of bleeding disorders.

Nature Publishing Group; 2004 Aug 1;3(8):684–694. Available from:

http://www.nature.com/articles/nrd1443

82. Houdebine L-M. Production of pharmaceutical proteins by transgenic animals.

Pergamon; 2009 Mar 1;32(2):107–121. Available from:

https://www.sciencedirect.com/science/article/pii/S0147957107001300

83. Spotnitz WD. Fibrin Sealant: The Only Approved Hemostat, Sealant, and

Adhesive-a Laboratory and Clinical Perspective. Hindawi; 2014 Mar

4;2014:203943. Available from: http://www.ncbi.nlm.nih.gov/pubmed/24729902

79 PMID: 24729902

84. Buchta C, Dettke M, Funovics PT, Hocker P, Knobl P, Macher M, Quehenberger

P, Treitl C, Worel N. Fibrin sealant produced by the CryoSealR FS System:

product chemistry, material properties and possible preparation in the autologous

preoperative setting. Wiley/Blackwell (10.1111); 2004 May 1;86(4):257–262.

Available from: http://doi.wiley.com/10.1111/j.0042-9007.2004.00516.x

85. Davidson BR, Burnett S, Javed MS, Seifalian A, Moore D, Doctor N.

Experimental study of a novel fibrin sealant for achieving haemostasis following

partial hepatectomy. Wiley-Blackwell; 2000 Jun 1;87(6):790–795. Available from:

http://www.blackwell-synergy.com/links/doi/10.1046/j.1365-2168.2000.01427.x

86. Kjaergard HK, Trumbull HR. Bleeding from the sternal marrow can be stopped

using vivostat patient-derived fibrin sealant. Elsevier; 2000 Apr 1;69(4):1173–

1175. Available from:

https://www.sciencedirect.com/science/article/pii/S000349759901560X

87. Drake DB, Wong LG. Hemostatic Effect of Vivostat Patient-Derived Fibrin

Sealant on Split-Thickness Skin Graft Donor Sites. 2003 Apr;50(4):367–372.

Available from: https://insights.ovid.com/crossref?an=00000637-200304000-

00007

88. Fukunaga N, Uchida T, Kaetsu H, Kawakami K, Ishihara Y, Funatsu A. Effective

application of fibrin sealant using a spray device with a double-tube structure.

Elsevier; 1998 Oct 1;18(5):345–349. Available from:

https://www.sciencedirect.com/science/article/pii/S0143749698000165

89. Holcomb JB, McClain JM, Pusateri AE, Beall D, Macaitis JM, Harris RA,

80 MacPhee MJ, Hess JR. Fibrin sealant foam sprayed directly on liver injuries

decreases blood loss in resuscitated rats. 2000;49(2):246–250. PMID: 10963535

90. Kheirabadi BS, Sieber J, Bukhari T, Rudnicka K, Murcin LA, Tuthill D. High-

Pressure Fibrin Sealant Foam: An Effective Hemostatic Agent for Treating Severe

Parenchymal Hemorrhage. Academic Press; 2008 Jan 1;144(1):145–150.

Available from:

https://www.sciencedirect.com/science/article/pii/S0022480407000704

91. FINLEY DS, LEE DI, EICHEL L, URIBE CA, McDOUGALL EM, CLAYMAN

R V. FIBRIN GLUE-OXIDIZED CELLULOSE SANDWICH FOR

LAPAROSCOPIC WEDGE RESECTION OF SMALL RENAL LESIONS.

Elsevier; 2005 May 1;173(5):1477–1481. Available from:

https://www.sciencedirect.com/science/article/pii/S0022534705606003

92. Friess W. Collagen – biomaterial for drug delivery. Elsevier; 1998 Mar

1;45(2):113–136. Available from:

https://www.sciencedirect.com/science/article/pii/S0939641198000174

93. Gelse K, Pöschl E, Aigner T. Collagens—structure, function, and biosynthesis.

Elsevier; 2003 Nov 28;55(12):1531–1546. Available from:

https://www.sciencedirect.com/science/article/pii/S0169409X03001820

94. Ruggeri ZM, Mendolicchio GL. Adhesion Mechanisms in Platelet Function. 2007

Jun 22;100(12):1673–1685. Available from:

http://circres.ahajournals.org/cgi/doi/10.1161/01.RES.0000267878.97021.ab

95. Savage B, Almus-Jacobs F, Ruggeri ZM. Specific Synergy of Multiple Substrate–

Receptor Interactions in Platelet Thrombus Formation under Flow. Cell Press;

81 1998 Sep 4;94(5):657–666. Available from:

https://www.sciencedirect.com/science/article/pii/S0092867400816074

96. CHEN J, LÓPEZ JA. Interactions of Platelets with Subendothelium and

Endothelium. 2005 Jan;12(3):235–246. Available from:

http://doi.wiley.com/10.1080/10739680590925484

97. Roberts H, Hoffman M, Monroe D. A Cell-Based Model of Thrombin Generation.

2006 Feb;32(S 1):032–038. Available from: http://www.thieme-

connect.de/DOI/DOI?10.1055/s-2006-939552

98. Hoffman M. Remodeling the Blood Coagulation Cascade. 2003 Aug;16(1/2):17–

20. Available from:

http://link.springer.com/10.1023/B:THRO.0000014588.95061.28

99. Farndale RW, Sixma JJ, Barnes MJ, De Groot PG. The role of collagen in

thrombosis and hemostasis. 2004;2(4):561–573. PMID: 15102010

100. R HM, Battista OA, Stark RB, McCord CW. Mirocrystalline collagen as a biologic

dressing, vascular prosthesis, and hemostatic agent. 1970;45(5):523.

101. Hait MR. Microcrystalline collagen. A new hemostatic agent. 1970

Sep;120(3):330. Available from: http://www.ncbi.nlm.nih.gov/pubmed/5456913

PMID: 5456913

102. Hait MR, Robb CA, Baxter CR, Borgmann AR, Tippett LO. Comparative

evaluation of avitene microcrystalline collagen hemostat in experimental animal

wounds. Elsevier; 1973 Mar 1;125(3):284–287. Available from:

https://www.sciencedirect.com/science/article/pii/0002961073900421

103. Silverstein ME, Keown K, Owen JA, Chvapil M. Collagen fibers as a fleece

82 hemostatic agent. 1980 Aug;20(8):688–94. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/7401211 PMID: 7401211

104. Wagner WR, Pachence JM, Ristich J, Johnson PC. Comparativein VitroAnalysis

of Topical Hemostatic Agents. Academic Press; 1996 Dec 1;66(2):100–108.

Available from:

https://www.sciencedirect.com/science/article/pii/S0022480496903790

105. Zwischenberger JB, Robert L. RLB, Swann JR, Conti VR. Comparison of Two

Topical Collagen-Based Hemostatic Sponges During Cardiothoracic Procedures.

Taylor & Francis; 1999 Jan 9;12(2):101–106. Available from:

http://www.tandfonline.com/doi/full/10.1080/089419399272656

106. Yang C, Hillas P, Tang J, Balan J, Notbohm H, Polarek J. Development of a

recombinant human collagen-type III based hemostat. 2004 Apr 15;69B(1):18–24.

Available from: http://doi.wiley.com/10.1002/jbm.b.20030

107. Browder IW, Litwin MS. Use of absorbable collagen for hemostasis in general

surgical patients. 1986 Sep;52(9):492–4. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/3752727 PMID: 3752727

108. Silverstein ME, Chvapil M. Experimental and clinical experiences with collagen

fleece as a hemostatic agent. 1981 May;21(5):388–93. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/7230285 PMID: 7230285

109. Prior JJ, Wallace DG, Harner A, Powers N. A sprayable hemostat containing

fibrillar collagen, bovine thrombin, and autologous plasma. Elsevier; 1999 Aug

1;68(2):479–485. Available from:

https://www.sciencedirect.com/science/article/pii/S0003497599005524

83 110. Nelson PA, Powers JN, Estridge TD, Elder EA, Alea AD, Sidhu PK, Sehl LC,

DeLustro FA. Serological analysis of patients treated with a new surgical hemostat

containing bovine proteins and autologous plasma. Wiley-Blackwell; 2001 Jan

1;58(6):710–719. Available from: http://doi.wiley.com/10.1002/jbm.10025

111. Chapman WC, Sherman R, Boyce S, Malawer M, Hill A, Buncke G, Block JE,

Fung JJ, Clavien P, Lee KF, Lebovic GS, Wren SM, Diethrich E, Goldsein R. A

novel collagen-based composite offers effective hemostasis for multiple surgical

indications: Results of a randomized controlled trial. Mosby; 2001 Apr

1;129(4):445–450. Available from:

https://www.sciencedirect.com/science/article/pii/S0039606001818900

112. Bochicchio G, Dunne J, Bochicchio K, Scalea T. The Combination of Platelet-

Enriched Autologous Plasma with Bovine Collagen and Thrombin Decreases the

Need for Multiple Blood Transfusions in Trauma Patients with Retroperitoneal

Bleeding. 2004 Jan;56(1):76–79. Available from:

https://insights.ovid.com/crossref?an=00005373-200401000-00012

113. Bigi A, Cojazzi G, Panzavolta S, Roveri N, Rubini K. Stabilization of gelatin films

by crosslinking with genipin. Elsevier; 2002 Dec 1;23(24):4827–4832. Available

from: https://www.sciencedirect.com/science/article/pii/S0142961202002351

114. Bertoni F, Barbani N, Giusti P, Ciardelli G. Transglutaminase Reactivity with

Gelatine: Perspective Applications in . Springer Netherlands;

2006 May 24;28(10):697–702. Available from:

http://link.springer.com/10.1007/s10529-006-9046-2

115. Kabiri M, Emami SH, Rafinia M, Tahriri M. Preparation and characterization of

84 absorbable hemostat crosslinked gelatin sponges for surgical applications. North-

Holland; 2011 May 1;11(3):457–461. Available from:

https://www.sciencedirect.com/science/article/pii/S1567173910002853

116. Traver MA, Assimos DG. New generation tissue sealants and hemostatic agents:

innovative urologic applications. MedReviews, LLC; 2006;8(3):104–11. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/17043707 PMID: 17043707

117. Xie X, Tian J, Lv F, Wu R, Tang W, Luo Y, Huang Y, Tang J. A novel hemostatic

sealant composed of gelatin, transglutaminase and thrombin effectively controls

liver trauma-induced bleeding in dogs. Nature Publishing Group; 2013 Jul

6;34(7):983–988. Available from: http://www.nature.com/articles/aps201317

118. Eastoe JE, Leach AA, editors. Chemical Constitution of Gelatin. London, UK:

Academic Press; 1977.

119. Jenkins H, Senz E, Owen H, Jampolis R. Present status of gelatin sponge for the

control of hemorrhage: With experimental data on its use for wounds of the great

vessels and the heart. 1946;132:614. Available from:

https://jamanetwork.com/data/journals/JAMA/7264/jama_132_11_002.pdf

120. FAIRBAIRN JW, WHITTET TD. Absorbable haemostatics, their uses and

identification. 1948 Feb 28;106(4400):149. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/18911857 PMID: 18911857

121. Permpongkosol S, Nicol TL, Bagga HS, Kohanim S, Kavoussi L, Solomon SB.

Prophylactic Gelatin Sponge Tract Injection to Prevent Bleeding after

Percutaneous Renal Cryoablation in a Swine Model. Elsevier; 2006 Sep

1;17(9):1505–1509. Available from:

85 https://www.sciencedirect.com/science/article/pii/S1051044307609336

122. GRIFFITH BC, MOREY AF, ROZANSKI TA, HARRIS R, DALTON SR,

TORGERSON SJ, PARTYKA SR. Central Renal Stab Wounds:: Treatment With

Augmented Fibrin Sealant in a Porcine Model. Elsevier; 2004 Jan 1;171(1):445–

447. Available from:

https://www.sciencedirect.com/science/article/pii/S0022534705627772

123. Lewis KM, Atlee HD, Mannone AJ, Dwyer J, Lin L, Goppelt A, Redl H.

Comparison of Two Gelatin and Thrombin Combination Hemostats in a Porcine

Liver Abrasion Model. Taylor & Francis; 2013 Jun 20;26(3):141–148. Available

from: http://www.tandfonline.com/doi/full/10.3109/08941939.2012.724519

124. L’Esperance JO, Sung JC, Marguet CG, Maloney ME, Springhart WP, Preminger

GM, Albala DM. Controlled Survival Study of the Effects of Tisseel or a

Combination of FloSeal and Tisseel on Major Vascular Injury and Major

Collecting-System Injury during Partial Nephrectomy in a Porcine Model. Mary

Ann Liebert, Inc. 2 Madison Avenue Larchmont, NY 10538 USA ; 2005 Nov

11;19(9):1114–1121. Available from:

http://www.liebertpub.com/doi/10.1089/end.2005.19.1114

125. BAK JB, SINGH A, SHEKARRIZ B. Use of Gelatin Matrix Thrombin Tissue

Sealant as an Effective Hemostatic Agent During Laparoscopic Partial

Nephrectomy. Elsevier; 2004 Feb 1;171(2):780–782. Available from:

https://www.sciencedirect.com/science/article/pii/S0022534705625803

126. Weaver FA, Hood DB, Zatina M, Messina L, Badduke B. Gelatin–thrombin-based

Hemostatic Sealant for Intraoperative Bleeding in Vascular Surgery. Springer-

86 Verlag; 2002 May;16(3):286–293. Available from:

http://linkinghub.elsevier.com/retrieve/pii/S0890509606622772

127. Oz MC, Rondinone JF, Shargill NS. Floseal Matrix:. New Generation Topical

Hemostatic Sealant. Blackwell Science Inc; 2003 Nov;18(6):486–493. Available

from: http://doi.wiley.com/10.1046/j.0886-0440.2003.00302.x

128. Leixnering M, Reichetseder J, Schultz A, Figl M, Wassermann E, Thurnher M,

Redl H. Gelatin Thrombin Granules for Hemostasis in a Severe Traumatic Liver

and Spleen Rupture Model in Swine. 2008 Feb;64(2):456–461. Available from:

https://insights.ovid.com/crossref?an=00005373-200802000-00032

129. Bachet J, Goudot B, Dreyfus G, Banfi C, Ayle NA, Aota M, Brodaty D, Dubois C,

Delentdecker P, Guilmet D. The proper use of glue: a 20-year experience with the

GRF glue in acute aortic dissection. 1997;12(2 Suppl):243-53; discussion 253–5.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/9271753 PMID: 9271753

130. Kodama K, Doi O, Higashiyama M, Yokouchi H. Pneumostatic effect of gelatin-

resorcinol formaldehyde-glutaraldehyde glue on thermal injury of the lung: an

experimental study on rats [Internet]. 1997. Available from:

https://academic.oup.com/ejcts/article-abstract/11/2/333/374926

131. Kunihara T, Shiiya N, Matsuzaki K, Murashita T, Matsui Y. Recommendation for

Appropriate Use of GRF Glue in the Operation for Acute Aortic Dissection

[Internet]. 2008. Available from:

https://pdfs.semanticscholar.org/8312/a0ec657dc25906092a289592d9acee8cbe42.

pdf

132. Olsen D, Yang C, Bodo M, Chang R, Leigh S, Baez J, Carmichael D, Perälä M,

87 Hämäläinen E-R, Jarvinen M, Polarek J. Recombinant collagen and gelatin for

drug delivery. Elsevier; 2003 Nov 28;55(12):1547–1567. Available from:

https://www.sciencedirect.com/science/article/pii/S0169409X03001819

133. Yang C, Hillas PJ, Baez JA, Nokelainen M, Balan J, Tang J, Spiro R, Polarek JW.

The Application of Recombinant Human Collagen in Tissue Engineering. Springer

International Publishing; 2004;18(2):103–119. Available from:

http://link.springer.com/10.2165/00063030-200418020-00004

134. Rathinamoorthy R, Sasikala L. Polysaccharide Fibers in Wound Management.

2011;3(3):38–44. Available from:

https://innovareacademics.in/journal/ijpps/Vol3Issue3/2282.pdf

135. Lee KY, Mooney DJ. Alginate: Properties and biomedical applications. Pergamon;

2012 Jan 1;37(1):106–126. Available from:

https://www.sciencedirect.com/science/article/pii/S0079670011000918

136. Knill C., Kennedy J., Mistry J, Miraftab M, Smart G, Groocock M., Williams H.

Alginate fibres modified with unhydrolysed and hydrolysed chitosans for wound

dressings. Elsevier; 2004 Jan 1;55(1):65–76. Available from:

https://www.sciencedirect.com/science/article/pii/S0144861703002285

137. Thomas S. Alginate dressings in surgery and wound management — part 1. MA

Healthcare London ; 2000 Feb 29;9(2):56–60. Available from:

http://www.magonlinelibrary.com/doi/10.12968/jowc.2000.9.2.26338

138. Segal HC, Hunt BJ, Gilding K. The Effects of Alginate and Non-Alginate Wound

Dressings on Blood Coagulation and Platelet Activation. Sage PublicationsSage

CA: Thousand Oaks, CA; 1998 Jan 1;12(3):249–257. Available from:

88 http://journals.sagepub.com/doi/10.1177/088532829801200305

139. Matthew IR, Browne RM, Frame JW, Millar BG. Alginate fiber dressing for oral

mucosal wounds. Elsevier; 1994 May 1;77(5):456–460. Available from:

http://linkinghub.elsevier.com/retrieve/pii/0030422094902232

140. Matthew IR, Browne RM, Frame JW, Millar BG. Tissue response to a haemostatic

alginate wound dressing in tooth extraction sockets. Elsevier; 1993 Jun

1;31(3):165–9. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8512910

PMID: 8512910

141. Paul W, Sharma C. Chitosan and alginate wound dressings: a short review.

2004;18(1):18–23. Available from: http://medind.nic.in/taa/t04/i1/taat04i1p18.pdf

142. Li D, Li P, Zang J, Liu J. Enhanced hemostatic performance of tranexamic acid-

loaded chitosan/alginate composite microparticles. 2012;2012. PMID: 23193369

143. Rong J, Liang M, Xuan F, Sun J, Zhao L, Zhen H, Tian X, Liu D, Zhang Q, Peng

C, Yao T, Li F, Wang X, Han Y, Yu W. Alginate-calcium microsphere loaded

with thrombin: A new composite biomaterial for hemostatic embolization.

Elsevier; 2015 Apr 1;75:479–488. Available from:

https://www.sciencedirect.com/science/article/pii/S0141813015000033

144. Qin Y. Alginate fibers: an overwiew of the production processes and applications

in wound management. 2008;57(April):171–180. Available from:

http://onlinelibrary.wiley.com/doi/10.1002/pi.2098/full PMID: 19327838

145. Thu H-E, Zulfakar MH, Ng S-F. Alginate based bilayer hydrocolloid films as

potential slow-release modern wound dressing. Elsevier; 2012 Sep 15;434(1–

2):375–383. Available from:

89 https://www.sciencedirect.com/science/article/pii/S0378517312005443

146. Mogoşanu GD, Grumezescu AM. Natural and synthetic polymers for wounds and

dressing. Elsevier; 2014 Mar 25;463(2):127–136. Available from:

https://www.sciencedirect.com/science/article/pii/S0378517313010831

147. Balakrishnan B, Mohanty M, Umashankar PR, Jayakrishnan A. Evaluation of an in

situ forming hydrogel wound dressing based on oxidized alginate and gelatin.

Elsevier; 2005 Nov 1;26(32):6335–6342. Available from:

https://www.sciencedirect.com/science/article/pii/S0142961205003194

148. Kumar Dutta P, Dutta J, Tripathi VS. Chitin and chitosan: Chemistry, properties

and applications [Internet]. 2004. Available from:

http://nopr.niscair.res.in/bitstream/123456789/5397/1/JSIR 63%281%29 20-31.pdf

149. Malette WG, Quigley HJ, Gaines RD, Johnson ND, Rainer WG. Chitosan: A New

Hemostatic. Elsevier; 1983 Jul 1;36(1):55–58. Available from:

http://linkinghub.elsevier.com/retrieve/pii/S0003497510606492

150. Bennett BL, Littlejohn L. Review of New Topical Hemostatic Dressings for

Combat Casualty Care. 2014;179(5):497–514. Available from:

http://publications.amsus.org/doi/abs/10.7205/MILMED-D-13-00199 PMID:

24806495

151. Chan LW, Kim CH, Wang X, Pun SH, White NJ, Kim TH. PolySTAT-modified

chitosan gauzes for improved hemostasis in external hemorrhage. Acta Materialia

Inc.; 2016;31:178–185. Available from:

http://dx.doi.org/10.1016/j.actbio.2015.11.017 PMID: 26593785

152. Hattori H, Amano Y, Nogami Y, Takase B, Ishihara M. Hemostasis for Severe

90 Hemorrhage with Photocrosslinkable Chitosan Hydrogel and Calcium Alginate.

Springer US; 2010 Dec 9;38(12):3724–3732. Available from:

http://link.springer.com/10.1007/s10439-010-0121-4

153. Whang HS, Kirsch W, Zhu YH, Yang CZ, Hudson SM. Hemostatic Agents

Derived from Chitin and Chitosan. Taylor & Francis Group ; 2005 Oct;45(4):309–

323. Available from:

http://www.tandfonline.com/doi/abs/10.1080/15321790500304122

154. Okamoto Y, Yano R, Miyatake K, Tomohiro I, Shigemasa Y, Minami S. Effects of

chitin and chitosan on blood coagulation. Elsevier; 2003 Aug 15;53(3):337–342.

Available from:

https://www.sciencedirect.com/science/article/pii/S0144861703000766

155. Chou T-C, Fu E, Wu C-J, Yeh J-H. Chitosan enhances platelet adhesion and

aggregation. Academic Press; 2003 Mar 14;302(3):480–483. Available from:

https://www.sciencedirect.com/science/article/pii/S0006291X03001736

156. Benesch J, Tengvall P. Blood protein adsorption onto chitosan. Elsevier; 2002 Jun

1;23(12):2561–2568. Available from:

https://www.sciencedirect.com/science/article/pii/S014296120100391X

157. Pusateri AE, McCarthy SJ, Gregory KW, Harris RA, Cardenas L, McManus AT,

Goodwin CW. Effect of a chitosan-based hemostatic dressing on blood loss and

survival in a model of severe venous hemorrhage and hepatic injury in swine.

2003;54(1):177–182. PMID: 12544915

158. Wedmore I, McManus JG, Pusateri AE, Holcomb JB. A Special Report on the

Chitosan-based Hemostatic Dressing: Experience in Current Combat Operations.

91 2006 Mar;60(3):655–658. Available from:

https://insights.ovid.com/crossref?an=00005373-200603000-00030

159. Brown MA, Daya MR, Worley JA. Experience with Chitosan Dressings in a

Civilian EMS System. Elsevier; 2009 Jul 1;37(1):1–7. Available from:

https://www.sciencedirect.com/science/article/pii/S0736467907004775

160. Sohn VY, Eckert MJ, Martin MJ, Arthurs ZM, Perry JR, Beekley A, Rubel EJ,

Adams RP, Bickett GL, Rush RM. Efficacy of Three Topical Hemostatic Agents

Applied by Medics in a Lethal Groin Injury Model. Academic Press; 2009 Jun

15;154(2):258–261. Available from:

https://www.sciencedirect.com/science/article/pii/S0022480408005222

161. Kheirabadi BS, Edens JW, Terrazas IB, Estep JS, Klemcke HG, Dubick MA,

Holcomb JB. Comparison of New Hemostatic Granules/Powders With Currently

Deployed Hemostatic Products in a Lethal Model of Extremity Arterial

Hemorrhage in Swine. 2009 Feb;66(2):316–328. Available from:

https://insights.ovid.com/crossref?an=00005373-200902000-00005

162. Gustafson SB, Fulkerson P, Bildfell R, Aguilera L, Hazzard TM. Chitosan

Dressing Provides Hemostasis in Swine Femoral Arterial Injury Model. Taylor &

Francis; 2007 Jan 2;11(2):172–178. Available from:

http://www.tandfonline.com/doi/full/10.1080/10903120701205893

163. Millner RWJ, Lockhart AS, Bird H, Alexiou C. A New Hemostatic Agent: Initial

Life-Saving Experience With Celox (Chitosan) in .

Elsevier; 2009 Feb 1;87(2):e13–e14. Available from:

https://www.sciencedirect.com/science/article/pii/S0003497508020420

92 164. Pozza M, Millner RWJ. Celox (chitosan) for haemostasis in massive traumatic

bleeding. 2011 Feb;18(1):31–33. Available from:

https://insights.ovid.com/crossref?an=00063110-201102000-00007

165. Quayle JM, Thomas GOR. A pre-hospital technique for controlling haemorrhage

from traumatic perineal and high amputation injuries. British Medical Journal

Publishing Group; 2011 Dec 1;157(4):419–20. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/22319995 PMID: 22319995

166. Nie W, Yuan X, Zhao J, Zhou Y, Bao H. Rapidly in situ forming chitosan/ε-

polylysine hydrogels for adhesive sealants and hemostatic materials. Elsevier;

2013 Jul 1;96(1):342–348. Available from:

https://www.sciencedirect.com/science/article/pii/S0144861713003615

167. Dowling MB, Smith W, Balogh P, Duggan MJ, MacIntire IC, Harris E, Mesar T,

Raghavan SR, King DR. Hydrophobically-modified chitosan foam: description

and hemostatic efficacy. Academic Press; 2015 Jan 1;193(1):316–323. Available

from: https://www.sciencedirect.com/science/article/pii/S0022480414005691

168. Mathias J-D, Tessier-Doyen N, Michaud P. Development of a Chitosan-Based

Biofoam: Application to the Processing of a Porous Ceramic Material. Molecular

Diversity Preservation International; 2011 Feb 16;12(2):1175–1186. Available

from: http://www.mdpi.com/1422-0067/12/2/1175

169. Testouri A, Honorez C, Barillec A, Langevin D, Drenckhan W. Highly Structured

Foams from Chitosan Gels. American Chemical Society; 2010 Jul

27;43(14):6166–6173. Available from:

http://pubs.acs.org/doi/abs/10.1021/ma100819j

93 170. Vournakis JN, Demcheva M, Whitson a, Guirca R, Pariser ER. Isolation,

purification, and characterization of poly-N-acetyl glucosamine use as a

hemostatic agent. 2004;57(1, S):S2–S6. PMID: 15280743

171. Fischer TH, Bode AP, Demcheva M, Vournakis JN. Hemostatic properties of

glucosamine-based materials. 2007 Jan;80A(1):167–174. Available from:

http://doi.wiley.com/10.1002/jbm.a.30877

172. Connolly RJ. Application of the Poly-N-Acetyl Glucosamine-Derived Rapid

Deployment Hemostat Trauma Dressing in Severe/Lethal Swine Hemorrhage

Trauma Models. 2004 Jul;57(Supplement):S26–S28. Available from:

https://insights.ovid.com/crossref?an=00005373-200407001-00006

173. Jewelewicz DD, Cohn SM, Crookes BA, Proctor KG. Modified Rapid Deployment

Hemostat Bandage Reduces Blood Loss and Mortality in Coagulopathic Pigs with

Severe Liver Injury. 2003 Aug;55(2):275–281. Available from:

https://insights.ovid.com/crossref?an=00005373-200308000-00011

174. King DR, Cohn SM, Proctor KG. Modified Rapid Deployment Hemostat Bandage

Terminates Bleeding in Coagulopathic Patients with Severe Visceral Injuries. 2004

Oct;57(4):756–759. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=00

005373-200410000-00010

175. Ersoy G, Kaynak MF, Yilmaz O, Rodoplu U, Maltepe F, Gokmen N. Hemostatic

effects of Microporous Polysaccharide Hemosphere® in a rat model with severe

femoral artery bleeding. Springer Healthcare Communications; 2007

May;24(3):485–492. Available from:

94 http://link.springer.com/10.1007/BF02848770

176. Burgert JM, Gegel BT, Austin R, Davila A, Deeds J, Hodges L, Hover A,

Lockhart C, Roy J, Simpson G, Weaver S, Wolfe W, Johnson D. Effects of arterial

blood pressure on rebleeding using Celox and TraumaDEX in a porcine model of

lethal femoral injury. 2010 Jun;78(3):230–6. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/20572410 PMID: 20572410

177. Gegel BT, Burgert JM, Lockhart C, Austin R, Davila A, Deeds J, Hodges L, Hover

A, Roy J, Simpson G, Weaver S, Wolfe W, Johnson D. Effects of Celox and

TraumaDEX on hemorrhage control in a porcine model. 2010 Apr;78(2):115–20.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/20583456 PMID: 20583456

178. Alam HB, Chen Z, Jaskille A, Querol RILC, Koustova E, Inocencio R, Conran R,

Seufert A, Ariaban N, Toruno K, Rhee P. Application of a Zeolite Hemostatic

Agent Achieves 100% Survival in a Lethal Model of Complex Groin Injury in

Swine. 2004 May;56(5):974–983. Available from:

https://insights.ovid.com/crossref?an=00005373-200405000-00007

179. Pusateri AE, Delgado A V., Dick EJ, Martinez RS, Holcomb JB, Ryan KL.

Application of a Granular Mineral-Based Hemostatic Agent (QuikClot) to Reduce

Blood Loss After Grade V Liver Injury in Swine. 2004 Sep;57(3):555–562.

Available from: https://insights.ovid.com/crossref?an=00005373-200409000-

00014

180. Inaba K, Rhee P, Teixeira PG, Barmparas G, Putty B, Branco BC, Cohn S,

Demetriades D. Intracorporeal use of advanced local hemostatics in a damage

control swine model of grade IV liver injury. 2011 Nov;71(5):1312–8. Available

95 from: https://insights.ovid.com/crossref?an=00005373-201111000-00032 PMID:

22002611

181. Li J, Cao W, Lv X, Jiang L, Li Y, Li W, Chen S, Li X. Zeolite-based hemostat

QuikClot releases calcium into blood and promotes blood coagulation in vitro.

Nature Publishing Group; 2013 Mar 21;34(3):367–372. Available from:

http://www.nature.com/articles/aps2012159

182. Rhee P, Brown C, Martin M, Salim A, Plurad D, Green D, Chambers L,

Demetriades D, Velmahos G, Alam H. QuikClot Use in Trauma for Hemorrhage

Control: Case Series of 103 Documented Uses. 2008 Apr;64(4):1093–1099.

Available from: https://insights.ovid.com/crossref?an=00005373-200804000-

00034

183. Gegel BT, Austin PN, Johnson AD. An evidence-based review of the use of a

combat gauze (QuikClot) for hemorrhage control. 2013 Dec;81(6):453–8.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/24597007 PMID: 24597007

184. MARGULIS V, MATSUMOTO ED, SVATEK R, KABBANI W, CADEDDU

JA, LOTAN Y. APPLICATION OF NOVEL HEMOSTATIC AGENT DURING

LAPAROSCOPIC PARTIAL NEPHRECTOMY. Elsevier; 2005 Aug

1;174(2):761–764. Available from:

https://www.sciencedirect.com/science/article/pii/S0022534701683873

185. Johnson D, Bates S, Nukalo S, Staub A, Hines A, Leishman T, Michel J, Sikes D,

Gegel B, Burgert J. The effects of QuikClot Combat Gauze on hemorrhage control

in the presence of hemodilution and hypothermia. 2014;3(2):21–25. PMID:

25568780

96 186. Wright JK, Kalns J, Wolf EA, Traweek F, Schwarz S, Loeffler CK, Snyder W,

Yantis LD, Eggers J. Thermal Injury Resulting from Application of a Granular

Mineral Hemostatic Agent. 2004 Aug;57(2):224–230. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=00

005373-200408000-00004

187. McManus J, Hurtado T, Pusateri A, Knoop KJ. A case series describing thermal

injury resulting from zeolite use for hemorrhage control in combat operations.

2007 Jan 2;11(1):67–71. Available from:

http://www.tandfonline.com/doi/full/10.1080/10903120601021176 PMID:

17169881

188. Ostomel TA, Stoimenov PK, Holden PA, Alam HB, Stucky GD. Host-guest

composites for induced hemostasis and therapeutic healing in traumatic injuries.

Kluwer Academic Publishers; 2006 Aug;22(1):55–67. Available from:

http://link.springer.com/10.1007/s11239-006-7658-y

189. Li Y, Li H, Xiao L, Zhou L, Shentu J, Zhang X, Fan J, Li Y, Li H, Xiao L, Zhou L,

Shentu J, Zhang X, Fan J. Hemostatic Efficiency and Wound Healing Properties of

Natural Zeolite Granules in a Lethal Rabbit Model of Complex Groin Injury.

Multidisciplinary Digital Publishing Institute; 2012 Dec 3;5(12):2586–2596.

Available from: http://www.mdpi.com/1996-1944/5/12/2586

190. Ostomel TA, Shi Q, Stoimenov PK, Stucky GD. Metal Oxide Surface Charge

Mediated Hemostasis. American Chemical Society; 2007;23(22):11233–11238.

Available from: https://pubs.acs.org/doi/abs/10.1021/la701281t

191. Arnaud F, Tomori T, Carr W, McKeague A, Teranishi K, Prusaczyk K, McCarron

97 R. Exothermic Reaction in Zeolite Hemostatic Dressings: QuikClot ACS and

ACS+®. Springer US; 2008 Oct 20;36(10):1708–1713. Available from:

http://link.springer.com/10.1007/s10439-008-9543-7

192. Gegel B, Burgert J, Gasko J, Campbell C, Martens M, Keck J, Reynolds H,

Loughren M, Johnson D. The Effects of QuikClot Combat Gauze and Movement

on Hemorrhage Control in a Porcine Model. Oxford University Press; 2012 Dec

1;177(12):1543–1547. Available from:

https://academic.oup.com/milmed/article/177/12/1543-1547/4336763

193. Kheirabadi BS, Scherer MR, Estep JS, Dubick MA, Holcomb JB. Determination

of Efficacy of New Hemostatic Dressings in a Model of Extremity Arterial

Hemorrhage in Swine. 2009 Sep;67(3):450–460. Available from:

https://insights.ovid.com/crossref?an=00005373-200909000-00006

194. Arnaud F, Teranishi K, Okada T, Parreño-Sacdalan D, Hupalo D, McNamee G,

Carr W, Burris D, McCarron R. Comparison of Combat Gauze and TraumaStat in

Two Severe Groin Injury Models. Academic Press; 2011 Jul 1;169(1):92–98.

Available from:

https://www.sciencedirect.com/science/article/pii/S0022480409004661

195. Arnaud F, Tomori T, Saito R, McKeague A, Prusaczyk WK, McCarron RM.

Comparative Efficacy of Granular and Bagged Formulations of the Hemostatic

Agent QuikClot. 2007 Oct;63(4):775–782. Available from:

https://insights.ovid.com/crossref?an=00005373-200710000-00009

196. Ward KR, Tiba MH, Holbert WH, Blocher CR, Draucker GT, Proffitt EK, Bowlin

GL, Ivatury RR, Diegelmann RF. Comparison of a New Hemostatic Agent to

98 Current Combat Hemostatic Agents in a Swine Model of Lethal Extremity Arterial

Hemorrhage. 2007 Aug;63(2):276–284. Available from:

https://insights.ovid.com/crossref?an=00005373-200708000-00005

197. Carraway JW, Kent D, Young K, Cole A, Friedman R, Ward KR. Comparison of a

new mineral based hemostatic agent to a commercially available granular zeolite

agent for hemostasis in a swine model of lethal extremity arterial hemorrhage.

Elsevier; 2008 Aug 1;78(2):230–235. Available from:

https://www.sciencedirect.com/science/article/pii/S0300957208001305

198. Arnaud F, Teranishi K, Tomori T, Carr W, McCarron R. Comparison of 10

hemostatic dressings in a groin puncture model in swine. Mosby; 2009 Sep

1;50(3):632–639.e1. Available from:

https://www.sciencedirect.com/science/article/pii/S0741521409012622

199. Kheirabadi BS, Mace JE, Terrazas IB, Fedyk CG, Estep JS, Dubick MA,

Blackbourne LH. Safety Evaluation of New Hemostatic Agents, Smectite

Granules, and Kaolin-Coated Gauze in a Vascular Injury Wound Model in Swine.

2010 Feb;68(2):269–278. Available from:

https://insights.ovid.com/crossref?an=00005373-201002000-00003

200. Ginsburg I, de Vries A, Katchalski E. The Action of Some Water-Soluble Poly-a-

Amino Acids on Fibrinolysis. American Association for the Advancement of

Science; 1952 Jul 4;116(3001):15–6. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17815666 PMID: 17815666

201. Wang R, Zhou B, Liu W, Feng X, Li S, Yu D, Chang J, Chi B, Xu H. Fast in situ

generated ɛ-polylysine-poly (ethylene glycol) hydrogels as tissue adhesives and

99 hemostatic materials using an enzyme-catalyzed method. SAGE PublicationsSage

UK: London, England; 2015 Mar 2;29(8):1167–1179. Available from:

http://journals.sagepub.com/doi/10.1177/0885328214553960

202. Pedersen AH, Lund-Hansen T, Bisgaard-Frantzen H, Olsen F, Petersen LC.

Autoactivation of human recombinant coagulation factor VII. 1989

Nov;28(24):9331–9336. Available from:

http://pubs.acs.org/doi/abs/10.1021/bi00450a013

203. Fischer D, Li Y, Ahlemeyer B, Krieglstein J, Kissel T. In vitro cytotoxicity testing

of polycations: influence of polymer structure on cell viability and hemolysis.

Elsevier; 2003 Mar 1;24(7):1121–1131. Available from:

https://www.sciencedirect.com/science/article/pii/S0142961202004453

204. Wang MY, Armstrong JK, Fisher TC, Meiselman HJ, McComb GJ, Levy ML. A

New, Pluronic-based, Bone Hemostatic Agent That Does Not Impair Osteogenesis.

Oxford University Press; 2001 Oct 1;49(4):962–968. Available from:

https://academic.oup.com/neurosurgery/article/49/4/962/3772045

205. Wellisz T, Armstrong JK, Cambridge J, Fisher TC. Ostene, a new water-soluble

bone hemostasis agent. 2006 May;17(3):420–5. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/16770175 PMID: 16770175

206. Henderson PW, Kadouch DJM, Singh SP, Zawaneh PN, Weiser J, Yazdi S,

Weinstein A, Krotscheck U, Wechsler B, Putnam D, Spector JA. A rapidly

resorbable hemostatic biomaterial based on dihydroxyacetone. 2009;9999A:NA-

NA. Available from: http://doi.wiley.com/10.1002/jbm.a.32586

207. Wallace DG, Cruise GM, Rhee WM, Schroeder JA, Prior JJ, Ju J, Maroney M,

100 Duronio J, Ngo MH, Estridge T, Coker GC. A tissue sealant based on reactive

multifunctional polyethylene glycol. Wiley-Blackwell; 2001 Jan 1;58(5):545–555.

Available from: http://doi.wiley.com/10.1002/jbm.1053

208. Hill A, Estridge TD, Maroney M, Monnet E, Egbert B, Cruise G, Coker GT.

Treatment of suture line bleeding with a novel synthetic surgical sealant in a

canine iliac PTFE graft model. Wiley-Blackwell; 2001 Jan 1;58(3):308–312.

Available from: http://doi.wiley.com/10.1002/1097-

4636%282001%2958%3A3%3C308%3A%3AAID-JBM1022%3E3.0.CO%3B2-P

209. Buskens E, Meijboom MJ, Kooijman H, Van Hout BA. The use of a surgical

sealant (CoSeal ® ) in cardiac and vascular reconstructive surgery: an economic

analysis [Internet]. 2006. Available from:

https://pdfs.semanticscholar.org/2d9a/6cbbc75294564117f2013964658f7d239ebf.

pdf

210. Seifman BD, Rubin MA, Williams AL, Wolf JS. Use Of Absorbable

Cyanoacrylate Glue To Repair An Open Cystotomy. Elsevier; 2002 Apr

1;167(4):1872–1875. Available from:

https://www.sciencedirect.com/science/article/pii/S0022534705652524

211. Montanaro L, Arciola C., Cenni E, Ciapetti G, Savioli F, Filippini F, Barsanti L.

Cytotoxicity, blood compatibility and antimicrobial activity of two cyanoacrylate

glues for surgical use. Elsevier; 2000 Jan 1;22(1):59–66. Available from:

https://www.sciencedirect.com/science/article/pii/S0142961200001630

212. Al-Belasy FA, Amer MZ. Hemostatic effect of n-butyl-2-cyanoacrylate

(histoacryl) glue in warfarin-treated patients undergoing oral surgery. W.B.

101 Saunders; 2003 Dec 1;61(12):1405–1409. Available from:

https://www.sciencedirect.com/science/article/pii/S0278239103008395

213. Mehdizadeh M, Yang J. Design Strategies and Applications of Tissue

Bioadhesives. Wiley-Blackwell; 2013 Mar 1;13(3):271–288. Available from:

http://doi.wiley.com/10.1002/mabi.201200332

214. Singer AJ, McClain SA, Katz A. A Porcine Epistaxis Model: Hemostatic Effects

of Octylcyanoacrylate. SAGE PublicationsSage CA: Los Angeles, CA; 2004 May

17;130(5):553–557. Available from:

http://journals.sagepub.com/doi/10.1016/j.otohns.2003.09.035

215. Fan Y, Sun H, Pei G, Ruan C. Haemostatic efficacy of an ethyl-2-cyanoacrylate-

based aerosol in combination with tourniquet application in a large wound model

with an arterial injury. Elsevier; 2008 Jan 1;39(1):61–66. Available from:

https://www.sciencedirect.com/science/article/pii/S0020138307004196

216. Golling M, Schaudt A, Mehrabi A, Mood ZA, Bechstein WO. Clinical application

of soft polyglycolic acid felt for hemostasis and repair of a lacerated liver: Report

of two cases. Springer Japan; 2008 Feb 1;38(2):188–192. Available from:

http://link.springer.com/10.1007/s00595-007-3587-4

217. Horák D, Švec F, Isakov Y, Polyaev Y, Adamyan A, Konstantinov K, Shafranov

V, Voronkova O, Nikanorov A, Trostenyuk N. Use of poly(2-hydroxyethyl

methacrylate) for endovascular occlusion in . Elsevier; 1992 Jan

1;9(1):43–48. Available from:

https://www.sciencedirect.com/science/article/pii/026766059290009I

218. Antonik LM, Khabibulina AG, Fadeeva T V., Kogan AS, Malyshev A V.

102 Hemostatic complexes based on polyacrylic acid. Springer US; 2007

May;41(5):278–280. Available from: http://link.springer.com/10.1007/s11094-

007-0062-x

219. Carr M, Kenawy E, Layman J, Wnek G, Ward K. Development Of The

Biohemostat-A Treatment Modality For High Pressure Bleeding Based On Super

Absorbent Polymers. 2002;17:54. Available from:

https://journals.lww.com/shockjournal/Citation/2002/06001/DEVELOPMENT_O

F_THE_BIOHEMOSTAT___A_TREATMENT.163.aspx

220. Casey BJ, Behrens AM, Hess JR, Wu ZJ, Griffith BP, Kofinas P. FVII Dependent

Coagulation Activation in Citrated Plasma by Polymer Hydrogels. American

Chemical Society; 2010 Dec 13;11(12):3248–3255. Available from:

http://pubs.acs.org/doi/abs/10.1021/bm101147w

221. Pasic M, Unbehaun A, Drews T, Hetzer R. Late wound healing problems after use

of BioGlue for apical hemostasis during transapical aortic valve implantation.

2011;13(5):532–4. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/21807815 PMID: 21807815

222. Mehdizadeh M, Weng H, Gyawali D, Tang L, Yang J. Injectable citrate-based

mussel-inspired tissue bioadhesives with high wet strength for sutureless wound

closure. Elsevier; 2012 Nov 1;33(32):7972–7983. Available from:

https://www.sciencedirect.com/science/article/pii/S0142961212008502

223. Ji Y, Ji T, Liang K, Zhu L. Mussel-inspired soft-tissue adhesive based on poly(diol

citrate) with catechol functionality. Springer US; 2016 Feb 24;27(2):30. Available

from: http://link.springer.com/10.1007/s10856-015-5649-2

103 224. Zhang H, Bré LP, Zhao T, Zheng Y, Newland B, Wang W. Mussel-inspired

hyperbranched poly(amino ester) polymer as strong wet tissue adhesive. Elsevier;

2014 Jan 1;35(2):711–719. Available from:

https://www.sciencedirect.com/science/article/pii/S0142961213012283

225. Ye Z, Zhang H, Luo H, Wang S, Zhou Q, DU X, Tang C, Chen L, Liu J, Shi Y-K,

Zhang E-Y, Ellis-Behnke R, Zhao X. Temperature and pH effects on biophysical

and morphological properties of self-assembling peptide RADA16-I. Wiley-

Blackwell; 2008 Feb 1;14(2):152–162. Available from:

http://doi.wiley.com/10.1002/psc.988

226. Wang T, Zhong X, Wang S, Lv F, Zhao X, Wang T, Zhong X, Wang S, Lv F,

Zhao X. Molecular Mechanisms of RADA16-1 Peptide on Fast Stop Bleeding in

Rat Models. Multidisciplinary Digital Publishing Institute; 2012 Nov

19;13(12):15279–15290. Available from: http://www.mdpi.com/1422-

0067/13/11/15279

227. Cormier AR, Pang X, Zimmerman MI, Zhou H-X, Paravastu AK. Molecular

Structure of RADA16-I Designer Self-Assembling Peptide Nanofibers. American

Chemical Society; 2013 Sep 24;7(9):7562–7572. Available from:

http://pubs.acs.org/doi/10.1021/nn401562f

228. Hsu BB, Conway W, Tschabrunn CM, Mehta M, Perez-Cuevas MB, Zhang S,

Hammond PT. Clotting Mimicry from Robust Hemostatic Bandages Based on

Self-Assembling Peptides. American Chemical Society; 2015 Sep 22;9(9):9394–

9406. Available from: http://pubs.acs.org/doi/10.1021/acsnano.5b02374

229. Tyan Y-C, Liao* J-D, Lin S-P. Surface properties and in vitro analyses of

104 immobilized chitosan onto polypropylene non-woven fabric surface using antenna-

coupling microwave plasma. Kluwer Academic Publishers; 2003;14(9):775–781.

Available from: http://link.springer.com/10.1023/A:1025036421604

230. Lin W-C, Tseng C-H, Yang M-C. In-Vitro Hemocompatibility Evaluation of a

Thermoplastic Polyurethane Membrane with Surface-Immobilized Water-Soluble

Chitosan and Heparin. Wiley-Blackwell; 2005 Oct 20;5(10):1013–1021. Available

from: http://doi.wiley.com/10.1002/mabi.200500077

231. Zahedi P, Rezaeian I, Ranaei-Siadat S-O, Jafari S-H, Supaphol P. A review on

wound dressings with an emphasis on electrospun nanofibrous polymeric

bandages. Wiley-Blackwell; 2009 Feb 1;21(2):77–95. Available from:

http://doi.wiley.com/10.1002/pat.1625

232. Blackbourne LH, Baer DG, Eastridge BJ, Kheirabadi B, Kragh JF, Cap AP,

Dubick M a., Morrison JJ, Midwinter MJ, Butler FK, Kotwal RS, Holcomb JB.

Military medical revolution. 2012;73:S372–S377. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201212005-00002 PMID: 23192060

233. Naimer SA. A Review of Methods to Control Bleeding from Life-Threatening

Traumatic Wounds. 2014;6:479–490. Available from:

http://www.scirp.org/journal/healthhttp://dx.doi.org/10.4236/health.2014.66067htt

p://creativecommons.org/licenses/by/4.0/

234. Naimer SA, Anat N, Katif G. Evaluation of techniques for treating the bleeding

wound. Elsevier; 2004 Oct 1;35(10):974–979. Available from:

https://www.sciencedirect.com/science/article/pii/S0020138303003164

105 235. Shipman N, Lessard CS. Pressure Applied by the Emergency/Israeli Bandage.

Oxford University Press; 2009 Jan 1;174(1):086–092. Available from:

https://academic.oup.com/milmed/article/174/1/86-92/4333642

236. Naimer SA, Chemla F. Elastic adhesive dressing treatment of bleeding wounds in

trauma victims. W.B. Saunders; 2000 Nov 1;18(7):816–819. Available from:

https://www.sciencedirect.com/science/article/pii/S0735675700158061

237. Hess JR, Brohi K, Dutton RP, Hauser CJ, Holcomb JB, Kluger Y, Mackway-Jones

K, Parr MJ, Rizoli SB, Yukioka T, Hoyt DB, Bouillon B. The Coagulopathy of

Trauma: A Review of Mechanisms. 2008 Oct;65(4):748–754. Available from:

https://insights.ovid.com/crossref?an=00005373-200810000-00002

238. Lashof-Sullivan M, Shoffstall A, Lavik E. Intravenous hemostats: challenges in

translation to patients. 2013;5(22):10719–28. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/24088870 PMID: 24088870

239. Lomas-Niera JL, Perl M, Chung C-S, Ayala A. SHOCK AND HEMORRHAGE:

AN OVERVIEW OF ANIMAL MODELS. 2005 Dec;24(Supplement 1):33–39.

Available from: https://insights.ovid.com/crossref?an=00024382-200512001-

00006

240. Satterly S, Nelson D, Zwintscher N, Oguntoye M, Causey W, Theis B, Huang R,

Haque M, Martin M, Bickett G, Rush RM. Hemostasis in a Noncompressible

Hemorrhage Model: An End-User Evaluation of Hemostatic Agents in a Proximal

Arterial Injury. Elsevier; 2013 Mar 1;70(2):206–211. Available from:

https://www.sciencedirect.com/science/article/pii/S1931720412002887

241. Englehart MS, Cho SD, Tieu BH, Morris MS, Underwood SJ, Karahan A, Muller

106 PJ, Differding JA, Farrell DH, Schreiber MA. A Novel Highly Porous Silica and

Chitosan-Based Hemostatic Dressing Is Superior to HemCon and Gauze Sponges.

2008 Oct;65(4):884–892. Available from:

https://insights.ovid.com/crossref?an=00005373-200810000-00024

242. Sambasivan CN, Cho SD, Zink KA, Differding JA, Schreiber MA. A highly

porous silica and chitosan-based hemostatic dressing is superior in controlling

hemorrhage in a severe groin injury model in swine. Elsevier; 2009 May

1;197(5):576–580. Available from:

https://www.sciencedirect.com/science/article/pii/S0002961009000725

243. Kheirabadi BS, Klemcke HG. Symp. Paper RTO‐MP‐HFM‐109. 2004. p. 20.

244. Mueller GR, Pineda TJ, Xie HX, Teach JS, Barofsky AD, Schmid JR, Gregory

KW. A novel sponge-based wound stasis dressing to treat lethal noncompressible

hemorrhage. 2012 Aug;73:S134–S139. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201208001-00024

245. Kragh JF, Aden JK, Steinbaugh J, Bullard M, Dubick MA. Gauze vs XSTAT in

wound packing for hemorrhage control. W.B. Saunders; 2015 Jul 1;33(7):974–

976. Available from:

https://www.sciencedirect.com/science/article/pii/S0735675715001941?via%3Dih

ub

246. Duggan M, Rago A, Sharma U, Zugates G, Freyman T, Busold R, Caulkins J,

Pham Q, Chang Y, Mejaddam A, Beagle J, Velmahos G, deMoya M, Zukerberg L,

Ng TF, King LDR. Self-expanding polyurethane polymer improves survival in a

107 model of noncompressible massive abdominal hemorrhage. 2013 Jun;74(6):1462–

1467. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201306000-00012

247. Peev MP, Rago A, Hwabejire JO, Duggan MJ, Beagle J, Marini J, Zugates G,

Busold R, Freyman T, Velmahos GS, Demoya MA, Yeh DD, Fagenholz PJ,

Sharma U, King DR. Self-expanding foam for prehospital treatment of severe

intra-abdominal hemorrhage. 2014 Mar;76(3):619–624. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201403000-00008

248. Cresilon [Internet]. Available from: https://cresilon.com/

249. Simmons JW, Pittet J-F, Pierce B. Trauma-Induced Coagulopathy. 2014 Sep

1;4(3):189–199. Available from: http://link.springer.com/10.1007/s40140-014-

0063-8 PMID: 25587242

250. Cohen MJ. Acute traumatic coagulopathy: Clinical characterization and

mechanistic investigation. Pergamon; 2014 May 1;133:S25–S27. Available from:

https://www.sciencedirect.com/science/article/pii/S0049384814001388

251. Kheirabadi B. Evaluation of topical hemostatic agents for combat wound

treatment. 2011;25–37. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/21607904 PMID: 21607904

252. Rall JM, Cox JM, Songer AG, Cestero RF, Ross JD. Comparison of novel

hemostatic dressings with QuikClot combat gauze in a standardized swine model

of uncontrolled hemorrhage. 2013 Aug;75:S150–S156. Available from:

108 http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201308002-00009

109 Chapter 3: Intravenously Administered Biomaterials and Advanced Technologies for

Management of Traumatic Non-compressible Hemorrhage

Content based on: Hickman DA, Pawlowski CL, Sekhon UDS, Marks J, Gupta AS. Biomaterials and Advanced Technologies for Hemostatic Management of Bleeding. Adv Mater. 2017 Nov 22. doi: 10.1002/adma.201700859.

3.1 Introduction

As emphasized previously, bleeding can cause a variety of debilitating health issues, especially in traumatic and surgical settings. The majority of hemorrhage-related deaths in

the military and civilian populations occur from non-compressible and heavy intracavitary hemorrhage that cannot be completely managed by tourniquets, external dressings and bandages1,2. Certain dressings and bandages have undergone advancement in materials

components as well as technology designs to improve the treatment options for such heavy

bleeding, especially in intracavitary applications, as reviewed in the previous chapter.

However, intravenous fluid resuscitation and subsequent blood component transfusion

currently remains the primary option for improving survival of civilian and battlefield

casualties with heavy truncal and junctional hemorrhage and traumatic brain injuries, as

described previously3–5. Unfortunately, such procedures are often logistically challenging

in austere pre-hospital (e.g. first responder) conditions, as well as, in medical facilities

without access to blood components, due to the limitations of blood availability and

portability, lack of appropriately trained personnel, high contamination risk, short shelf-

life etc. Especially for platelets, the major hemostatic cellular component of blood, these

issues present severe functional and logistical barriers, and therefore the development of

intravenously transfusable semi-synthetic or synthetic platelet analogues that can provide

110 platelet’s hemostatic functions without its biologic issues and availability/accessibility limitations, has become an exciting area of biomaterials research. In parallel, synthetic mimicry of certain other non-cellular coagulatory components of blood have also gained momentum in recent years. All of these materials and technologies are to be intravenously administered and are currently in the pre-clinical evaluation stage, with anticipated translation not only for pre-hospital management of civilian and military trauma but also for transfusion applications in surgery and hematology/oncology areas in hospital settings.

3.2 Naturally derived intravascular pro-coagulant materials

As described in Chapter 1, overall hemostasis is a combined output of platelet plug formation (primary hemostasis) and coagulation cascade leading to crosslinked fibrin biopolymer formation (secondary hemostasis) at the bleeding site. Therefore, while one arm of materials research has focused on technologies that leverage and mimic platelet functions, the other arm has focused on various molecular aspects of modulating the initiation, propagation or stabilization of the coagulation cascade components and final fibrin clot. In modulating the molecular aspects of the coagulation cascade, the most obvious approach is to intravenously administer fibrinogen as well as coagulation factors.

In a bleeding patient, fibrinogen levels can be replenished by directly administering fibrinogen concentrate, or cryoprecipitate (a mixed product containing fibrinogen, factor

VIII, von Willebrand factor, and FXIII). These fibrinogen products have been extensively studied in pre-clinical animal models as well as in selected clinical trials and significant benefit of these products in hemostatic management of traumatic and surgical bleeding have been demonstrated6–9. Further clinical studies are needed to establish the short-term

111 and long-term systemic safety of such products. In the field of coagulation factors, FVIII

and FIX concentrates isolated from pooled donor plasma became commercially available

in the 1970s for the treatment of hemophilia, but were found to transmit the risk of viral

hepatitis10,11. In , the current clinical mainstay of factor replenishment

is the use of fresh frozen plasma (FFP), which is rich in multiple coagulation factors

(prothrombin or FII, FV, FVII, FIX and FX)12–14. However, FFP also suffers from issues

of availability, pathogenic contamination risks and immunologic side-effect risks10,15,16.

These issues have led to the research and development of a variety of coagulation factors

using recombinant technologies during the past two decades10,17–21. For example, following

the cloning of the human FVIII mRNA in 1984, a variety of recombinant FVIII proteins

have been developed, such as Kogenate® (Bayer), Recombinate® (Baxter), Refacto®

(Wyeth Pharmaceuticals), and ADVATE® (Baxter)22,23. All these recombinant FVIII preparations have demonstrated significant hemostatic efficacy in treating bleeding risks in patients with hemophilia A. Similarly, the cloning of FIX from a human liver cDNA library has led to development of recombinant FIX (e.g. BeneFix®, Wyeth

Pharmaceuticals) for bleeding risk treatment in haemophilia B patients24–26. It is important

to note here that the use of Factor VIII and Factor IX have been reported to induce

development of inhibitory alloantibodies in at least 30% of treated patients after multiple

prophylactic dosages, and this has led to the development of one of the most notable

recombinant coagulation factor products, the recombinant Factor VIIa (FVIIa)17,18. As was

depicted in Figure 1-2, this coagulation factor can interact with Tissue Factor to generate

moderate amounts of thrombin (Factor IIa) in the extrinsic pathway of coagulation cascade.

Factor VIIa (and thrombin) can also reportedly activate FIX to FIXa, which subsequently

112 takes part in activation and complexation with other coagulation factors (FXa and FVIIIa)

on the surface of activated platelet membrane in presence of Ca++ to form the

prothrombinase (FXa + FVa + FII) complex, further amplifying the generation of thrombin

(Figure 1-2) via the common pathway of coagulation27,28. Thrombin further activates

localized platelets and also breaks down fibrinogen to form fibrin, the final coagulation

cascade product. Therefore, the central theme in utilization of recombinant or plasma- derived FVIIa to improve hemostasis is to augment the generation of thrombin while bypassing the tenase amplification step, and this has proven highly beneficial in hemostatic treatment of patients with congenital coagulation factor defects, hematologic malignancies, thrombocytopenia and traumatic bleeding. FVII was originally cloned in the late 1980s from human liver and Hep G2 cells and has subsequently led to the development of the recombinant FVIIa product NovoSeven® (Novo Nordisk)29–31. This product was originally

approved clinically to treat hemophilia patients where lack of FVIII (Hemophilia A) or FIX

(Hemophilia B) affects the intrinsic pathway of thrombin generation, and therefore

intravenous administration of FVIIa was postulated to compensate for this by generating

thrombin via the tissue factor (extrinsic) pathway32,33. NovoSeven® has also been recently

approved for use in patients with Glanzmann Thrombasthenia where patient’s platelets

have qualitative or quantitative deficiencies that affect their ability to bind to fibrinogen

via integrin GPIIb/IIIa. For patients with thrombocytopenia (platelet deficiencies), it has

been demonstrated that combining FVIIa therapy with platelet transfusion may provide

additive benefit, since the common pathway of coagulation cascade that leads to amplified

thrombin generation and fibrin formation is facilitated by the presence of activated platelet

membrane18. It is important to note here that the hemostatic promise shown by intravenous

113 FVIIa has also led to its incorporation in externally applicable dressings like gelatin sponge

or microporous polysaccharide spheres, to improve the overall hemostatic capacity of these

technologies34,35. The clinical benefit of recombinant FVIIa in trauma, intracranial

bleeding, and other heavy bleeding surgeries are also being explored36–39. For example,

two randomized, placebo-controlled, double-blind trials (one in blunt trauma and one in

) have been carried out to evaluate the efficacy and safety of

recombinant FVIIa as adjunctive therapy for hemostatic management of trauma patients,

and the results have indicated a reduction in need of blood component transfusion in the

patients, along with systemic safety of the FVIIa in the dose ranges studied38.

3.3 Synthetically derived pro-coagulant systems

In recent years, research approaches have also focused on mimicking and leveraging the

body’s coagulation mechanisms using intravascularly applicable synthetic biomaterials.

For example, in the work of Morrissey and colleagues, it was demonstrated that inorganic

polyphosphate polymers (PolyP) secreted from activated platelets is capable of facilitating

activation of FXI to FXIa and FV to FVa, which subsequently results in the prothrombinase

complex leading to thrombin generation and back-activation of other coagulation factors40–

42. This property of platelet-derived PolyP was recently leveraged by Kudela et al to create

silica nanoparticles coated with this PolyP polymer (PolyP-SNP), which demonstrated promising hemostatic characteristics in vitro in the context of augmenting thrombin generation and reducing clotting time in thromboelastographic evaluation with pooled plasma43. To advance this technology towards safe clinical translation as an intravenous

hemostat, an important design requirement would be to control the exposure of the PolyP-

114 coated surface to coagulation factors selectively at the bleeding site such that the thrombin

and fibrin augmentation can be localized at that site while avoiding systemic off-target pro- coagulant effects. An additional challenge may be the maintenance of PolyP activity over reasonable periods of time, since these polyphosphate polymers are known to be rapidly deactivated by phosphatases in circulation. Masking the PolyP coating on particles from phosphatase action in circulation and then programming in a mechanism to have it exposed

(unmasked) selectively at the bleeding site can lead to a unique target-responsive hemostatic particle design. The PolyP material itself has shown additional interesting properties regarding enhancement of fibrin clot strength and stability, in that it can inhibit fibrinolysis by affecting thrombin-activated fibrinolysis inhibitor (TAFI), as well as, increase fibrin fiber thickness40. This property may be beneficial to intravenous hemostasis

applications if the clot formation can be directed selectively at the bleeding site, especially

in non-compressible hemorrhage (e.g. in trauma). In another interesting recent report,

Baylis et al have reported on ‘self-propelled’ particles made from calcium carbonate

+ (CaCO3) mixed with protonated tranexamic acid (TXA-NH3 ), where the propulsion

+ occurs due to carbon dioxide generation via the reaction between CaCO3 and TXA-NH3

in the milieu44. These particles were reported to be capable of being propelled through

blood against convective forces, and external application of these particles in several pre- clinical bleeding models (mice and pigs) demonstrated promising tissue-penetration to deliver loaded thrombin for hemostatic clot promotion. The particles were reported to have safe biodistribution when intravenously injected in mice. However, it was unclear from the report whether the particles, reportedly about 10μ in diameter, were designed for actual safe intravascular applications systemically, since solid particles of such large diameter are

115 known to pose occlusive risks in the microvasculature. Figure 3-1 shows selected results

from hemostasis-relevant studies carried out with PolyP-coated silica particle systems

+ (Figure 3-1A and B) and the ‘self-propelled’ CaCO3 particles mixed with TXA-NH3 and

thrombin payloads (Figure 3-1C, D and E). Figure 3-1A demonstrates that PolyP loaded in silica nanoparticles (PolyP-SNP) accelerates thrombin generation compared to SNP alone while Figure 3-1B demonstrates that PolyP delivered via SNP particles (PO3 in SNP)

reduces clotting time of blood (i.e. speeds up coagulation) compared to PolyP in solution.

Figure 3-1C shows representative fluorescence and scanning electron micrograph images

of CaCO3 particles along with a schematic of administration of such ‘self-propelled’

+ particles loaded with thrombin in TXA-NH3 medium to a liver puncture site. Figure 3-1D

shows fluorescence-based confirmation (green fluorescence) of particles in the liver injury

model in mice, while Figure 3-1E shows significant reduction of blood loss in the mouse

liver injury model when such ‘self-propelled’ particles were used to deliver thrombin deep

into the injury site.

3.4 Materials and technologies to enhance clot strength and stability

As described previously in Chapter 1, in physiological scenarios hemostasis is a highly

spatiotemporally regulated balance between clot formation and clot lysis45. However, in

pathological scenarios, dysregulated fibrinolysis can cause substantial hemostatic

abnormality (e.g. trauma-induced coagulopathy) leading to sub-optimal stoppage of

bleeding (weak fibrin clot and hyper-fibrinolysis) as well as disseminated intravascular

coagulopathy (DIC) risks46. In such cases, drugs, biomaterials and technologies that can

reduce fibrinolysis or increase fibrin strength (and stability) can provide significant

116 therapeutic benefit. One such drug is the antifibrinolytic agent tranexamic acid (TXA), which is a synthetic derivative of the amino acid lysine that inhibits fibrinolysis by blocking the lysine binding sites on plasminogen47,48. TXA has recently undergone extensive clinical studies as an intravenous hemostat, e.g. in the Clinical Randomization of an

Antifibrinolytic in Significant Hemorrhage 2 (CRASH-2) trial49,50. These studies have

Figure 3-1. Selected results from hemostasis-relevant studies carried out with PolyP-coated silica particle + systems (A and B) and thrombin-loaded CaCO3-based particles mixed in TXA-NH3 that can self-propel themselves into wound depth (C, D and E). A demonstrates that PolyP loaded in silica nanoparticles (PolyP- SNP) accelerates thrombin generation compared to SNP alone while B demonstrates that PolyP delivered via SNP particles (PO3 in SNP) reduces clotting time of blood (i.e. speeds up coagulation) compared to PolyP in solution; C shows representative fluorescence and scanning electron micrograph images of CaCO3 particles along with a schematic of experimental set-up to administer thrombin-loaded ‘self-propelled’ particles in + TXA-NH3 medium to a liver puncture site; D shows presence of green fluorescent particles deep within the liver injury site in mice; E demonstrates that liver injury site-localized delivery of thrombin-loaded ‘self- propelled’ CaCO3 particles in TXA-NH3+ results in significant reduction of blood loss, compared to non- propelled thrombin delivery or control treatment. Figure components adapted and reproduced with permission43,44. Copyright 2015, John Wiley & Sons Inc. and 2015, AAAS Science Advances CC-BY NC.

117 established that TXA administration can reduce exsanguination and thus the number of

transfusions, and also reduce mortality risks in trauma and surgery patients. However,

CRASH-2 trial studies were performed in civilian hospitals and did not have specificity

towards measures of coagulopathy, injury severity or mechanism of injury, which may not

be fully applicable to combat trauma. Therefore, TXA has been further studied for

hemostatic potential in severe combat trauma, e.g. in the Military Application of

Tranexamic Acid in Trauma Emergency Resuscitation (MATTERs) study51. These studies

have further concluded that the use of TXA with blood component–based resuscitation in

treating combat injury results in improved coagulation outcomes and survival in patients

requiring massive transfusion. Based on such promising findings, further studies with TXA

for mitigating hemorrhagic complications are currently ongoing, e.g. the Pre-hospital Anti- fibrinolytics for Traumatic Coagulopathy and Haemorrhage (PATCH) trial, the Study of

Tranexamic Acid during Air Medical Prehospital Transport (STAAMP) trail and the

efforts of the Trans-agency Collaboration for Trauma Induced Coagulopathy (TACTIC)

consortium52, that are expected to enhance the optimal therapeutic approaches involving

TXA.

It has also been well established that when platelets interact with fibrin polymers within blood clots, the biomechanical contraction of clot-associated active platelet cytoskeleton facilitates the retraction and stiffening of clots53,54. This property of platelets was used as

an inspiration to produce synthetic platelet-like particles (PLPs) that were formed of

deformable ultra-low crosslinked poly(N-isopropylacrylamide-co-acrylic acid) (pNIPAm-

AAc) gel microparticles surface-modified with fibrin protofibril-specific antibody

118 motifs55. These PLPs were tested in vitro by flowing them in thrombin-activated platelet- poor plasma (PPP) within endothelialized microfluidic devices, and the particles demonstrated formation of fibrin. It should be noted that in this study the PPP was already thrombin-activated, and classically it is the thrombin (and not platelets) that break down the plasma fibrinogen to fibrin. Platelets essentially augment this process by providing the active platelet surface for the formation of prothrombinase complex and PLPs were not reported to have such active ‘fibrin-generating’ surface but rather a ‘fibrin-binding’ surface. However, it was not clear from the studies how the thrombin-activated PPP itself

could not form sufficient fibrin, but the presence of the PLP particles in that thrombin-

activated PPP somehow resulted in fibrin formation at levels similar to that of platelet-rich

plasma. Nonetheless, the PLPs demonstrated fibrin-retracting property when incorporated

in clots, although this action occurred over much longer time-scales compared to retraction

time-scales shown by natural platelets. For in vivo studies, the PLPs were injected in rats

and 5 min after injection the rat femoral vein was injured to bleed. The PLPs were found

to get incorporated in the clots at the vein injury site and the bleeding time for PLP-injected

animals were found to be only slightly lower (~120 sec) than non-injected animals (~150 sec) and comparable to FVIIa-injected animals. Further evaluation of this technology will be necessary regarding evaluation of long-term biodistribution (higher than 5 min circulation), systemic side-effects and performance in animal models with platelet dysfunctions (e.g. thrombocytopenia, coagulopathy etc.).

Other reported literature on polymeric material for intravenous hemostat application was inspired by the physiological action of the transglutaminase coagulation factor FXIIIa,

119 which strengthens and stabilizes the fibrin clot by creating intra- and inter-fiber cross-links

Figure 3-2 Schematic of fibrinogen conversion to fibrin and assembly of cross-linked fibrin biopolymeric mesh catalyzed by thrombin (FIIa) and FXIIIa, and hemostatic materials and technologies inspired by these mechanisms.

through amide bond formation between the lysine and glutamic acid residues of self-

assembled fibrin protofibrils (native mechanism schematic shown in Figure 3-2)56,57. This

polymer, termed polySTAT, was formed by incorporating fibrin-binding peptide domains

on a linear water-soluble (hydroxyethyl)methacrylate (HEMA) and N-hydroxysuccinimide

methacrylate (NHSMA) co-polymer backbone [p(HEMA-co-NHSMA)]58. The fibrin-

binding peptide itself was adapted from the work previously reported by Kolodziej et al

regarding phage display-based development of fibrin-specific peptides59. The polySTAT

polymer was able to incorporate homogenously within fibrin clots when added to

fibrinogen in presence of thrombin. Rheometric and thromboelastographic

characterization of mechanical properties of polySTAT-enriched fibrin clot compared

against fibrin clots enriched with recombinant FXIIIa or with polySCRAM (control

polymer containing scrambled non-specific peptide) showed that the polySTAT polymer

enhanced clot formation, strengthening and stabilization capacities at levels significantly

higher than the control polymer and comparable to the physiologically relevant

120 recombinant FXIIIa activity. In a rat femoral artery bleeding model of trauma-induced coagulopathy, intravenous administration of the polySTAT polymer significantly enhanced clotting strength and stability, reduced re-bleeding and blood loss and rendered efficient stabilization of blood pressure with reduced need of volume resuscitation.

Biodistribution of the polymer indicated that the material is mostly cleared from plasma and sequestered in the kidney over time, which may be indicative of its circulation safety regarding off-target effects. Dose response effects of this polymer as well as its long-term effects from accumulation in the kidney need to be further investigated. Figure 3-3 shows representative results from hemostatic studies carried out with PLP particles (Figure 3-3A-

D) and with polySTAT polymer systems (Figure 3-3E-G). Figure 3-3A shows magnified images of fibrin-binding soft microgel platelet-like particles (PLPs), natural platelets

(PRP), non-binding control microgel particles (S11–ULCs), and fibrin-binding rigid polystyrene particles (H6-PS) within fibrin matrices 1 hr post fibrin-polymerization while

Figure 3-3B shows the calculated spread area of PLPs, natural platelets, S11-ULCs and

H6-PS particles within these fibrin matrices, demonstrating that soft PLPs have platelet- like spreading behavior (compared to control non-binding and control rigid particles).

Figure 3-3C and D show bleeding time and blood loss analysis data in rat femoral vein injury model where administration of PLPs resulted in reduced bleeding time as well as reduced blood loss similar to administration of FVIIa (compared to saline vehicle or control particles). Figure 3-3E shows the homogenous incorporation of the fibrin-strengthening polySTAT polymer (green fluorescence) within native fibrin matrix (red fluorescence).

Figure 3-3F shows that polySTAT incorporation and strengthening within fibrin results in reduced clot lysis over time (i.e. increased clot strength and stability) compared to control

121 saline and control polymer (polySCRAM). Figure 3-3G demonstrates that intravenous administration of polySTAT results in significant reduction of blood loss rate in a rat femoral artery hemorrhage model.

Figure 3-3 Selected results from hemostatic studies carried out with platelet-like microgel particles (PLPs, A-D) and with polySTAT injectable polymer systems (E-G). A shows magnified images of fibrin-binding soft ultra-low crosslinked microgel particles (PLPs), natural platelets (PRP), non-binding control microgel particles (S11–ULCs), and fibrin-binding rigid polystyrene particles (H6-PS) within fibrin matrices 1 hr post fibrin polymerization and B shows the calculated spread area of PLPs, PRP platelets, S11-ULCs and H6-PS particles within the fibrin matrices, demonstrating that soft PLPs have spreading behavior comparable to natural platelets (PRP) and greater than control particles; C and D show bleeding time and blood loss analysis data in rat femoral vein injury model where administration of PLPs resulted in reduced bleeding time and blood loss comparable to administration of FVIIa and significantly lower than treatment with vehicle or control particles; E shows the homogenous incorporation of the fibrin-binding polySTAT polymer (green fluorescence) within native fibrin matrix (red fluorescence); F shows that polySTAT incorporation within fibrin results in crosslinking based strengthening of fibrin and thus reduction of clot lysis over time (i.e. increased clot stability) compared to treatment with control saline and control polymer (polySCRAM); G demonstrates that intravenous administration of polySTAT results in significant reduction of blood loss rate in a rat femoral artery hemorrhage model. Figure components adapted and reproduced with permission.55,58 Copyright 2014, Macmillan Publishers Ltd and 2015, AAAS.

3.5 Materials and technologies for vascular embolization

Another interesting area of using synthetic materials in intravascular hemostatic applications, is the area of blood vessel embolization that is often used to staunch internal bleeding in aneurysms, gastro-intestinal bleeds etc.60–62. Catheter-mediated deployment of

122 permanent metallic coils or synthetic liquid polymers (e.g. Onyx, an ethylene-vinyl alcohol copolymer dispersed in dimethyl sulfoxide solvent) has already been clinically approved for such site-localized embolic applications60,63,64. Metallic coils and Onyx polymers have

been reported to present issues of material migration, secondary tissue damage, off-target

embolization, incomplete occlusion, biocompatibility issues etc.61,65. To resolve some of

these issues, Avery et al recently reported on development of a shear-thinning polymer system formed from a nanocomposite of gelatin and silica nanoparticles, that can be injected into vascular compartment for embolic effect66,67. The unique rheological

properties of this material, rendered by charge based ‘physical hydrogel’ interactions

between the silica and gelatin materials, allows for shear-driven (e.g. through a pressurized catheter) introduction into blood vessel due to shear-thinning of the material and upon release of pressure the nanocomposite can quickly recover its solid like behavior to render embolization. This unique polymeric system has shown promising embolic efficacy in small and large animal models, to staunch bleeding. Gelatin is a collagen-derived degradation product and therefore can directly activate platelets68. Silica, because of its

negative charge, is known to activate coagulation factors69. Therefore, while these materials attributes may be beneficial in augmentation of embolic and hemostatic effects selectively at the injury site, presence of these materials in systemic circulation as a result of polymer nanocomposite degradation, may pose systemic pro-thrombotic and pro- coagulant risks. Therefore, future advancement of this material and embolic technology would require in depth evaluation of its long-term systemic safety (including degradation products), clearance mechanisms, off-target effects on systemic hemostatic components

(e.g. platelets and coagulation factors) and effects on immune and complement system.

123 3.6 Natural platelet-derived systems

The platelet membrane presents multiple macromolecular entities that participate in

various aspects of primary (platelet adhesion and aggregation) and secondary (fibrin mesh

formation) hemostasis. Figure 3-4 shows a schematic of platelet’s hemostatically relevant surface-interactive adhesion and aggregation mechanisms and the technologies inspired therefrom. The GPIb-IX-V complex as well as GPIa/IIa and GPVI are responsible for

platelet adhesion at the bleeding site by binding to vWF and collagen, respectively70–73.

The platelet surface integrin GPIIb-IIIa, in its activated conformation, is responsible for

binding blood protein fibrinogen by which platelets can bridge with each other to form a

platelet plug74,75. Active platelet surface also expresses P-selectin (by secretion from

platelet α granules), which can bind to glycoprotein ligands on other platelets and

leukocytes at sites of platelet aggregation and inflammation76,77. The surface of active

platelets also present a high extent of anionic phospholipids (e.g. phosphatidylserines

translocated from inner leaflet to outer leaflet of the membrane) that, in presence of calcium

ion (Ca++), facilitates co-localization and activation of several coagulation factors to form

the extrinsic and intrinsic tenase as well as the prothrombinase complex, ultimately leading

to the conversion of Factor II (prothrombin) to Factor IIa (thrombin) which in turn breaks

down fibrinogen to fibrin. Activated platelets also secrete several granule contents, e.g.

PolyP and vWF that can modulate and amplify various aspects of coagulation. Because of

such multifactor role of platelets in clot promotion, transfusion of platelet concentrates

(PC) has become an important clinical component in the prophylactic management of

bleeding risks, as well as, emergency treatment of traumatic hemorrhage78,79. However, the

PC products at room temperature pose high risks of bacterial contamination, and also are

124

Figure 3-4 Schematic of platelet’s injury site-selective adhesion mechanisms (platelet GPIbα of the GPIb- IX-V complex binding to vWF, and GPIa-IIa as well as GPVI binding to collagen) and aggregation mechanism (fibrinogen-mediated bridging of active platelet surface integrin GPIIb-IIIa), and various hemostatic technologies inspired by these mechanisms.

known to undergo activation and degranulation during transport and storage, resulting in

very short shelf-life (3-5 days)80,81. Therefore, a significant volume of research is currently

being directed towards reducing contamination risks by pathogen reduction technologies82–

84, as well as, enhancing the platelet storage stability and shelf-life via cold-storage (4oC),

freezing and lyophilization85–92. In parallel, several research efforts have been directed

towards development of hemostatic technologies that use a variety of platelet-derived

biological components.

Since platelet’s hemostatic properties are significantly mediated by various membrane

components, platelet membranes isolated from outdated platelet concentrates have been

utilized to develop phospholipid vesicles, about 0.5 μ in diameter, through a serial process

of differential centrifugation, freeze-thawing, pasteurization and sonication, and these vesicles have been evaluated as intravenously administered hemostat under the product

125 name ‘infusible platelet membrane’ (IPM CyplexTM, Cypress Bioscience, San Diego,

California)93. Evaluation of IPM has been carried out in vitro, in vivo in pre-clinical models

and in multiple phases of clinical studies, but it is yet to be approved by the FDA for clinical

use, due to certain issues. For example, since the membrane is isolated from allogeneic

platelet concentrates, the pathogenic contamination risks associated with platelet

concentrates may also be present in IPM products, which in turn require rigorous

purification and treatment to ensure safety. It has been reported that such treatments can

lead to loss of functional components in the membrane that are otherwise needed for

hemostatic mechanisms. Several antigens (HLA I and II) and integrin GPIIb-IIIa have been

reported to be lost in IPM due to purification and pathogen reduction treatments94. While

loss of HLA can be viewed as an advantage in the context of the ability to use IPM without

the need for blood type matching, the loss of GPIIb-IIIa renders inability to form the fibrinogen-mediated platelet plug and thereby affects primary hemostatic capability of

IPM. Interestingly, presence of pro-coagulant phospholipid function was reported to be retained on IPM surface, implying that secondary hemostatic mechanisms may still be rendered by these vesicles95. Also, partial functionality of the GPIb-IX-V complex and P- selectin was reported to be retained in flow cytometric analysis of IPM, suggesting that primary hemostatic function may be partly possible by this product94. In pre-clinical animal

models, IPM has shown promising hemostatic efficacy (usually measured by reduction in

blood volume loss or improvement in time for blood to clot from a controlled injury),

without systemic toxicity and thromboembolic risks96–98. These promising studies have led

to clinical evaluation of IPM in healthy as well as thrombocytopenic patients. In early phase

clinical studies, transfusion of IPM showed hemostatic capability comparable to

126 transfusion of allogeneic platelet concentrates and did not show major adverse effects of

immune response and refractoriness. Clinical efficacy of IPM products, especially in

patients who show refractory immune response to repeated platelet transfusions as well as

in patients who are thrombocytopenic with multiple other indications, is yet to be

established. Efficacy in non-compressible hemorrhage (e.g. in trauma) is also yet to be

established. A variant of the IPM strategy named Thrombosomes® (CellPhire, Maryland

USA) was recently reported, where freeze-dried platelets stabilized in trehalose were prepared from a pool of Group O in-date leukoreduced apheresis platelet units99,100. This

product was tested for its physical and surface characteristics, systemic distribution,

toxicity, sequestration and hemostatic efficacy in vitro as well as in vivo in pre-clinical animal models of bleeding (in rabbits, beagles and rhesus macaque), and the results reported were quite promising in terms of in vitro platelet aggregatory and pro-coagulant capacity and in vivo systemic circulation and safety. In vivo hemostatic efficacy has been reported for this product in terms of reducing bleeding in pre-clinical animal models.

Interestingly, isolated platelet membrane-based technologies have also been recently reported by several research groups in the context of coating drug delivery vehicles (e.g. polymeric and silica nano/microparticles) to demonstrate targeted action in treating cancer and bacterial infection, thereby emphasizing the potential utilization of platelet’s role in multiple for unique treatment strategies101–104. Figure 3-5 shows selected

results from hemostatic studies with IPM, where addition of IPM in thrombocytopenic

blood perfused ex vivo in rabbit aorta segments dose-dependently increased fibrin

deposition at low and high shear rates (shown in Figure 3-5A) and in vivo

127

Figure 3-5 Selected results from hemostatic studies with infusible platelet membrane (IPM) technology, where (A) addition of IPM in thrombocytopenic blood perfused ex vivo in rabbit aorta segments dose- dependently increased fibrin deposition at low and high shear rates and (B) in vivo administration of IPM in thrombocytopenic rabbits resulted in significant reduction of bleeding time. Figure components adapted and reproduced with permission.97,98 Copyright 2000, John Wiley & Sons Inc. and 2012, IJPSR.

administration of IPM in thrombocytopenic rabbits resulted in reduction of bleeding time

(shown in Figure 3-5B).

Instead of using intact isolated platelet membrane to form freeze-dried vesicles or coated

particle systems, some approaches have focused on incorporating specific platelet surface

protein and receptor components in lipidic or polymeric particle systems. The earliest

report of this approach was in a product called ‘plateletsome’, where deoxycholate-induced extraction of platelet membrane fraction with multiple protein components including GPIb,

GPIIb/IIIa and GPIV was incorporated into the membrane of liposomal vesicles formed of sphingomyelin, phosphatidylcholine and monosialylganglioside using reverse-phase

evaporation sonication technique105. This product was tested for hemostatic capability and

demonstrated substantial reduction of tail-bleeding in thrombocytopenic rats while

showing systemic safety. However, no further reports have been available about this

product after the mid 1990s. In an analogous approach, Takeoka and colleagues from Keio

128 University and Waseda University in Japan developed recombinant GPIbα (rGPIbα) and

GPIa-IIa (rGPIa-IIa) proteins and conjugated them on the surface of liposomes, latex beads

or albumin particles to render synthetic particulate systems with platelet-inspired vWF- binding and collagen-binding capabilities106–109. In vitro, these designs have shown

efficient binding to vWF-coated surfaces as well as collagen surfaces in presence of soluble

vWF, with more binding evident at higher shear rates, thereby closely mimicking the

mechanisms of natural platelet adhesion at high shear. Takeoka and colleagues have also

investigated leveraging the co-operative activity of platelet surface motifs that facilitate

hemostasis in a physiological scenario. Platelets render primary hemostasis by a co-

operative interplay between pro-adhesive (vWF-binding and collagen-binding) and pro-

aggregatory (fibrinogen-mediated GPIIb-IIIa binding) functionalities110,111. Takeoka et al

demonstrated that platelet’s pro-adhesive mechanisms of vWF-binding and collagen- binding could be combined on a synthetic particle surface by conjugating rGPIbα and rGPIa-IIa motifs together on the outer leaflet of liposomes108,109. The rGPIbα motif was

utilized to enable reversible binding to vWF under shear flow, while the rGPIa-IIa motif

was utilized to enable adhesion stabilization by concomitant binding to collagen.

Compared to liposomes bearing rGPIbα only or rGPIa-IIa only, those bearing a

combination of these motifs showed significantly higher adhesion and retention on

collagen-coated surfaces in presence of soluble vWF under high shear flow conditions. In another recent interesting approach, Doshi et al created spherical polystyrene as well as crosslinked discoid albumin particles, surface-decorated either with recombinant vWF-A1

domain fragment or recombinant platelet GPIbα domain fragment to mimic platelet-to-

vWF interaction under flow112. These particle systems were evaluated in vitro where the

129 discoid particles showed better adhesion than spherical particles on target surfaces, and the

particles could get incorporated within platelet thrombi on collagen-coated surfaces in

microfluidic chambers. Figure 3-6 shows selected study results of the above-described

technologies that leverage platelet’s adhesion mechanisms. Aggregometry studies shown

in Figure 3-6A demonstrates that addition of rGPIbα-decorated albumin microsphere particles significantly enhances that percent (%) aggregation of platelets in presence of vWF and ristocetin, compared to without ristocetin, thereby confirming that the aggregation enhancement is due to the binding of rGPIbα to ristocetin-induced unfolded vWF. Figure 3-6B shows that polystyrene microparticles surface-decorated with vWF’s

A1 domain or GPIbα fragment can undergo shear-responsive enhanced adhesion on GPIbα

–coated or vWF-coated microfluidic surfaces respectively. Figure 3-6C shows a schematic of rGPIbα-coated red fluorescent liposomes flowed over vWF-coated surfaces, with the results demonstrating enhanced adhesion and accumulation of these liposomes over time at higher shear, mimicking the shear-responsive interaction of platelet GPIbα to vWF in platelet’s adhesion mechanism.

The designs described above represent primarily the adhesive component of platelet function in hemostasis, but do not have the ability to aggregate among themselves or site- selectively amplify the aggregation of active platelets. Platelet aggregation is a function of fibrinogen-mediated bridging of integrin GPIIb-IIIa on active platelets and therefore certain designs have focused on leveraging this mechanism by coating particles with fibrinogen. In the 1980s, Coller et al reported conjugation of fibrinogen on acrylonitrile beads, where the resultant beads enhanced agglomeration of ADP-activated platelets in

130

Figure 3-6 Selected results from studies involving technologies that leverage platelet’s bleeding site- selective adhesion mechanisms. A shows platelet aggregometry studies to demonstrate that addition of rGPIbα-decorated albumin microsphere particles significantly enhances the aggregation level of platelets in presence of vWF and ristocetin, compared to without ristocetin, thereby confirming that the aggregation enhancement is due to the binding of rGPIbα to ristocetin-induced unfolded vWF; B shows that polystyrene microparticles surface-decorated with vWF’s A1 domain or platelet’s GPIbα fragment can undergo shear- responsive enhanced adhesion on GPIbα–coated or vWF-coated surfaces respectively, under flow, mimicking platelet-relevant adhesion mechanisms; C shows an experimental schematic of flowing rGPIbα- decorated red fluorescent liposomes over vWF-coated surfaces, with the corresponding fluorescence microscopy results demonstrating enhanced adhesion and accumulation of these liposomes over time at higher shear, mimicking the shear-responsive interaction of platelet GPIbα to vWF. Figure components adapted and reproduced with permission109,112. Copyright 2012, John Wiley & Sons Inc. and 2000, American Chemical Society.

vitro113. Fibrinogen-coated systems were also reported in the early 1990s by Agam and

Livne, where erthyrocytes (RBCs) were coated with fibrinogen, and these modified RBCs showed induction of platelet aggregation in vitro and reduction of prolonged bleeding time in thrombocytopenic rats in vivo114. In the mid 1990s, Yen et al reported a biomaterials technology named ‘ThrombospheresTM’ consisting of albumin microspheres coated with fibrinogen, and these particles rendered aggregation of active platelets in vitro as well as

131 reduction of bleeding time in an ear-punch model in rabbits in vivo115,116. A similar design

termed ‘SynthocytesTM’ was reported by Levi et al in the late 1990s, where fibrinogen-

coated albumin microcapsules were able to reduce the bleeding time as well as blood

volume loss in an ear-punch bleeding model and abdominal surgical incision model in

thrombocytopenic rabbits117. In vitro platelet-aggregating capacity of the fibrinogen-coated

albumin particles under shear flow conditions was also reported in early 2000 by the

research group from Waseda University and Keio University in Japan118. Another similar

design of fibrinogen-coated albumin particles was reported under the product name

FibrinoplateTM by a company called Advanced Therapeutics & Co, but further information

about this product was not available after 2011. Figure 3-7 shows selected results from studies carried out with various fibrinogen-coated particle designs. Figure 3-7A shows the effect of administering fibrinogen-coated albumin microcapsules (SynthocyteTM) in

reducing ear punch bleeding time in thrombocytopenic rabbits, where induction of

thrombocytopenia (TCP) significantly increased bleeding time from normal and

administration of SynthocyteTM particles (SC) could significantly reduce bleeding time in

these animals, compared to administration of control particles (CP) or saline. Figure 3-7B

shows scanning electron micrograph of the SynthocyteTM particles incorporated with

platelets and fibrin in hemostatic clots. Figure 3-7C shows the ability of fibrinogen-coated

albumin particles to bind to platelet-immobilized surfaces in absence versus presence of

Ca++ ions, compared to binding of unmodified albumin particles, while Figure 3-7D shows representative fluorescence microscopy images of these binding studies. These results

establish that fibrinogen-coated particles can interact and bind with active (e.g. Ca++

activated) platelets under flow environment, possibly via interaction with platelet surface

132

Figure 3-7 Selected results from studies carried out with various fibrinogen-coated particle designs. A shows the effect of administering fibrinogen-coated albumin microcapsules (SynthocyteTM) in reducing ear punch bleeding time in thrombocytopenic rabbits, where induction of thrombocytopenia (TCP) significantly increased bleeding time from normal levels and administration of SynthocyteTM particles (SC) could significantly reduce bleeding time in these animals, compared to administration of control particles (CP) or saline; B shows scanning electron micrograph of the SynthocyteTM particles incorporated with platelets and fibrin in hemostatic clots; C shows the ability of fibrinogen-coated albumin particles to bind to platelet- immobilized surfaces in absence versus presence of Ca++ ions, compared to binding of unmodified albumin particles and D shows representative fluorescence microscopy images of these binding studies, confirming high binding of fibrinogen-coated particles and establishing that these particles can interact with active platelets under flow environment, possibly via interaction with platelet surface integrin GPIIb-IIIa, thereby mimicking and amplifying the active platelet aggregation component of hemostasis. Figure components adapted and reproduced with permission117,118. Copyright 1999, Macmillan Publishers Ltd and 2001, American Chemical Society.

integrin GPIIb-IIIa. Thus, these technologies seem to efficiently leverage and amplify the active platelet aggregation component of hemostasis by means of a ‘super-fibrinogen’ design. Off-target systemic thromboembolic risks can be a possible issue with such designs, and therefore rigorous in vivo pre-clinical studies are needed to establish safety and clearance.

133 3.7 Synthetic biomaterials-based platelet mimics and substitutes

Creating platelet-inspired designs by utilizing isolated platelet membranes, fibrinogen

coatings and platelet surface-relevant recombinant protein fragments can present issues of batch-to-batch variation in bioactivity, protein stability and solubility, potential immunogenicity and high cost. Therefore, a significant research interest has been focused on utilization of synthetic ligand systems (e.g. peptides) decorated on semi-synthetic or synthetic particle platforms that can render hemostatic mechanisms of platelets while providing advantages of controllable and scalable manufacture, consistent bioactivity, long shelf-life and reduced biological/immunological side effects. To this end, the majority of research approaches has been focused in the area of mimicking fibrinogen-binding to active platelet surface integrin GPIIb-IIIa by decorating particle surfaces with fibrinogen-relevant synthetic peptide ligands, in essence once again rendering a ‘super-fibrinogen’ type design.

Fibrinogen is a hepatic protein, approximately 340,000 in molecular weight and present in the blood at concentration ranges of 200-400 mg/dL. Fibrinogen has an approximately 45 nm long macromolecular structure consisting of two outer D-domains, each connected by a coiled-coil segment to its central E domain. The molecule contains two sets of three polypeptide chains, termed Aα, Bβ and γ, which are joined together in the N-terminal E domain by five symmetrical disulfide bridges75,119. Upon activation of platelets in a

hemostatic (or thrombotic) environment, the surface integrin GPIIb/IIIa assumes a ligand-

binding conformation that can bind to fibrinogen via the Aα chain tetrapeptide Arg-Gly-

Asp-Ser (RGDS) and the γ chain 12-residue peptide His-Leu-Gly-Gly-Ala-Lys-Gln-Ala-

Gly-Asp-Val (HLGGAKQAGDV)120,121. Since these peptide sequences are present on both

termini of fibrinogen, the fibrinogen molecule can act as a bridging system between active

134 platelets (schematic shown previously in Figure 3-4), resulting in aggregation of these

platelets to form the primary hemostatic plug. Based on these mechanisms, one of the

earliest semi-synthetic designs was that of ‘Thromboerythrocytes’ reported by Coller et al,

where RGD sequence Ac-CGGRGDF-NH2 was covalently conjugated to the surface amino

group of RBCs using a heterobifunctional crosslinking reagent (N-maleimido-6-

aminocaproyl ester of 1-hydroxy-2-nitrobenzene4-sulfonic acid)122. These RGD-decorated

RBCs showed binding to active platelet integrin GPIIb-IIIa, were capable of co-

aggregating with activated platelets and could bind to active platelets on collagen-coated

surfaces. Decoration of synthetic particle platforms (latex beads, poly-lactic acid based

polymeric particles etc.) with RGD sequences (e.g. CGGRGDF on latex beads and GRGDS

on poly-lactic acid-based particles) have been reported by Takeoka and colleagues123 as

well as Lavik and colleagues124,125, in the context of mimicking and amplifying the

fibrinogen-mediated interactions with integrin GPIIb-IIIa on activated platelets.

Interestingly, in the studies conducted by Takeoka et al with CGGRGDF-decorated particle systems, it was found that these particles could promote activation (and subsequent aggregation) of inactive platelets123. Similar effects of linear RGDS-containing peptides

have also been reported by Du et al and Beer et al120,126. Furthermore, the RGD tri-peptide sequence is known to have ubiquitous cell- and matrix-recognition capabilities127,128. In

fact, this sequence has been reported in a variety of research including targeting to tumor

cells, endothelial cells, bone cells etc. beyond its interaction with platelets. Therefore,

utilizing CGGRGDF or GRGDS type of linear ubiquitous RGD peptides to mimic

hemostasis-relevant fibrinogen-mimetic platelet aggregation mechanisms on synthetic

platelet designs may pose major systemic thromboembolic risks (due to activation of

135 circulating resting platelets) as well as render cross-reactivity with other cellular entities in

vivo that may affect clinical translation. Because of the issues associated with these linear

ubiquitous RGD peptides, Takeoka and colleagues have also researched utilizing

fibrinogen γ-chain relevant peptides, specifically the sequence HLGGAKQAGDV

otherwise known as the H12 peptide, to decorate particle platforms. This peptide was used

to modify the surface of latex beads, albumin particles and liposomes, and the resultant

systems were evaluated for their platelet-inspired pro-aggregatory hemostatic activities129–

133. In vitro, the H12-decorated latex beads showed minimal aggregation effect on inactive

platelets, suggesting that this peptide may confer improved systemic safety compared to

RGDS peptides. Albumin microparticles surface-decorated with H12 peptides were able to

enhance platelet aggregation on collagen-immobilized surfaces in vitro and improve ear

punch bleeding time in thrombocytopenic rabbits and tail bleeding time in

thrombocytopenic rats, both by approximately 50%. Liposomes surface-decorated with

H12 peptides showed similar hemostatic capability, and encapsulating ADP in such

liposomes enabled augmentation of platelet activation and aggregation as well as correction

of ear bleeding time in thrombocytopenic rabbits and tail bleeding time in

thrombocytopenic rats. These studies however did not provide any information regarding

the overall biodstribution of the particles. Figure 3-8 shows selected results from in vitro

and in vivo studies that have been carried out using particle designs surface-decorated with

fibrinogen-relevant RGD or H12 peptides to mimic and amplify fibrinogen-mediated

platelet aggregation. Figure 3-8A shows platelet aggregometry studies with RGD peptide- decorated RBCs (Thromboerythrocytes), where unmodified control erythrocytes were unable to enhance aggregation of ADP-activated platelets and Thromboerythrocytes were

136 unable to enhance aggregation of inactive platelets or integrin GPIIb-IIIa-blocked platelets, but Thromboerythrocytes were able to significantly enhance the aggregation of ADP- activated platelets. Figure 3-8B shows representative fluorescent images as well as quantitative surface-coverage data of H12 peptide-decorated, or RGD peptide-decorated or undecorated fluorescently-labeled latex beads suspended in reconstituted blood binding to platelet-immobilized surfaces, indicating the enhanced binding capability of H12- decorated and RGD-decorated beads to platelets. Figure 3-8C shows the capability of H12 peptide-decorated PEGylated albumin particles (H12-PEG-Alb) to significantly reduce ear punch bleeding time upon intravenous administration in thrombocytopenic rabbits, compared to administration of control undecorated PEG-albumin (PEG-Alb) particles.

Figure 3-8D shows the capability of H12 peptide-decorated PEGylated liposomal vesicles

(H12-PEG-vesicles) to significantly reduce tail bleeding time in thrombocytopenic rats, compared to administration of control undecorated PEG-vesicles. Figure 3-8E shows quantitative bleeding time reduction data and representative fluorescence image of hemostasized vessel, where administration of RGD-decorated polymeric nanoparticles

(GRGDS-PLGA-PLL particles) significantly reduced bleeding time in a rat femoral artery injury model, compared to administration of undecorated particles or saline.

Regarding use of peptides to mimic platelet’s adhesion mechanisms, an early demonstration is found in reports from Gyongyossy-Issa and colleagues involving the development of platelet GPIbα-relevant 15-mer peptides that have binding capability to vWF134. These peptides were tethered onto liposomal surfaces stabilized via cross-linking,

and the resultant liposomal constructs, 150-200 nm in diameter, were tested as ‘synthetic

137

Figure 3-8 Selected results from studies carried out with various particle platforms surface-decorated with fibrinogen-relevant RGD or H12 peptides to mimic platelet aggregation mechanism. A shows platelet aggregometry studies with RGD-decorated RBCs (Thromboerythrocytes), where unmodified control erythrocytes were unable to enhance aggregation of ADP-activated platelets and Thromboerythrocytes were unable to enhance aggregation of inactive platelets or integrin GPIIb-IIIa-blocked platelets, but they could significantly enhance the aggregation of ADP-activated platelets, thereby mimicking fibrinogen and GPIIb- IIIa mediated platelet aggregation mechanism; B shows representative fluorescent images as well as quantitative surface-coverage data of H12 peptide-decorated, or RGD peptide-decorated or undecorated fluorescently-labeled latex beads suspended in reconstituted blood and flowed over platelet-immobilized surfaces, indicating the enhanced GPIIb-IIIa-binding capability of H12-decorated and RGD-decorated beads to platelets; C shows the capability of H12-decorated PEGylated albumin particles (H12-PEG-Alb) to significantly reduce ear punch bleeding time upon intravenous administration in thrombocytopenic rabbits, compared to administration of control undecorated PEG-albumin (PEG-Alb) particles; D shows the capability of H12-decorated PEGylated liposomal vesicles (H12-PEG-vesicles) to significantly reduce tail bleeding time in thrombocytopenic rats, compared to administration of control undecorated PEG-vesicles; E shows quantitative bleeding time reduction data and representative fluorescence image of hemostasized vessel, where administration of RGD-decorated polymeric nanoparticles (GRGDS-PLGA-PLL particles)

138 significantly reduced bleeding time in a rat femoral artery injury model, compared to undecorated particles or saline. Figure components adapted and reproduced with permission122–124,129,130. Copyright 1992, American Society for Clinical Investigation; 2003, Elsevier; 2008, John Wiley & Sons Inc.; 2005, American Chemical Society and 2009, AAAS.

platelet substitutes’135. Although the biochemical characterizations of these constructs have been reported, the actual hemostatic efficacy evaluation of these constructs in vivo have not been reported in any pre-clinical animal model.

3.8 Discussion

Regarding mimicry of platelet functions, one straightforward approach that several research groups have employed involves detergent-based extraction of natural platelet’s membrane and then coating synthetic particles with this membrane material. Since many of platelets’ hemostatic mechanisms are associated with its surface biology, this approach, theoretically, can capture and leverage this surface biology on synthetic platforms for hemostatic (and drug delivery) applications. Additionally, due to a significant history of research on the evaluation of infusible platelet membrane (IPM) vesicles as intravenous hemostats, the membrane-coated particle approaches can potentially streamline their design optimization. One concern in this approach is the loss of bioactivity of certain platelet surface motifs that is rendered by the various membrane extraction and purification steps, which may in turn lead to capturing only a part of hemostatic mechanisms or presenting batch-to-batch inconsistency in hemostatic function. Another concern is that the limited availability of platelet concentrates may in turn affect the availability of extractable platelet membranes and this can become a logistical barrier to translation. Such issues can be potentially resolved by utilizing recombinant protein fragments or small peptide ligands

139 that can mimic the functions of platelets’ hemostasis-relevant surface components without having to mimic the complete surface membrane and complex biology. Examples of this approach are seen in designs that utilize recombinant GPIbα, GPIa-IIa and vWF-A1 protein fragments, as well as, in strategies that employ vWF-binding, collagen-binding and integrin

GPIIb-IIIa-binding peptides to decorate synthetic particle platforms.

A variety of particle fabrication technologies to modulate physico-mechanical and geometric parameters have been established in recent years, and this is expected to benefit the incorporation of such parameters in refining the materials and technologies for synthetic platelet research. Upon adhesion and activation at the injury site due to physical

(shear) and biochemical (receptor-based signaling cascades) triggers, platelets transform from the ‘resting’ biconvex discoid morphology to an ‘activated’ stellate pseudopodal morphology with significant reduction in modulus and remain irreversibly adhered at the injury site for hemostasis. This phenomenon can inspire the development of materials with environment-responsive physico-chemical and mechanical properties for development of platelet-inspired hemostats. Regarding utilization of semi-synthetic and synthetic particle based systems for designing platelet surrogates, it is critical to study and mitigate systemic risks associated with the particle material itself, as well as, with the surface chemistries and ligands utilized for particle surface decoration. In previous sections, the potential risks associated with using ubiquitous cross-reactive ligands (e.g. RGDS) have been discussed.

In the same note, every ligand motif should be studied both in free form as well as particle- tethered form for systemic effects on platelets (e.g. by aggregometry) and plasma (e.g. by thrombin generation assays). In addition, the base materials used for particle construction

140 should be chosen carefully to ensure minimal toxicity and high biocompatibility. It is also important to determine the spatiotemporal fate of these particles and materials once administered in vivo, in the context of biodegradation, clearance mechanisms and elimination pathways. Maintaining systemic safety while rendering targeted hemostatic action at the bleeding site within an appropriate time window and undergoing effective clearance, will be the key to success for any of these technologies.

Regarding utilization of coagulation cascade components for intravascular hemostasis, the direct use of recombinant coagulation factors is clinically well established in selective groups of patients. The bulk of these approaches are essentially directed at amplifying the generation of thrombin through tissue-factor pathway or common pathway of coagulation, such that the thrombin can convert fibrinogen to fibrin to facilitate clotting. It is important to note that the generation of thrombin is significantly enhanced in presence of active platelet membrane via formation of tenase complex (FVIIIa + FIXa) and prothrombinase complex (FXa + FVa) on the phosphatidyl serine-rich membrane surface. Therefore, use of coagulation factors in patients with platelet deficiency or dysfunction may have only limited benefit towards hemostasis. In comparison, use of materials and technologies that can promote activation of certain coagulation cascade components independent of platelet presence may show some benefit towards subsequently amplifying hemostasis. One such interesting material is the PolyP coating on particles that attempts to mimic the action of polyphosphates secreted from active platelets by activating FXI (to FXIa) and FIX (to

FIXa) en route to activation of FX and subsequently amplifying thrombin generation through the tenase and prothrombinase complexes on active platelet surface. While this

141 mechanism and technology have generated substantial interest in recent hemostasis research, it would be important to limit the action of PolyP-coated particles at the bleeding site, to avoid systemic hypergeneration of thrombin and fibrin. One potential approach to shielding of the PolyP coating can be the use of a mask that may be cleaved site-selectively by bleeding site relevant enzyme action, thereby exposing the PolyP only locally to amplify coagulation cascade activation and propagation. Precise regulation of these mechanisms will be important to avoid systemic effects. The same rationale holds true for technologies that may attempt to coat synthetic or semi-synthetic particle platforms with tissue factor or negatively charged phospholipids. Other pro-coagulant technologies that strengthen the fibrin clot (e.g. fibrin-crosslinking microgel particles or FXIIIa action mimicking polymers) may have to depend upon sufficient prior generation of fibrin at the bleeding site to be able to significantly benefit in vivo from their mechanism of fibrin-strengthening.

This in turn would require prior sufficient generation of thrombin and resultant conversion of fibrinogen in a site-selective manner at the bleeding site, which in turn requires the involvement of tissue factor-dependent thrombin generation and platelet-dependent thrombin amplification pathways at the site. Therefore, future research should be directed at modularly combining the hemostasis mimicking and enhancing mechanisms of the different materials and technologies to find an optimum system that (i) promotes and enhances active platelet or platelet-inspired aggregation at the bleeding site, (ii) augments thrombin generation and fibrin formation selectively at that site and (iii) strengthens the formed fibrin biopolymeric network to reduce clot lysis and amplify hemostasis. While individual components of such systems are currently under pre-clinical research, these and potential integrated systems need to be rigorously studied for biodistribution, systemic

142 safety and site-selective hemostatic action in multiple well-characterized bleeding models in vivo. Through robust interdisciplinary research between materials scientists, chemical and biomedical engineers, molecular biologists, biochemists and clinicians, future advancement of hemostatic materials and technologies can be envisioned to revolutionize bleeding treatment and transfusion medicine approaches.

3.9 Acknowledgments

This work was supported by the National Heart Lung and Blood Institute (NHLBI) of the

National Institutes of Health under award numbers R01 HL121212 (PI: Sen Gupta). The content of this publication is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

3.10 References

1. Kauvar DS, Lefering R, Wade CE. Impact of Hemorrhage on Trauma Outcome:

An Overview of Epidemiology, Clinical Presentations, and Therapeutic

Considerations. 2006 Jun;60(Supplement):S3–S11. Available from:

https://insights.ovid.com/crossref?an=00005373-200606001-00002

2. Kheirabadi BS, Klemcke HG. Symp. Paper RTO‐M P‐ HFM ‐ 109. 2004. p. 20.

3. Cotton BA, Gunter OL, Isbell J, Au BK, Robertson AM, Morris JA, St. Jacques P,

Young PP. Damage Control Hematology: The Impact of a Trauma Exsanguination

Protocol on Survival and Blood Product Utilization. 2008 May;64(5):1177–1183.

Available from: https://insights.ovid.com/crossref?an=00005373-200805000-

00005

143 4. Briggs A, Askari R. Damage control resuscitation. Elsevier; 2016 Sep 1;33:218–

221. Available from:

http://linkinghub.elsevier.com/retrieve/pii/S1743919116300243

5. Pohlman TH, Walsh M, Aversa J, Hutchison EM, Olsen KP, Lawrence Reed R.

Damage control resuscitation. Churchill Livingstone; 2015 Jul 1;29(4):251–262.

Available from:

https://www.sciencedirect.com/science/article/pii/S0268960X1500003X

6. Fries D, Martini WZ. Role of fibrinogen in trauma-induced coagulopathy. Oxford

University Press; 2010 Aug 1;105(2):116–121. Available from:

http://linkinghub.elsevier.com/retrieve/pii/S0007091217335298

7. RAHE-MEYER N, SØRENSEN B. Fibrinogen concentrate for management of

bleeding. Wiley/Blackwell (10.1111); 2011 Jan 1;9(1):1–5. Available from:

http://doi.wiley.com/10.1111/j.1538-7836.2010.04099.x

8. Levy JH, Goodnough LT. How I use fibrinogen replacement therapy in acquired

bleeding. American Society of Hematology; 2015 Feb 26;125(9):1387–93.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/25519751 PMID: 25519751

9. Curry N, Rourke C, Davenport R, Stanworth S, Brohi K. Fibrinogen replacement

in trauma haemorrhage. BioMed Central; 2014 Jul 7;22(Suppl 1):A5. Available

from: http://sjtrem.biomedcentral.com/articles/10.1186/1757-7241-22-S1-A5

10. Bishop P, Lawson J. Recombinant biologics for treatment of bleeding disorders.

Nature Publishing Group; 2004 Aug 1;3(8):684–694. Available from:

http://www.nature.com/articles/nrd1443

11. EVATT BL, FARRUGIA A, SHAPIRO AD, WILDE JT. Haemophilia 2002:

144 emerging risks of treatment. Wiley/Blackwell (10.1111); 2002 May 1;8(3):221–

229. Available from: http://doi.wiley.com/10.1046/j.1365-2516.2002.00612.x

12. Liumbruno G, Bennardello F, Lattanzio A, Piccoli P, Rossetti G, Italian Society of

Transfusion Medicine and Immunohaematology (SIMTI) Work Group as IS of

TM and I (SIMTI) W. Recommendations for the transfusion of plasma and

platelets. SIMTI Servizi; 2009 Apr;7(2):132–50. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/19503635 PMID: 19503635

13. Yang L, Stanworth S, Hopewell S, Doree C, Murphy M. Is fresh-frozen plasma

clinically effective? An update of a systematic review of randomized controlled

trials (CME). Wiley/Blackwell (10.1111); 2012 Aug 1;52(8):1673–1686.

Available from: http://doi.wiley.com/10.1111/j.1537-2995.2011.03515.x

14. Arya RC, Wander G, Gupta P. Blood component therapy: Which, when and how

much. Wolters Kluwer -- Medknow Publications; 2011 Apr;27(2):278–84.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/21772701 PMID: 21772701

15. Klein HG. Allogeneic transfusion risks in the surgical patient. Elsevier; 1995 Dec

1;170(6A Suppl):21S–26S. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/8546242 PMID: 8546242

16. Goodnough LT, Brecher ME, Kanter MH, AuBuchon JP. Transfusion Medicine —

Blood Transfusion. Massachusetts Medical Society ; 1999 Feb 11;340(6):438–

447. Available from:

http://www.nejm.org/doi/abs/10.1056/NEJM199902113400606

17. Midathada M V, Mehta P, Waner M, Fink LM. Recombinant Factor VIIa in the

Treatment of Bleeding. 2004;121(1):124–137. Available from:

145 http://www.embase.com/search/results?subaction=viewrecord&from=export&id=

L38090163%5Cnhttp://dx.doi.org/10.1309/D0G0-C96V-05CJ-5B4J PMID:

14750250

18. Gerotziafas GT, Zervas C, Gavrielidis G, Tokmaktsis A, Hatjiharissi E,

Papaioannou M, Lazaridou A, Constantinou N, Samama MM, Christakis J.

Effective hemostasis with rFVIIa treatment in two patients with severe

thrombocytopenia and life-threatening hemorrhage. Wiley-Blackwell; 2002 Mar

1;69(3):219–222. Available from: http://doi.wiley.com/10.1002/ajh.10056

19. Pipe SW. Recombinant clotting factors. 2008;99:840–850. Available from:

www.thrombosis-online.com

20. Ofosu FA, Freedman J, Semple JW. Plasma-derived biological used to

promote haemostasis. Schattauer GmbH; 2008 Nov 30;99(11):851–862. Available

from: http://www.thieme-connect.de/DOI/DOI?10.1160/TH07-10-0592

21. PIPE SW. The promise and challenges of bioengineered recombinant clotting

factors. Wiley/Blackwell (10.1111); 2005 Aug 1;3(8):1692–1701. Available from:

http://doi.wiley.com/10.1111/j.1538-7836.2005.01367.x

22. Toole JJ, Knopf JL, Wozney JM, Sultzman LA, Buecker JL, Pittman DD,

Kaufman RJ, Brown E, Shoemaker C, Orr EC, Amphlett GW, Foster WB, Coe M

Lou, Knutson GJ, Fass DN, Hewick RM. Molecular cloning of a cDNA encoding

human antihaemophilic factor. 1984 Nov 22;312(5992):342–347. Available from:

http://www.nature.com/articles/312342a0

23. Lusher JM, Arkin S, Abildgaard CF, Schwartz RS. Recombinant Factor VIII for

the Treatment of Previously Untreated Patients with Hemophilia A -- Safety,

146 Efficacy, and Development of Inhibitors. 1993 Feb 18;328(7):453–459. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/8421474 PMID: 8421474

24. Kurachi K, Davie EW. Isolation and characterization of a cDNA coding for human

factor IX. National Academy of Sciences; 1982 Nov 1;79(21):6461–4. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/6959130 PMID: 6959130

25. Kaufman RJ, Wasley LC, Furie BC, Furie B, Shoemaker CB. Expression,

purification, and characterization of recombinant gamma-carboxylated factor IX

synthesized in Chinese hamster ovary cells. 1986 Jul 25;261(21):9622–8.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/3733688 PMID: 3733688

26. White GC, Beebe A, Nielsen B. Recombinant factor IX. 1997 Jul;78(1):261–5.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/9198163 PMID: 9198163

27. Butenas S, Brummel KE, Branda RF, Paradis SG, Mann KG. Mechanism of factor

VIIa-dependent coagulation in hemophilia blood. 2002;99(3):923–930. PMID:

11806995

28. Allen GA, Monroe DM, Roberts HR, Hoffman M. The effect of factor X level on

thrombin generation and the procoagulant effect of activated factor VII in a cell-

based model of coagulation. 2000;11(4 SUPPL. 1):3–7. PMID: 10850556

29. Hagen FS, Gray CL, O’Hara P, Grant FJ, Saari GC, Woodbury RG, Hart CE,

Insley M, Kisiel W, Kurachi K. Characterization of a cDNA coding for human

factor VII. National Academy of Sciences; 1986 Apr 1;83(8):2412–6. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/3486420 PMID: 3486420

30. O’Hara PJ, Grant FJ, Haldeman BA, Gray CL, Insley MY, Hagen FS, Murray MJ.

Nucleotide sequence of the gene coding for human factor VII, a -

147 dependent protein participating in blood coagulation. National Academy of

Sciences; 1987 Aug 1;84(15):5158–62. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/3037537 PMID: 3037537

31. Hedner U. Dosing with based on current evidence.

2004;41(1 Suppl 1):35–39. Available from:

http://ovidsp.ovid.com/ovidweb.cgi?T=JS&PAGE=reference&D=med5&NEWS=

N&AN=14872419

32. Hedner U, Lee CA. First 20 years with recombinant FVIIa (NovoSeven).

2011;17(1):172–182. PMID: 20609014

33. Levi M, Levy JH, Andersen HF, Truloff D. Safety of Recombinant Activated

Factor VII in in Randomized Clinical Trials. 2010;1791–1800.

34. Björses K, Holst J. Various Local Hemostatic Agents with Different Modes of

Action; an in vivo Comparative Randomized Vascular Surgical Experimental

Study. W.B. Saunders; 2007 Mar 1;33(3):363–370. Available from:

https://www.sciencedirect.com/science/article/pii/S1078588406005879

35. Björses K, Holst J. Topical haemostatics in renal trauma-an evaluation of four

different substances in an experimental setting [Internet]. 2009. Available from:

http://www.scielo.br/pdf/ibju/v35n3/v35n3a22

36. Dutton RP, Parr M, Tortella BJ, Champion HR, Bernard GR, Boffard K, Bouillon

B, Croce MA, Dimsits J, Holcomb JB, Leppaniemi A, Vincent J-L, Hauser CJ.

Recombinant Activated Factor VII Safety in Trauma Patients: Results From the

CONTROL Trial. 2011 Jul;71(1):12–19. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=00

148 005373-201107000-00003

37. Mitra B, Cameron PA, Parr MJ, Phillips L. Recombinant factor VIIa in trauma

patients with the “triad of death.” Elsevier Ltd; 2012;43(9):1409–1414. Available

from: http://dx.doi.org/10.1016/j.injury.2011.01.033 PMID: 21345431

38. Boffard KD, Riou B, Warren B, Choong PIT, Rizoli S, Rossaint R, Axelsen M,

Kluger Y. Recombinant Factor VIIa as Adjunctive Therapy for Bleeding Control

in Severely Injured Trauma Patients: Two Parallel Randomized, Placebo-

Controlled, Double-Blind Clinical Trials. 2005 Jul;59(1):8–18. Available from:

https://insights.ovid.com/crossref?an=00005373-200507000-00002

39. Mayo A, Misgav M, Kluger Y, Geenberg R, Pauzner D, Klausner J, Ben-Tal O.

Recombinant activated factor VII (NovoSevenTM): Addition to replacement

therapy in acute, uncontrolled and life-threatening bleeding. 2004;87(1):34–40.

PMID: 15260820

40. Smith S, Mutch N, Baskar D-. Polyphosphate modulates blood coagulation and

fibrinolysis. 2006; Available from: http://www.pnas.org/content/103/4/903.short

41. Morrissey J, Choi S, Smith S. Polyphosphate: an ancient molecule that links

platelets, coagulation and inflammation. 2012; Available from:

http://www.bloodjournal.org/content/early/2012/04/18/blood-2012-03-

306605.short

42. Morrissey JH. Polyphosphate: a link between platelets, coagulation and

inflammation. Springer Japan; 2012 Apr 5;95(4):346–352. Available from:

http://link.springer.com/10.1007/s12185-012-1054-5

43. Kudela D, Smith SA, May-Masnou A, Braun GB, Pallaoro A, Nguyen CK,

149 Chuong TT, Nownes S, Allen R, Parker NR, Rashidi HH, Morrissey JH, Stucky

GD. Clotting Activity of Polyphosphate-Functionalized Silica Nanoparticles.

Wiley-Blackwell; 2015 Mar 23;54(13):4018–4022. Available from:

http://doi.wiley.com/10.1002/anie.201409639

44. Baylis JR, Yeon JH, Thomson MH, Kazerooni A, Wang X, St. John AE, Lim EB,

Chien D, Lee A, Zhang JQ, Piret JM, Machan LS, Burke TF, White NJ, Kastrup

CJ. Self-propelled particles that transport cargo through flowing blood and halt

hemorrhage. American Association for the Advancement of Science; 2015 Oct

2;1(9):e1500379–e1500379. Available from:

http://advances.sciencemag.org/cgi/doi/10.1126/sciadv.1500379

45. Chapin J, Hajjar K. Fibrinolysis and the control of blood coagulation. 2015;

Available from:

https://www.sciencedirect.com/science/article/pii/S0268960X14000721

46. Walsh M, Shreve J, Thomas S, Moore E, Moore H, Hake D, Pohlman T, Davis P,

Ploplis V, Piscoya A, Wegner J, Bryant J, Crepinsek A, Lantry J, Sheppard F,

Castellino F. Fibrinolysis in Trauma: “Myth,” “Reality,” or “Something in

Between.” Thieme Medical Publishers; 2017 Feb 20;43(02):200–212. Available

from: http://www.thieme-connect.de/DOI/DOI?10.1055/s-0036-1597900

47. OKAMOTO S, OKAMOTO U. AMINO-METHYL-CYCLOHEXANE-

CARBOXYLIC ACID: AMCHA. The Keio Journal of Medicine; 1962;11(3):105–

115. Available from:

http://joi.jlc.jst.go.jp/JST.Journalarchive/kjm1952/11.105?from=CrossRef

48. Okamoto S, Hijikata-Okunomiya A, Wanaka K, Okada Y, Okamoto U. Enzyme-

150 Controlling Medicines: Introduction. Copyright © 1997 by Thieme Medical

Publishers, Inc.; 1997 Dec 8;23(06):493–501. Available from: http://www.thieme-

connect.de/DOI/DOI?10.1055/s-2007-996127

49. CRASH-2 trial collaborators H, Shakur H, Roberts I, Bautista R, Caballero J,

Coats T, Dewan Y, El-Sayed H, Gogichaishvili T, Gupta S, Herrera J, Hunt B,

Iribhogbe P, Izurieta M, Khamis H, Komolafe E, Marrero M-A, Mejía-Mantilla J,

Miranda J, Morales C, Olaomi O, Olldashi F, Perel P, Peto R, Ramana P V, Ravi

RR, Yutthakasemsunt S. Effects of tranexamic acid on death, vascular occlusive

events, and blood transfusion in trauma patients with significant haemorrhage

(CRASH-2): a randomised, placebo-controlled trial. 2010 Jul 3;376(9734):23–32.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/20554319 PMID: 20554319

50. Roberts I, Prieto-Merino D. Applying results from clinical trials: tranexamic acid

in trauma patients. BioMed Central; 2014 Dec 5;2(1):56. Available from:

http://www.jintensivecare.com/content/2/1/56

51. Morrison JJ, Dubose JJ, Rasmussen TE, Midwinter MJ. Military Application of

Tranexamic Acid in Trauma Emergency Resuscitation (MATTERs) Study.

American Medical Association; 2012 Feb 1;147(2):113. Available from:

http://archsurg.jamanetwork.com/article.aspx?doi=10.1001/archsurg.2011.287

52. Brown JB, Neal MD, Guyette FX, Peitzman AB, Billiar TR, Zuckerbraun BS,

Sperry JL. Design of the Study of Tranexamic Acid during Air Medical

Prehospital Transport (STAAMP) Trial: Addressing the Knowledge Gaps. NIH

Public Access; 2015;19(1):79–86. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/25076119 PMID: 25076119

151 53. Lam WA, Chaudhuri O, Crow A, Webster KD, Li T-D, Kita A, Huang J, Fletcher

DA. Mechanics and contraction dynamics of single platelets and implications for

clot stiffening. Nature Publishing Group; 2011 Jan 5;10(1):61–66. Available from:

http://www.nature.com/articles/nmat2903

54. Jen CJ, McIntire L V. The structural properties and contractile force of a clot.

1982;2(5):445–455. Available from: http://doi.wiley.com/10.1002/cm.970020504

55. Brown AC, Stabenfeldt SE, Ahn B, Hannan RT, Dhada KS, Herman ES, Stefanelli

V, Guzzetta N, Alexeev A, Lam W a, Lyon LA, Barker TH. Ultrasoft microgels

displaying emergent platelet-like behaviours. 2014;13(September):1–7. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/25194701 PMID: 25194701

56. Nikolajsen CL, Dyrlund TF, Poulsen ET, Enghild JJ, Scavenius C. Coagulation

factor XIIIa substrates in human plasma: identification and incorporation into the

clot. American Society for Biochemistry and Molecular Biology; 2014 Mar

7;289(10):6526–34. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/24443567 PMID: 24443567

57. Muszbek L, Bereczky Z, Bagoly Z, Komáromi I, Katona É. Factor XIII: A

Coagulation Factor With Multiple Plasmatic and Cellular Functions. 2011

Jul;91(3):931–972. Available from:

http://www.physiology.org/doi/10.1152/physrev.00016.2010

58. Chan LW, Wang X, Wei H, Pozzo LD, White NJ, Pun SH. A synthetic fibrin

cross-linking polymer for modulating clot properties and inducing hemostasis.

2015;7(277):1–11 (online). Available from:

http://stm.sciencemag.org/cgi/doi/10.1126/scitranslmed.3010383 PMID: 25739763

152 59. Kolodziej AF, Nair SA, Graham P, McMurry TJ, Ladner RC, Wescott C, Sexton

DJ, Caravan P. Fibrin Specific Peptides Derived by Phage Display:

Characterization of Peptides and Conjugates for Imaging. American Chemical

Society; 2012 Mar 21;23(3):548–556. Available from:

http://pubs.acs.org/doi/10.1021/bc200613e

60. Gillespie CJ, Sutherland AD, Mossop PJ, Woods RJ, Keck JO, Heriot AG.

Mesenteric Embolization for Lower Gastrointestinal Bleeding. 2010

Sep;53(9):1258–1264. Available from:

https://insights.ovid.com/crossref?an=00003453-201009000-00005

61. Lubarsky M, Ray C, Funaki B. Embolization agents-which one should be used

when? Part 2: small-vessel embolization. Thieme Medical Publishers; 2010

Mar;27(1):99–104. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/21359018 PMID: 21359018

62. Vaidya S, Tozer KR, Chen J. An overview of embolic agents. Thieme Medical

Publishers; 2008 Sep;25(3):204–15. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/21326511 PMID: 21326511

63. Panagiotopoulos V, Gizewski E, Asgari S, Regel J, Forsting M, Wanke I.

Embolization of intracranial arteriovenous malformations with ethylene-vinyl

alcohol copolymer (Onyx). American Journal of ; 2009 Jan

1;30(1):99–106. Available from: http://www.ncbi.nlm.nih.gov/pubmed/18842759

PMID: 18842759

64. Thiex R, Williams A, Smith E, Scott RM, Orbach DB. The use of Onyx for

embolization of central nervous system arteriovenous lesions in pediatric patients.

153 American Journal of Neuroradiology; 2010 Jan 1;31(1):112–20. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/19749215 PMID: 19749215

65. Bakar B, Oruckaptan HH, Hazer BD, Saatci I, Atilla P, Kılıc K, Muftuoglu SF.

Evaluation of the toxicity of onyx compared with n-butyl 2-cyanoacrylate in the

subarachnoid space of a rabbit model: an experimental research. Springer-Verlag;

2010 Feb 15;52(2):125–134. Available from:

http://link.springer.com/10.1007/s00234-009-0594-8

66. Gaharwar AK, Avery RK, Assmann A, Paul A, McKinley GH, Khademhosseini

A, Olsen BD. Shear-Thinning Nanocomposite Hydrogels for the Treatment of

Hemorrhage. American Chemical Society; 2014 Oct 28;8(10):9833–9842.

Available from: http://pubs.acs.org/doi/10.1021/nn503719n

67. Avery RK, Albadawi H, Akbari M, Zhang YS, Duggan MJ, Sahani D V, Olsen

BD, Khademhosseini A, Oklu R. An injectable shear-thinning biomaterial for

endovascular embolization. American Association for the Advancement of

Science; 2016 Nov 16;8(365):365ra156. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/27856795 PMID: 27856795

68. Fedel M, Endogan T, Hasirci N, Maniglio D, Morelli A, Chiellini F, Motta A.

Blood compatibility of polymers derived from natural materials. SAGE

PublicationsSage UK: London, England; 2012 Jul 18;27(4):295–312. Available

from: http://journals.sagepub.com/doi/10.1177/0883911512446060

69. Gryshchuk V, Galagan N. Silica Nanoparticles Effects on Blood Coagulation

Proteins and Platelets. 2016;2016:2959414. Available from:

http://www.hindawi.com/journals/bri/2016/2959414/ PMID: 26881078

154 70. Ruggeri ZM, Mendolicchio GL. Adhesion Mechanisms in Platelet Function. 2007

Jun 22;100(12):1673–1685. Available from:

http://circres.ahajournals.org/cgi/doi/10.1161/01.RES.0000267878.97021.ab

71. Varga-Szabo D, Pleines I, Nieswandt B. Cell Adhesion Mechanisms in Platelets.

2008 Mar 1;28(3):403–412. Available from:

http://atvb.ahajournals.org/cgi/doi/10.1161/ATVBAHA.107.150474

72. Andrews RK, Berndt MC. Platelet physiology and thrombosis. Pergamon; 2004

Jan 1;114(5–6):447–453. Available from:

https://www.sciencedirect.com/science/article/pii/S0049384804004220

73. Arnout J, Hoylaerts MF, Lijnen HR. Haemostasis. Springer Berlin Heidelberg;

2006. p. 1–41. Available from: http://link.springer.com/10.1007/3-540-36028-X_1

74. Bennett JS, Shattil SJ, Power JW, Gartner TK. Interaction of fibrinogen with its

platelet receptor. Differential effects of alpha and gamma chain fibrinogen

peptides on the glycoprotein IIb-IIIa complex. 1988 Sep 15;263(26):12948–53.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/3417645 PMID: 3417645

75. Fuss C, Palmaz JC, Sprague EA. Fibrinogen: Structure, Function, and Surface

Interactions. 2001;12(6):677–682.

76. Blann A, Nadar SK, Lip GYH. The adhesion molecule P-selectin and

cardiovascular disease. Oxford University Press; 2003 Dec 1;24(24):2166–2179.

Available from: https://academic.oup.com/eurheartj/article-

lookup/doi/10.1016/j.ehj.2003.08.021

77. Andre P. P-selectin in haemostasis. Wiley/Blackwell (10.1111); 2004 Aug

1;126(3):298–306. Available from: http://doi.wiley.com/10.1111/j.1365-

155 2141.2004.05032.x

78. Etchill EW, Myers SP, Raval JS, Hassoune A, SenGupta A, Neal MD. Platelet

Transfusion in Critical Care and Surgery. 2017 May;47(5):537–549. Available

from: http://insights.ovid.com/crossref?an=00024382-201705000-00002

79. Heal JM, Blumberg N. Optimizing platelet transfusion therapy. Churchill

Livingstone; 2004 Sep 1;18(3):149–165. Available from:

https://www.sciencedirect.com/science/article/pii/S0268960X03000572

80. Boscarino C, Tien H, Acker J, Callum J, Hansen AL, Engels P, Glassberg E,

Nathens A, Beckett A. Feasibility and transport of packed red blood cells into

Special Forces operational conditions. 2014 Apr;76(4):1013–1019. Available

from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201404000-00015

81. Spinella PC, Dunne J, Beilman GJ, O’Connell RJ, Borgman MA, Cap AP, Rentas

F. Constant challenges and evolution of US military transfusion medicine and

blood operations in combat. Wiley/Blackwell (10.1111); 2012 May 1;52(5):1146–

1153. Available from: http://doi.wiley.com/10.1111/j.1537-2995.2012.03594.x

82. Solheim BG. Pathogen reduction of blood components. Pergamon; 2008 Aug

1;39(1):75–82. Available from:

https://www.sciencedirect.com/science/article/pii/S1473050208000888

83. Seghatchian J, de Sousa G. Pathogen-reduction systems for blood components: the

current position and future trends. 2006 Dec;35(3):189–96. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17110168 PMID: 17110168

156 84. Picker SM. Pathogen Reduction Technologies: The Best Solution for Safer Blood?

2012;3:133. Available from:

https://pdfs.semanticscholar.org/226b/64bc04e67f1a24573ce604aaae25ef35f327.p

df

85. Bode AP, Read MS. Lyophilized platelets: continued development. Pergamon;

2000 Feb 1;22(1–2):99–105. Available from:

https://www.sciencedirect.com/science/article/pii/S095538860000028X

86. Hoffmeister KM, Josefsson EC, Isaac NA, Clausen H, Hartwig JH, Stossel TP.

Glycosylation restores survival of chilled blood platelets. American Association

for the Advancement of Science; 2003 Sep 12;301(5639):1531–4. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/12970565 PMID: 12970565

87. Jhang JS, Spitalnik SL. Glycosylation and cold platelet storage. 2005

Nov;4(6):483–7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/16232387

PMID: 16232387

88. Bode AP, Fischer TH. Lyophilized Platelets: Fifty Years in the Making. Taylor &

Francis; 2007 Jan 24;35(1):125–133. Available from:

http://www.tandfonline.com/doi/full/10.1080/10731190600974962

89. Cap AP, Perkins JG. Lyophilized Platelets: Challenges and Opportunities. 2011

May;70:S59–S60. Available from:

https://insights.ovid.com/crossref?an=00005373-201105001-00022

90. Lozano M. Platelets come back in from the cold. American Society of

Hematology; 2008 Mar 15;111(6):2951–2951. Available from:

http://www.bloodjournal.org/cgi/doi/10.1182/blood-2008-01-131920

157 91. Reddoch KM, Pidcoke HF, Montgomery RK, Fedyk CG, Aden JK,

Ramasubramanian AK, Cap AP. Hemostatic Function of Apheresis Platelets

Stored at 4°C and 22°C. 2014 May;41:54–61. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=00

024382-201405001-00012

92. Nair PM, Pidcoke HF, Cap AP, Ramasubramanian AK. Effect of cold storage on

shear-induced platelet aggregation and clot strength. 2014 Sep;77:S88–S93.

Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201409002-00007

93. Nasiri S. Infusible platelet membrane as a platelet substitute for transfusion: an

overview. SIMTI Servizi; 2013 Jul;11(3):337–42. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/23736926 PMID: 23736926

94. Chao FC, Kim BK, Houranieh AM, Liang FH, Konrad MW, Swisher SN, Tullis

JL. Infusible platelet membrane microvesicles: a potential transfusion substitute

for platelets. 1996 Jun;36(6):536–42. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/8669086 PMID: 8669086

95. Biró E, Akkerman JWN, Hoek FJ, Gorter G, Pronk LM, Sturk A, Nieuwland R.

The phospholipid composition and cholesterol content of platelet-derived

microparticles: a comparison with platelet membrane fractions. 2005

Dec;3(12):2754–63. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/16359513 PMID: 16359513

96. Graham SS, Gonchoroff NJ, Miller JL. Infusible Platelet Membranes Retain

158 Partial Functionality of the Platelet GPIb/IX/V Receptor Complex. Oxford

University Press; 2001 Jan 1;115(1):144–147. Available from:

https://academic.oup.com/ajcp/article-lookup/doi/10.1309/CCDV-3BEP-XXKP-

BKDM

97. Nasiri S, Heidari M, Rivandi S. INFUSIBLE PLATELET MEMBRANES

IMPROVE HEMOSTASIS IN THROMBOCYTOPENIC RABBITS: STUDIES

WITH TWO DIFFERENT INJECTION DOSES. 2012;3(12). Available from:

www.ijpsr.com

98. Galán AM, Bozzo J, Hernández MR, Pino M, Reverter JC, Mazzara R, Escolar G,

Ordinas A. Infusible platelet membranes improve hemostasis in thrombocytopenic

blood: experimental studies under flow conditions. 2000 Sep;40(9):1074–80.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/10988310 PMID: 10988310

99. Fitzpatrick GM, Cliff R, Tandon N. Thrombosomes: a platelet-derived hemostatic

agent for control of noncompressible hemorrhage. 2013 Jan;53 Suppl 1:100S–

106S. Available from: http://www.ncbi.nlm.nih.gov/pubmed/23301961 PMID:

23301961

100. Sum R, Hager S, Pietramaggiori G, Orgill DP, Dee J, Rudolph A, Orser C,

Fitzpatrick GM, Ho D. Wound-healing properties of trehalose-stabilized freeze-

dried outdated platelets. 2007 Apr;47(4):672–9. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17381626 PMID: 17381626

101. Hu C-MJ, Fang RH, Wang K-C, Luk BT, Thamphiwatana S, Dehaini D, Nguyen

P, Angsantikul P, Wen CH, Kroll A V., Carpenter C, Ramesh M, Qu V, Patel SH,

Zhu J, Shi W, Hofman FM, Chen TC, Gao W, Zhang K, Chien S, Zhang L.

159 Nanoparticle biointerfacing by platelet membrane cloaking. Nature Publishing

Group; 2015 Oct 16;526(7571):118–121. Available from:

http://www.nature.com/articles/nature15373

102. Hu Q, Sun W, Qian C, Wang C, Bomba HN, Gu Z. Anticancer Platelet-Mimicking

Nanovehicles. Wiley-Blackwell; 2015 Nov 1;27(44):7043–7050. Available from:

http://doi.wiley.com/10.1002/adma.201503323

103. Li J, Ai Y, Wang L, Bu P, Sharkey CC, Wu Q, Wun B, Roy S, Shen X, King MR.

Targeted drug delivery to circulating tumor cells via platelet membrane-

functionalized particles. Elsevier; 2016 Jan 1;76:52–65. Available from:

https://www.sciencedirect.com/science/article/pii/S014296121500856X

104. Modery-Pawlowski CL, Kuo H-H, Baldwin WM, Gupta A Sen. A platelet-inspired

paradigm for nanomedicine targeted to multiple diseases. Future Medicine Ltd

London, UK ; 2013 Oct 30;8(10):1709–1727. Available from:

https://www.futuremedicine.com/doi/10.2217/nnm.13.113

105. Rybak MEM, Renzulli LA. A Based Platelet Substitute, the

Plateletsome, with Hemostatic Efficacy. Taylor & Francis; 1993 Jan 11;21(2):101–

118. Available from:

http://www.tandfonline.com/doi/full/10.3109/10731199309117350

106. Takeoka S, Teramura Y, Okamura Y, Tsuchida E, Handa M, Ikeda Y. Rolling

properties of rGPIbα-conjugated phospholipid vesicles with different membrane

flexibilities on vWf surface under flow conditions. Academic Press; 2002 Aug

23;296(3):765–770. Available from:

https://www.sciencedirect.com/science/article/pii/S0006291X02009348

160 107. Nishiya T, Kainoh M, Murata M, Handa M, Ikeda Y. Reconstitution of adhesive

properties of human platelets in liposomes carrying both recombinant

glycoproteins Ia/IIa and Ib?? under flow conditions: Specific synergy of receptor-

ligand interactions. 2002;100(1):136–142. PMID: 12070018

108. Nishiya T, Kainoh M, Murata M, Handa M, Ikeda Y. PLATELET

INTERACTIONS WITH LIPOSOMES CARRYING RECOMBINANT

PLATELET MEMBRANE GLYCOPROTEINS OR FIBRINOGEN: APPROACH

TO PLATELET SUBSTITUTES. 2001 Jan 11;29(6):453–464. Available from:

http://www.tandfonline.com/doi/full/10.1081/BIO-100108550

109. Takeoka S, Teramura Y, Ohkawa H, Ikeda Y, Tsuchida E. Conjugation of von

Willebrand factor-binding domain of platelet glycoprotein Ib alpha to size-

controlled albumin microspheres. 2000;1(2):290–5. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/11710113 PMID: 11710113

110. Versteeg HH, Heemskerk JWM, Levi M, Reitsma PH. New Fundamentals in

Hemostasis. American Physiological Society Bethesda, MD; 2013 Jan;93(1):327–

358. Available from: http://www.physiology.org/doi/10.1152/physrev.00016.2011

111. Roberts H, Hoffman M, Monroe D. A Cell-Based Model of Thrombin Generation.

2006 Feb;32(S 1):032–038. Available from: http://www.thieme-

connect.de/DOI/DOI?10.1055/s-2006-939552

112. Doshi N, Orje JN, Molins B, Smith JW, Mitragotri S, Ruggeri ZM. Platelet

Mimetic Particles for Targeting Thrombi in Flowing Blood. Wiley-Blackwell;

2012 Jul 24;24(28):3864–3869. Available from:

http://doi.wiley.com/10.1002/adma.201200607

161 113. Coller B. Interaction of normal, thrombasthenic, and Bernard-Soulier platelets with

immobilized fibrinogen: defective platelet-fibrinogen interaction in

thrombasthenia. 1980;55(2). Available from:

http://www.bloodjournal.org/content/55/2/169.short

114. Agam G, Livne AA. Erythrocytes with covalently bound fibrinogen as a cellular

replacement for the treatment of thrombocytopenia. 1992 Feb;22(2):105–12.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/1572388 PMID: 1572388

115. Lee DH, Blajchman MA, Ho TWC, Yen RCK. In vitro characteristics of

Thrombospheres, a novel hemostatic agent. 1995;86(suppl 1):902a.

116. Yen RCK, Ho TWC, Blajchman MA. A novel approach to correcting the bleeding

associated with thormbocytopenia. 1995;35:41S.

117. Levi M, Friederich PW, Middleton S, de Groot PG, Wu YP, Harris R, Biemond

BJ, Heijnen HF, Levin J, ten Cate JW. Fibrinogen-coated albumin microcapsules

reduce bleeding in severely thrombocytopenic rabbits. 1999;5(1):107–11.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/9883848 PMID: 9883848

118. Takeoka S, Teramura Y, Okamura Y, Handa M, Ikeda Y, Tsuchida E. Fibrinogen-

conjugated albumin polymers and their interaction with platelets under flow

conditions. 2001;2(4):1192–7. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/11777392 PMID: 11777392

119. MOSESSON MW. Fibrinogen and fibrin structure and functions. Blackwell

Science Inc; 2005 Aug;3(8):1894–1904. Available from:

http://doi.wiley.com/10.1111/j.1538-7836.2005.01365.x

120. Du XP, Plow EF, Frelinger AL, O’Toole TE, Loftus JC, Ginsberg MH. Ligands

162 "activate" integrin alpha IIb beta 3 (platelet GPIIb-IIIa). 1991 May

3;65(3):409–16. Available from: http://www.ncbi.nlm.nih.gov/pubmed/2018974

PMID: 2018974

121. Bennett JS. Structural biology of glycoprotein IIb-IIIa. Elsevier; 1996 Jan

1;6(1):31–36. Available from:

https://www.sciencedirect.com/science/article/pii/1050173895001263

122. Coller BS, Springer KT, Beer JH, Mohandas N, Scudder LE, Norton KJ, West SM.

Thromboerythrocytes. In vitro studies of a potential autologous, semi-artificial

alternative to platelet transfusions. American Society for Clinical Investigation;

1992 Feb;89(2):546–55. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/1737845 PMID: 1737845

123. Takeoka S, Okamura Y, Teramura Y, Watanabe N, Suzuki H, Tsuchida E, Handa

M, Ikeda Y. Function of fibrinogen γ-chain dodecapeptide-conjugated latex beads

under flow. 2003;312(3):773–779.

124. Bertram JP, Williams C a, Robinson R, Segal SS, Flynn NT, Lavik EB.

Intravenous hemostat: to halt bleeding. 2009;1(11):11ra22. PMID:

20371456

125. Lashof-Sullivan MM, Shoffstall E, Atkins KT, Keane N, Bir C, VandeVord P,

Lavik EB. Intravenously administered nanoparticles increase survival following

blast trauma. National Academy of Sciences; 2014 Jul 15;111(28):10293–10298.

Available from: http://www.pnas.org/cgi/doi/10.1073/pnas.1406979111

126. Beer JH, Springer KT, Coller BS. Immobilized Arg-Gly-Asp (RGD) peptides of

varying lengths as structural probes of the platelet glycoprotein IIb/IIIa receptor.

163 1992 Jan 1;79(1):117–28. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/1728303 PMID: 1728303

127. Ruoslahti E. RGD and other recognition sequences for integrins. 1996;12:697–

715. Available from: http://www.ncbi.nlm.nih.gov/pubmed/8970741 PMID:

8970741

128. Humphries JD, Byron A, Humphries MJ. Integrin ligands at a glance. The

Company of Biologists Ltd; 2006 Oct 1;119(Pt 19):3901–3. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/16988024 PMID: 16988024

129. Okamura Y, Takeoka S, Teramura Y, Maruyama H, Tsuchida E, Handa M, Ikeda

Y. Hemostatic effects of fibrinogen gamma-chain dodecapeptide-conjugated

polymerized albumin particles in vitro and in vivo. Wiley/Blackwell (10.1111);

2005 Jul 1;45(7):1221–1228. Available from: http://doi.wiley.com/10.1111/j.1537-

2995.2005.00173.x

130. OKAMURA Y, TAKEOKA S, ETO K, MAEKAWA I, FUJIE T, MARUYAMA

H, IKEDA Y, HANDA M. Development of fibrinogen γ-chain peptide-coated,

adenosine diphosphate-encapsulated liposomes as a synthetic platelet substitute.

Wiley/Blackwell (10.1111); 2009 Mar 1;7(3):470–477. Available from:

http://doi.wiley.com/10.1111/j.1538-7836.2008.03269.x

131. Okamura Y, Fujie T, Nogawa M, Maruyama H, Handa M, Ikeda Y, Takeoka S.

Haemostatic effects of polymerized albumin particles carrying fibrinogen γ-chain

dodecapeptide as platelet substitutes in severely thrombocytopenic rabbits.

Wiley/Blackwell (10.1111); 2008 Jun 1;18(3):158–166. Available from:

http://doi.wiley.com/10.1111/j.1365-3148.2008.00860.x

164 132. Okamura Y, Fujie T, Maruyama H, Handa M, Ikeda Y, Takeoka S. Prolonged

hemostatic ability of polyethylene glycol-modified polymerized albumin particles

carrying fibrinogen γ-chain dodecapeptide. Wiley/Blackwell (10.1111); 2007 Jul

1;47(7):1254–1262. Available from: http://doi.wiley.com/10.1111/j.1537-

2995.2007.01265.x

133. Okamura Y, Maekawa I, Teramura Y, Maruyama H, Handa M, Ikeda Y, Takeoka

S. Hemostatic Effects of Phospholipid Vesicles Carrying Fibrinogen γ Chain

Dodecapeptide in Vitro and in Vivo. 2005 Nov;16(6):1589–1596. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/16287259 PMID: 16287259

134. del Carpio Munoz C, Campbell W, Constantinescu I, Gyongyossy-Issa MIC.

Rational design of antithrombotic peptides to target the von Willebrand Factor

(vWf) - GPIb integrin interaction. Springer-Verlag; 2008 Dec 16;14(12):1191–

1202. Available from: http://link.springer.com/10.1007/s00894-008-0375-z

135. Gyongyossy‐Issa MIC, Kizhakkedathu J, Constantinescu I, Campbell W, del

Carpio Munoz CA. Synthetic platelets [Internet]. US; CA2599849C, 2007.

Available from:

https://patents.google.com/patent/CA2599849C/en?oq=Gyongyossy‐

Issa%2C+Kizhakkedathu%2C+Constantinescu%2C+Campbell%2C+del+Carpio+

Munoz

165 Chapter 4: Designing an Ideal Platelet-inspired System for the Intravenous

Management of Traumatic Bleeding

Content based on: Hickman DA, Pawlowski CL, Sekhon UDS, Marks J, Gupta AS. Biomaterials and Advanced Technologies for Hemostatic Management of Bleeding. Adv Mater. 2017 Nov 22. doi: 10.1002/adma.201700859.

Shukla M, Sekhon UD, Betapudi V, Li W, Hickman DA, Pawlowski CL, Dyer MR, Neal MD, McCrae KR, Sen Gupta A. In vitro characterization of SynthoPlate™ (synthetic platelet) technology and its in vivo evaluation in severely thrombocytopenic mice. J Thromb Haemost. 2017 Feb;15(2):375-387. doi: 10.1111/jth.13579. PubMed PMID: 27925685; PubMed Central PMCID: PMC5305617.

4.1 Introduction

The past two chapters have been a comprehensive review of the various topical, intra-

cavitary and intravenous hemostatic technologies in terms of materials, mechanisms and

state-of-art, and discussed challenges and opportunities to help advancement of the field.

For externally accessible injuries, a variety of natural and synthetic biomaterials have

undergone robust research, leading to hemostatic technologies including glues, bandages,

tamponades, tourniquets, dressings and pro-coagulant powders. In contrast, treatment of

internal non-compressible hemorrhage still heavily depends on transfusion of whole blood

or blood’s hemostatic components (platelets, fibrinogen and coagulation factors).

Transfusion of platelets poses significant challenges of limited availability, high cost,

contamination risks, short shelf-life, low portability, performance variability and

immunological side-effects, while use of fibrinogen or coagulation factors provides only

partial mechanisms for hemostasis. With such considerations, significant interdisciplinary

research endeavors have been focused on developing materials and technologies that can

166 be manufactured conveniently, sterilized to minimize contamination and enhance shelf-

life, and administered intravenously to mimic, leverage and amplify physiological

hemostatic mechanisms. We have developed an I.V.-administrable ‘synthetic platelet’

nanoconstruct (SynthoPlateTM), that can mimic and amplify the body’s natural hemostatic

mechanisms specifically at the bleeding site while maintaining systemic safety. We have also begun to explore ways to exploit SynthoPlateTM’s lipid core for injury-site specific

delivery of hemostasis augmenting drugs such as Tranexamic Acid (TXA, anti-fibrinolytic

agent), recombinant coagulation factors, polyphosphates etc.

4.2 Current State of Traumatic Bleeding Management

While externally administered and intra-cavitary hemostatic materials and technologies

have significantly improved the treatment options for accessible injuries, emergency or

prophylactic management of internal non-compressible hemorrhage is still heavily

dependent on transfusion of whole blood or blood components (RBC, plasma and

platelets). These transfusion methods and products are essentially meant to compensate for

blood volume loss and to facilitate or augment physiological inherent mechanisms of

primary and secondary hemostasis. Along with blood transfusion, there are a few

pharmaceutical compounds like desmopressin, aprotinin and tranexamic acid (TXA) that

have been either approved or under clinical trial for intravenous administration to augment

hemostasis1. These pharmaceutical compounds act essentially as facilitators or stabilizers of hemostatic pathway components, but are not directly pro-hemostatic agents. For

example, desmopressin acts by mimicking the function of natural vasopressin to

systemically increase the plasma concentration of Factor VIII and vWF, and therefore can

167 be used to treat mild hemophilia and vWF disease2. However, meta-analysis reports of

controlled clinical trials with desmopressin have indicated that such systemic amplification

of coagulation facilitators may pose a risk of off-target thrombosis3,4. Aprotinin is a broad-

spectrum inhibitor of serine proteases, and therefore its use in hemostasis is debatable since

it can block pro-hemostatic factors (e.g. thrombin) as well as anti-coagulatory factors (e.g

activated protein C) and fibrinolytic factors (e.g. plasmin)5. Nonetheless, the use of this

agent has shown benefit in management of bleeding complications in several surgical

procedures6,7. As described previously, TXA does not promote clotting but rather maintains clot strength and stability by preventing its plasmin-induced breakdown8. This property has

shown benefits in preventing bleeding related morbidities and mortalities9. Along with the

use of these pharmaceutical agents, blood component (especially platelet) transfusion

remains the primary intravascular strategy in rendering hemostasis, and the use of blood

products present significant challenges due to limited availability, high cost of purification

and storage, contamination risks, and immunological side-effects. While efficient blood-

banking protocols and modern pathogen reduction technologies work relentlessly to

resolve some of these challenges in locations with suitable medical facilities, the challenges

persist in remote locations and austere battlefield conditions. This has triggered robust

research endeavors in development of semi-synthetic or synthetic materials-based systems

that can provide the hemostatic functions of blood components while circumventing the

issues associated with natural blood products. In the context of hemostasis, these research

endeavors are currently directed towards materials that capture the biology or functions of

platelets and the coagulation cascade components.

168 4.3 Designing an Ideal System

Prompt resuscitation and point-of-injury treatment of ongoing hemorrhage is of upmost importance to improve survival10,11. In recent years, resuscitation of bleeding trauma patients has shifted away from crystalloid fluids to balanced blood product transfusion12–

14. Important in the current resuscitation strategies is the transfusion of pooled platelets,

which have been shown to be a critical component of modern hemostatic resuscitation

strategies15,16. As mentioned in Chapter 1, the PAMPer trial prehospital adminstration of thawed plasma resulted in lower 30-day mortality and a lower median prothrombin time

than standard-care resuscitation while the PROPPR trial demonstrated that early platelet

administration is associated with improved hemostasis and reduced mortality in traumatic

bleeding patients compared to patients who received blood products without platelets17,18.

However, platelet transfusions are fraught with several challenges including increased risk

of bacterial contamination in storage, platelet activation/degranulation during storage, and

a limited shelf life of only 5 days to 7 days19–21. As well, platelet transfusions account for

25% to 30% of all transfusion reactions while only accounting for around 10% of blood

product transfusions19,22. Finally, given the challenging requirements for storage of pooled platelets, there is very limited availability of platelets outside of large central blood banks, which presents a major problem for hemorrhage resuscitation in remote civilian locations

and austere battlefield conditions19,22. This has prompted significant research in platelet

substitutes that may resolve the above issues while providing platelet-inspired hemostatic action.

169 For an ideal design of a platelet-inspired hemostatic particle, the injury site-selective adhesive and aggregatory mechanisms of platelets should be functionally mimicked and integrated on a biocompatible, intravenously injectable particle platform. To this end,

Takeoka and colleagues have demonstrated the possibility of combining the platelet- inspired pro-adhesive function with pro-aggregatory function by either co-conjugation of rGPIbα and H12 peptide on the same latex bead, or mixing rGPIbα-decorated latex beads with H12 peptide-decorated latex beads23. When the two motifs were conjugated onto the surface of the same bead, the large size of the rGPIbα motif sterically masked the action of the smaller H12 peptide, and therefore the effect seen was mostly of platelet-inspired adhesion but no aggregation (Figure 4-1). In contrast, when each motif was conjugated to separate beads and the beads were mixed together, the resultant mixture significantly enhanced the aggregation and surface-coverage of platelets on collagen-coated surfaces in presence of soluble vWF under shear flow conditions, when compared to H12 peptide- decorated beads only. These studies suggested that if platelet-relevant pro-adhesive and pro-aggregatory functionalities could be combined on a single particle platform without the issue of mutual steric interference, the resultant design would have superior hemostatic capacity compared to designs that bear only pro-adhesive or only pro-aggregatory function.

Figure 4-1. Platelet aggregometry studies where undecorated latex beads, or H12 peptide-decorated latex

170 beads, or latex beads co-decorated with H12 peptides and rGPIbα fragments were added to ADP-activated platelets and the results showed that the co-decorated beads had no additional aggregation enhancement effects, possibly due to the masking of small H12 peptides by the large rGPIbα fragments.

4.4 Design of SynthoPlateTM Nanoparticle System

Rationalizing from this idea, we have utilized the technique of heteromultivalent surface-

modification to decorate synthetic particle surfaces simultaneously with vWF-binding peptides (VBP), collagen-binding peptides (CBP) and active platelet GPIIb-IIIa-binding fibrinogen-mimetic peptides (FMP), to create a unique synthetic platelet design, the

SynthoPlateTM 24,25. The VBP, Thr-Arg-Tyr-Leu-Arg-Ile-His-Pro-Glu-Ser-Trp-Val-His-

Glu-Ile (TRYLRIHPQSWVHQI) was derived from the C2 domain (residues 2303-2332) of the coagulation factor FVIII, based on reports of FVIII binding domains on vWF26. The

CBP sequence was a single chain 7-mer repeat of the Glycine(G)- -Proline(P)-

Figure 4-2. Manufacture and characterization of SynthoPlate constructs.

171

Hydroxyproline(O) tri-peptide (i.e. -[GPO]7-) that has helicogenic affinity to fibrillar

collagen but minimal ability to activate platelets via platelet collagen receptors GPIa/IIa

and GPVI (hence no systemic thrombotic risk)27,28. The FMP sequence was the cyclic RGD

peptide cyclo-CNPRGDY[-OEt]RC, which was reported to have high affinity and

selectivity for active platelet GPIIb-IIIa but minimal affinity for resting GPIIb-IIIa (hence

reduced systemic thromboembolic risk) as well as minimal selectivity for other common

cell integrins (hence reduced cross-reactivity in vivo)29. Each peptide was conjugated to a

poly(ethylene glycol)-terminated lipid (i.e. PEG-lipid) molecule via amide, thioether or

‘click’ chemistry to form lipid-PEG-peptide conjugates25. Resultant conjugates could be

self-assembled and extruded into ~150nm unilamellar liposomal vesicles such that the

Figure 4-3. Ex vivo aggregometry studies for SynthoPlate with mouse platelets.

172 vesicle’s outer leaflet had heteromultivalent decorations of VBP, CBP and FMP motifs

(Figure 4-2). Due to the small size of the peptides and adequate PEG spacer length, they

did not sterically interfere with respective bioactivities. This integrative synthetic platelet

design (bearing VBP + CBP + FMP peptides together) was named SynthoPlateTM. The

SynthoPlateTM nanoparticles were able to demonstrate platelet-inspired hemostasis-

relevant bioactivities in vitro.

4.5 SynthoPlateTM In Vitro Evaluation

We observe, utilizing aggregometry studies with mouse platelets, that SynthoPlate does

not aggregate or activate resting platelets and that SynthoPlate can rescue aggregation

when added to diluted platelets (Figure 4-3). Looking more closely at the classical

secondary hemostasis pathway, thrombin generation assays were performed. These studies

demonstrated that SynthoPlate does not have the ability to rapidly and spontaneously

activate coagulation co-factors in plasma to generate thrombin. In the presence of activated platelets, it does slightly enhance the generation of thrombin likely by recruiting and clustering activated platelets (Figure 4-4A). In parallel plate flow chamber experiments

with plasma in the presence of platelet agonist (ADP), flowing over a ‘vWF + collagen’

surface, larger clots are generated in the presence of SynthoPlate (Figure 4-4B and C). D‐

dimer analysis of the generated/deposited clots on the ‘vWF + collagen’ surface confirms

that the clots resulting from enhanced clot formation in the SynthoPlate‐added group are

indeed rich in fibrin, and that the process does not just involve fibrinogen binding to

recruited active platelets (Figure 4-4D).

173

Figure 4-4. Effect of SynthoPlate on secondary hemostasis

4.6 In Vivo Evaluation of SynthoPlateTM In Thrombocytopenia Model

We showed that SynthoPlate significantly reduced tail transection bleeding time in normal mice at superior levels compared to particles decorated with pro-adhesive (VBP + CBP) peptides only or pro-aggregatory (FMP) peptides only30–32. We then

174

Figure 4-5. Development of thrombocytopenia model in mouse and evaluation of SynthoPlate capability in correcting tail transection bleeding time in thrombocytopenic mouse.

developed a thrombocytopenia mouse model to evaluate SynthoPlate capability in

correcting tail transection bleeding time in thrombocytopenic mouse. Mice underwent

dose‐dependent depletion of platelets resulting in a corresponding increase in bleeding time

after tail transection (10 μg of antibody cocktail dose resulted in ~90% platelet depletion and a corresponding four‐fold to five‐fold increase in bleeding time as compared with normal mice) (Figure 4-5A). Severely thrombocytopenic (TCP) mice injected with

175 phosphate‐buffered saline (PBS) or control (unmodified) particles at various doses show

no improvement (reduction) in bleeding time, but TCP mice injected with SynthoPlate

vesicles show a dose‐dependent significant improvement in bleeding time, with the highest

dose (1000 nL−1, equivalent to the normal murine platelet count) reducing the bleeding

Figure 4-6. Characterization of SynthoPlate‐mediated hemostatic clot in transected tail of thrombocytopenic mouse.

time to ~150s (close to the bleeding time of ~100s for normal mice) (Figure 4-5B).

Immunofluorescence analysis showed that the SynthoPlateTM particles could co-localize with injured endothelium and locally available active platelets to enhance primary hemostasis (platelet recruitment and aggregation) and could in effect also enhance

176 secondary hemostasis (fibrin generation) (Figure 4-6). In addition, it was demonstrated

that SynthoPlateTM was systemically safe (no systemic pro-thrombotic or pro-coagulant

risk) and could be eliminated mostly via liver over time (Figure 4-7)25. The size, shape and

stiffness aspects of platelets influence their vascular margination towards the wall under a hemodynamic flow environment33,34. Based upon such observations, the heteromultivalent

surface- modification strategy used for SynthoPlateTM

Figure 4-7. Evaluation of systemic pro‐coagulant risk and biodistribution of SynthoPlate in mice.

177

particles, could also be adapted to other particle platforms (e.g. discoid albumin particles)

to integrate platelet-relevant hemostatic surface chemistry with platelet’s fluid dynamic

margination-influencing biophysical (e.g. geometric) properties35.

4.7 Discussion

It is important to note here that platelet’s hemostasis-relevant surface mechanisms are driven by multiple interactions occurring in tandem, and therefore the design of a synthetic platelet mimic can significantly benefit from incorporating such heterotypic interactions via materials engineering. This aspect was first emphasized in studies where multiple recombinant protein fragments (e.g. rGPIbα that binds vWF and rGPIa/IIa that binds collagen) were combined on latex beads or albumin particles. The recombinant technology usually involves high cost, and the large size of these recombinant protein fragments often poses mutual steric interference to their respective bioactivity when conjugated together on micro- or nano-scale particles. Protein fragments may also pose risks of immunogenicity over time. These issues can be potentially resolved by identifying and utilizing small peptides instead of protein fragments that can perform the same platelet-inspired functions.

It was demonstrated in our recent reports that platelet’s hemostatically relevant adhesive and aggregatory functions could be efficiently captured via heteromultivalent modification of lipidic and polymeric particles with multiple types of small peptides. Such approaches may allow easy manufacture and scale up by leveraging modern technologies of automated peptide synthesis, purification and high throughput particle manufacture methods. These strategies can also incorporate biophysical components of particle design (e.g. shape, size,

178 flexibility etc.) that mimic the fluid dynamic behavior of platelets regarding their

interaction with the RBC volume and efficient margination to vessel wall. Exploiting such

biophysical parameters for particle design can open new doors not only for synthetic

platelet technologies, but also for efficient platforms in vascular drug delivery36.

While such heteromultivalent integrative designs demonstrate combinatorial functional advantages in mimicking platelet mechanisms, in advancement and translation of such technologies, there may be potential challenges regarding large-scale manufacturing of such multi-component designs especially in the context of Good Manufacturing Practice

(GMP) scale process development37. Also, these liposome-based and albumin-based

particle designs as well as polymeric nanoparticle-based designs described previously, should be extensively studied for dose limitations in terms of toxicity (i.e. maximum tolerance), systemic risks, and possible immune response risks. Biodistribution is an important technique in formulating drugs as it assists in determining the pharmacokinetic profile and toxicity. There are limitations to utilizing any type of biodistribution protocol.

In Figure 4-7, data is presented that was obtained by intravenously injecting RhB labeled particles into the mice and allowing the particles to circulate for 2hrs. After this time, the mice were euthanized, the organs were collected and homogenized, and the RhB labeled

particles were extracted and quantified via UPLC. The concentration of the nanoparticles

(mg of nanoparticles/mg of tissue) in each organ,based off a calibration curve, and the

injected dose was utilized to determine the percent biodistribution in each organ.

Unfortunately, this protocol only allows us to determine the biodistribution at one time

point of a given animal and each sample must undergo considerable preparation to be

179 analyzed. In this same study, separate mice were injected with radiolabeled SynthoPlate

3 111 ( H tagged cholesteryl ester in the vesicle membrane or encapsulated InCl3 in the vesicle core). These mice were whole body imaged by single photon emission computed

111 tomography for assessment of the InCl3-labeled particle distribution or killed for

harvesting of organs and blood for scintillation counting of 3H-labeled particles. This

protocol allowed for real-time kinetic analysis of particle distribution in a single animal.

Unfortunately, the use of radioactive tracers in these protocols require extra institutional

approval, experimental constraints, safety precautions and waste processing.

In addition to biodistribution studies, in order to utilize the nanoparticle designs discussed

as intravenous hemostat technologies, it would be critical to carry out dose escalation

studies in appropriate in vivo models, along with systemic monitoring of hemodynamics

and vitals. For example, it has been demonstrated in the field of nanoparticle-based drug

delivery research that certain liposome-based formulations can result in complement- activation related pseudoallergy (CARPA) reactions38,39 depending on dose, particle size

and surface-charge and administration speed.

In the following chapters, I will describe the warranted studies that I carried out to validate

and characterize the efficacy and risk of our ‘synthetic platelet surrogate’ designs and the

methods established to enhance the efficacy and mitigate the risks of our designs.

4.8 Acknowledgments

This work was supported by the National Heart Lung and Blood Institute (NHLBI) of the

180 National Institutes of Health under award numbers R01 HL121212 (PI: Sen Gupta). The content of this publication is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

4.9 References

1. Ryan KL, Cortez DS, Dick EJ, Pusateri AE. Efficacy of FDA-approved hemostatic

drugs to improve survival and reduce bleeding in rat models of uncontrolled

hemorrhage. Elsevier; 2006 Jul 1;70(1):133–144. Available from:

https://www.sciencedirect.com/science/article/pii/S0300957205005083

2. Franchini M. The use of desmopressin as a hemostatic agent: A concise review.

Wiley-Blackwell; 2007 Aug 1;82(8):731–735. Available from:

http://doi.wiley.com/10.1002/ajh.20940

3. Laupacis A, Fergusson D. Drugs to minimize perioperative blood loss in cardiac

surgery: meta-analyses using perioperative blood transfusion as the outcome. The

International Study of Peri-operative Transfusion (ISPOT) Investigators. 1997

Dec;85(6):1258–67. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/9390590 PMID: 9390590

4. Wells PS. Safety and efficacy of methods for reducing perioperative allogeneic

transfusion: a critical review of the literature. 2002;9(5):377–88. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/12237729 PMID: 12237729

5. Levy JH. Hemostatic agents. 2004 Dec;44(S2):58S–62S. Available from:

http://doi.wiley.com/10.1111/j.0041-1132.2004.04173.x

6. Levi M, Cromheecke ME, de Jonge E, Prins MH, de Mol BJ, Briët E, Büller HR.

181 Pharmacological strategies to decrease excessive blood loss in : a

meta-analysis of clinically relevant endpoints. Elsevier; 1999 Dec

4;354(9194):1940–1947. Available from:

https://www.sciencedirect.com/science/article/pii/S0140673699012647

7. Capdevila X, Calvet Y, Biboulet P, Biron C, Rubenovitch J, d’Athis F. Aprotinin

decreases blood loss and homologous transfusions in patients undergoing major

. 1998 Jan;88(1):50–7. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/9447855 PMID: 9447855

8. Roberts I, Prieto-Merino D, Manno D. Mechanism of action of tranexamic acid in

bleeding trauma patients: an exploratory analysis of data from the CRASH-2 trial.

2014;18(6):1–5. Available from: http://ccforum.com/content/18/6/685 PMID:

25498484

9. Roberts I, Prieto-Merino D. Applying results from clinical trials: tranexamic acid

in trauma patients. BioMed Central; 2014 Dec 5;2(1):56. Available from:

http://www.jintensivecare.com/content/2/1/56

10. Sauaia A, Moore F, Moore E, Moser K. Epidemiology of trauma deaths: a

reassessment. 1995; Available from:

https://journals.lww.com/jtrauma/Abstract/1995/02000/Epidemiology_of_Trauma

_Deaths__A_Reassessment.6.aspx

11. Russo RM, Williams TK, Grayson JK, Lamb CM, Cannon JW, Clement NF,

Galante JM, Neff LP. Extending the golden hour. 2016 Mar;80(3):372–380.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/26670114 PMID: 26670114

12. Neal MD, Hoffman MK, Cuschieri J, Minei JP, Maier R V, Harbrecht BG, Billiar

182 TR, Peitzman AB, Moore EE, Cohen MJ, Sperry JL, Investigators TI and the HR

to I. Crystalloid to packed red blood cell transfusion ratio in the massively

transfused patient: when a little goes a long way. NIH Public Access; 2012

Apr;72(4):892–8. Available from: http://www.ncbi.nlm.nih.gov/pubmed/22491601

PMID: 22491601

13. Schöchl H, Maegele M, Solomon C, Görlinger K, Voelckel W. Early and

individualized goal-directed therapy for trauma-induced coagulopathy. BioMed

Central; 2012 Feb 24;20:15. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/22364525 PMID: 22364525

14. Kutcher ME, Kornblith LZ, Narayan R, Curd V, Daley AT, Redick BJ, Nelson

MF, Fiebig EW, Cohen MJ. A Paradigm Shift in Trauma Resuscitation. 2013 Sep

1;148(9):834. Available from: http://www.ncbi.nlm.nih.gov/pubmed/23864019

PMID: 23864019

15. Holcomb JB, Tilley BC, Baraniuk S, Fox EE, Wade CE, Podbielski JM, del Junco

DJ, Brasel KJ, Bulger EM, Callcut R a, Cohen MJ, Cotton B a, Fabian TC, Inaba

K, Kerby JD, Muskat P, O’Keeffe T, Rizoli S, Robinson BRH, Scalea TM,

Schreiber M a, Stein DM, Weinberg J a, Callum JL, Hess JR, Matijevic N, Miller

CN, Pittet J, Hoyt DB, Pearson GD, Leroux B, van Belle G. Transfusion of

Plasma, Platelets, and Red Blood Cells in a 1:1:1 vs a 1:1:2 Ratio and Mortality in

Patients With Severe Trauma. 2015;313:471. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/25647203 PMID: 25647203

16. Holcomb JB, del Junco DJ, Fox EE, Wade CE, Cohen MJ, Schreiber MA, Alarcon

LH, Bai Y, Brasel KJ, Bulger EM, Cotton BA, Matijevic N, Muskat P, Myers JG,

183 Phelan HA, White CE, Zhang J, Rahbar MH, PROMMTT Study Group. The

prospective, observational, multicenter, transfusion (PROMMTT)

study: comparative effectiveness of a time-varying treatment with competing risks.

NIH Public Access; 2013 Feb;148(2):127–36. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/23560283 PMID: 23560283

17. Cardenas JC, Zhang X, Fox EE, Cotton BA, Hess JR, Schreiber MA, Wade CE,

Holcomb JB, PROPPR Study Group. Platelet transfusions improve hemostasis and

survival in a substudy of the prospective, randomized PROPPR trial. American

Society of Hematology; 2018 Jul 24;2(14):1696–1704. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/30030268 PMID: 30030268

18. Sperry JL, Guyette FX, Brown JB, Yazer MH, Triulzi DJ, Early-Young BJ, Adams

PW, Daley BJ, Miller RS, Harbrecht BG, Claridge JA, Phelan HA, Witham WR,

Putnam AT, Duane TM, Alarcon LH, Callaway CW, Zuckerbraun BS, Neal MD,

Rosengart MR, Forsythe RM, Billiar TR, Yealy DM, Peitzman AB, Zenati MS.

Prehospital Plasma during Air Medical Transport in Trauma Patients at Risk for

Hemorrhagic Shock. 2018;379(4):315–326. Available from:

http://www.nejm.org/doi/10.1056/NEJMoa1802345 PMID: 30044935

19. Annen K, Olson JE. Optimizing platelet transfusions. 2015 Nov;22(6):559–564.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/26390163 PMID: 26390163

20. Spinella PC, Dunne J, Beilman GJ, O’Connell RJ, Borgman MA, Cap AP, Rentas

F. Constant challenges and evolution of US military transfusion medicine and

blood operations in combat. Wiley/Blackwell (10.1111); 2012 May 1;52(5):1146–

1153. Available from: http://doi.wiley.com/10.1111/j.1537-2995.2012.03594.x

184 21. Modery-Pawlowski CL, Tian LL, Pan V, McCrae KR, Mitragotri S, Sen Gupta A.

Approaches to synthetic platelet analogs. Elsevier Ltd; 2013;34(2):526–541.

Available from: http://linkinghub.elsevier.com/retrieve/pii/S0142961212010976

22. Garraud O, Cognasse F, Tissot J-D, Chavarin P, Laperche S, Morel P, Lefrère J-J,

Pozzetto B, Lozano M, Blumberg N, Osselaer J-C. Improving platelet transfusion

safety: biomedical and technical considerations. SIMTI Servizi; 2016

Mar;14(2):109–22. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/26674828 PMID: 26674828

23. Okamura Y, Handa M, Suzuki H, Ikeda Y, Takeoka S. New strategy of platelet

substitutes for enhancing platelet aggregation at high shear rates: cooperative

effects of a mixed system of fibrinogen γ-chain dodecapeptide- or glycoprotein

Ibα-conjugated latex beads under flow conditions. Springer-Verlag; 2006 Dec

21;9(4):251–258. Available from: http://link.springer.com/10.1007/s10047-006-

0345-0

24. Sen Gupta A, Ravikumar M, Modery‐Pawlowski C. Synthetic platelets [Internet].

US; US9107845B2, 2012. Available from:

https://patents.google.com/patent/US9107845?oq=9107845%2C+2015

25. Shukla M, Sekhon UDS, Betapudi V, Li W, Hickman DA, Pawlowski CL, Dyer

MR, Neal MD, McCrae KR, Sen Gupta A. In vitro characterization of

SynthoPlateTM (synthetic platelet) technology and its in vivo evaluation in severely

thrombocytopenic mice. 2017;15(2):375–387. PMID: 27925685

26. Nogami K, Shima M, Giddings JC, Takeyama M, Tanaka I, Yoshioka A.

Relationship between the binding sites for von Willebrand factor, phospholipid,

185 and human factor VIII C2 inhibitor alloantibodies within the factor VIII C2

domain. 2007 May;85(4):317–22. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17483075 PMID: 17483075

27. Mo X, An Y, Yun C-S, Yu SM. Nanoparticle-Assisted Visualization of Binding

Interactions between Collagen Mimetic Peptide and Collagen Fibers. Wiley-

Blackwell; 2006 Mar 27;118(14):2325–2328. Available from:

http://doi.wiley.com/10.1002/ange.200504529

28. Cejas MA, Chen C, Kinney WA, Maryanoff BE. Nanoparticles That Display Short

Collagen-Related Peptides. Potent Stimulation of Human Platelet Aggregation by

Triple Helical Motifs. 2007 Jul;18(4):1025–1027. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17583929 PMID: 17583929

29. Cheng S, Craig WS, Mullen D, Tschopp JF, Dixon D, Pierschbacher MD. Design

and synthesis of novel cyclic RGD-containing peptides as highly potent and

selective integrin alpha IIb beta 3 antagonists. 1994 Jan 7;37(1):1–8. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/7507165 PMID: 7507165

30. Ravikumar M, Modery CL, Wong TL, Dzuricky M, Sen Gupta A. Mimicking

Adhesive Functionalities of Blood Platelets using Ligand-Decorated Liposomes.

American Chemical Society; 2012 Jun 20;23(6):1266–1275. Available from:

http://pubs.acs.org/doi/10.1021/bc300086d

31. Ravikumar M, Modery CL, Wong TL, Sen Gupta A. Peptide-Decorated

Liposomes Promote Arrest and Aggregation of Activated Platelets under Flow on

Vascular Injury Relevant Protein Surfaces in Vitro. American Chemical Society;

2012 May 14;13(5):1495–1502. Available from:

186 http://pubs.acs.org/doi/10.1021/bm300192t

32. Modery-Pawlowski CL, Tian LL, Ravikumar M, Wong TL, Gupta A Sen. In vitro

and in vivo hemostatic capabilities of a functionally integrated platelet-mimetic

liposomal nanoconstruct. Elsevier Ltd; 2013;34(12):3031–3041. Available from:

http://dx.doi.org/10.1016/j.biomaterials.2012.12.045 PMID: 23357371

33. AlMomani T, Udaykumar HS, Marshall JS, Chandran KB. Micro-scale Dynamic

Simulation of Erythrocyte–Platelet Interaction in Blood Flow. Springer US; 2008

Jun 11;36(6):905–920. Available from: http://link.springer.com/10.1007/s10439-

008-9478-z

34. Tokarev AA, Butylin AA, Ermakova EA, Shnol EE, Panasenko GP,

Ataullakhanov FI. Finite Platelet Size Could Be Responsible for Platelet

Margination Effect. Cell Press; 2011 Oct 19;101(8):1835–1843. Available from:

https://www.sciencedirect.com/science/article/pii/S0006349511010046

35. Anselmo AC, Modery-Pawlowski CL, Menegatti S, Kumar S, Vogus DR, Tian

LL, Chen M, Squires TM, Sen Gupta A, Mitragotri S. Platelet-like Nanoparticles:

Mimicking Shape, Flexibility, and Surface Biology of Platelets To Target Vascular

Injuries. American Chemical Society; 2014 Nov 25;8(11):11243–11253. Available

from: http://pubs.acs.org/doi/10.1021/nn503732m

36. Sen Gupta A. Role of particle size, shape, and stiffness in design of intravascular

drug delivery systems: insights from computations, experiments, and nature. 2016

Mar;8(2):255–270. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/26306941 PMID: 26306941

37. Sercombe L, Veerati T, Moheimani F, Wu SY, Sood AK, Hua S. Advances and

187 Challenges of Liposome Assisted Drug Delivery. Frontiers; 2015 Dec 1;6:286.

Available from: http://journal.frontiersin.org/article/10.3389/fphar.2015.00286

38. Szebeni J, Baranyi L, Savay S, Bodo M, Morse DS, Basta M, Stahl GL, Bünger R,

Alving CR. Liposome-induced pulmonary hypertension: properties and

mechanism of a complement-mediated pseudoallergic reaction. American

Physiological SocietyBethesda, MD ; 2000 Sep;279(3):H1319–H1328. Available

from: http://www.physiology.org/doi/10.1152/ajpheart.2000.279.3.H1319

39. Szebeni J. Hemocompatibility testing for nanomedicines and biologicals:

predictive assays for complement mediated infusion reactions. 2012;4(1):33–53.

Available from: http://www.degruyter.com/view/j/ejnm.2012.4.issue-1/ejnm-2012-

0002/ejnm-2012-0002.xml

188 Chapter 5: Intravenous administration of synthetic platelets (SynthoPlate) in a mouse

liver injury model of uncontrolled hemorrhage improves hemostasis

Content based on: Dyer MR, Hickman DA, Luc N, Haldeman S, Loughran P, Pawlowski CL, Sen Gupta A, Neal MD. Intravenous administration of synthetic platelets (SynthoPlateTM) in a mouse liver injury model of uncontrolled hemorrhage improves hemostasis. J. Trauma Acute Care Surg. 2018; (March). doi: 10.1097/TA.0000000000001893. PMID: 29538234

5.1 Introduction

Traumatic hemorrhage continues to be a leading cause of preventable death in civilian and battlefield scenarios. Given the challenges previously described associated with platelet transfusions, there is significant clinical interest in extending the shelf-life and hemostatic viability of platelets via cold storage and lyophilization1,2, as well as in developing

transfusable synthetic platelet substitutes that allow long-term storage and platelet-inspired hemostatic action3–5. To this end, we have developed a synthetic platelet nanotechnology,

namely SynthoPlate, that incorporates the proadhesive and proaggregatory properties of native platelets by virtue of decorating a combination of platelet function-mimicking peptides (von Willebrand Factor-binding peptides [VBP], collagen binding peptides or collagen binding peptide [CBP] and active platelet GPIIb-IIIa–binding fibrinogen-mimetic peptides [FMP]) on a biocompatible lipid vesicle platform6. SynthoPlate is shelf-stable, portable, and easily administered, addressing many of the challenges and limitations of human platelet usage in remote and prehospital settings. Detailed chemistry of these peptide decorations and functions, in vitro characterization of the resultant SynthoPlate technology and its promising hemostatic effect in correcting tail bleeding time in normal

as well as thrombocytopenic mice, have been reported previously7,8. Building on these

189 findings, we hypothesized that SynthoPlate treatment would result in decreased blood loss

in acute, hemorrhagic shock. To test this, we used a validated, reproducible, and high- throughput model of murine liver injury that results in uncontrolled intraperitoneal

hemorrhage and severe shock9.

5.2 Materials and Methods

5.2.1 Animals

C57BL/6 WT mice (8–12 weeks, male) were purchased from the Jackson Laboratories

(Bar Harbor, ME). Mice were housed in accordance with University of Pittsburgh

(Pittsburgh, PA) and National Institutes of Health (Bethesda, MD) animal care guidelines

in specific pathogen-free conditions with 12-hour light-dark cycles and free access to

standard feed and water. All animal experiments were approved conducted in accordance

with the guidelines set forth by the Animal Research and Care Committee at the University

of Pittsburgh.

5.2.2 SynthoPlate Manufacture

SynthoPlate is made of multiple lipid, lipid-peptide, and cholesterol components, as has

been described previously7,8. 1,2-Distearoyl-sn-glycero-3-phosphocholine, 1,2-Distearoyl-

sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-1000] (DSPE-

mPEG1000), 1,2-Distearoyl-sn-glycero-3-phosphoethanolamine-N-

[maleimide(polyethylene glycol)-2000] (DSPE-PEG2000-Mal), and 1,2-distearoyl-sn-

glycero-3-phosphoethanolamine-N-[azido(polyethylene glycol)-2000] (DSPE-PEG2000-

Azide) were purchased from Avanti Polar Lipids (Alabaster, AL). Rhodamine B-

dihexadecanoyl-sn-glycero-3-phosphoethanolamine (DHPE-RhB) was purchased from

190 Invitrogen (Carlsbad, CA). The peptides VBP (sequence CTRYLRIHPQSWVHQI), CBP

(sequence C[GPO]7) and FMP (sequence cyclo-{Pra}CNPRGD{Tyr(OEt)}RC) were custom-synthesized by GeScript (Piscataway, NJ). vWF-binding peptide and CBP were

10 conjugated to DSPE-PEG2000-Mal using thiol-maleimide coupling and FMP was

conjugated to DSPE-PEG2000-azide via copper-catalyzed alkyne-azide cycloaddition

“click” chemistry (CuCAAC)11. The conjugates were purified by dialysis and characterized

by MALDI-TOF mass spectrometry. Cholesterol was purchased from Sigma-Aldrich

(Saint Louis, MO). The lipid vesicle fabrication technique of “film rehydration and

extrusion” was used to manufacture SynthoPlate as described previously2,4.

Heteromultivalently decorated SynthoPlate vesicles were produced and dynamic light

scattering and electron microscopy characterization indicated fresh-made vesicles were approximately 200 nm in diameter as we have previously reported7.

5.2.3 Liver Laceration Model of Uncontrolled Hemorrhage and SynthoPlate Evaluation

We developed and validated a murine model of uncontrolled hemorrhage using a liver

laceration9. A schematic and timeline of the experimental design is presented in Figure

5-1. Briefly, mice were anesthetized with sodium pentobarbital (70 mg/kg) via

intraperitoneal injection, and real-time hemodynamic monitoring is achieved through femoral arterial cannulation. A midline laparotomy is performed, preweighed absorption

191 triangles are inserted away from the liver, and 75% of the left middle lobe of the liver is

Figure 5-1 Experimental timeline, liver laceration model, and SynthoPlate mechanism. (A) Experimental design and timeline for pretreatment studies. (B) Experimental Design and timeline for rescue studies. (C) The liver laceration procedure is performed through a midline laparotomy and 75% of the left middle lobe is lacerated. This results in uncontrolled intraperitoneal hemorrhage and associated hemorrhagic shock. SynthoPlate particles, a novel synthetic platelet nanotechnology, mimic the adhesion, and aggregatory properties of native platelets by decorating lipid vesicles with platelet functioning-mimicking peptides including include a vWF-binding peptide (VBP), a CBP and an active platelet GPIIb-IIIa–binding FMP. These peptides are those able to bind to sites of hemorrhage and other activated platelets to assist and amplify hemostasis.

lacerated. Mice were treated via tail-vein injection with SynthoPlate particles (30 mg/kg), control particles (unmodified particles, 30 mg/kg), or normal saline (0.9%) either 30 minutes before (pretreatment) to or 1 minute after (rescue) performing the liver laceration.

Animal surgeons were blinded to treatment groups. For rescue strategy experiments, the tail vein was cannulated with a mouse tail vein catheter (SAI Infusion Technologies), and the treatment was administered through the catheter. The dose for SynthoPlate infusion

192 was determined from prior work in tail transection bleeding models, with final infusion

volumes of 250 μL to 300 μL7. When administered in the rescue setting, the SynthoPlate infusion was followed by a flush of normal saline, equal volume, to ensure particles were infused into the circulation. To determine the effect of SynthoPlate on acute blood loss mitigation, mice were sacrificed 20 minutes following liver laceration. Blood loss was calculated by assessing the difference in the weight of absorption triangles pre- and post-

liver laceration. The weight of the resected liver was determined immediately following

sacrifice. Hemodynamic data were collected using DMSI-400 (Micro-Med) software.

5.2.4 In Vivo Biodistribution of SynthoPlate Particles

In consideration of a theoretical risk of systemic off-target sequestration of SynthoPlate resulting in microvascular complications, we sought to evaluate the biodistribution and determine clearance of SynthoPlate following hemorrhagic shock in mice subjected to liver laceration as described above. Tissue homogenates were analyzed using ultra high- performance liquid chromatography (UPLC). For this, organs (brain, heart, lung, liver, injured liver, kidney, and spleen) were harvested at the time of sacrifice, snap-frozen in liquid nitrogen, dried in a lyophilizer, and dry weight was recorded. The organs were homogenized at 4,000 rpm for two cycles of 25 seconds using a BeadBug Microtube

Homogenizer (Benchmark Scientific, Edison, NJ) with 3.0-mm high-impact zirconium beads. Samples were shaken overnight at 750 rpm at 37 °C in a 1:1 solution of methanol/chloroform to extract the RhB-conjugated lipids. These samples were centrifuged at 12,000g for 20 minutes, and the supernatant, containing RhB-labeled lipids from SynthoPlate, was collected. The RhB-labeled lipids were resolved using Waters

193 Acquity UPLC system (column: Waters BEH C8 1.7 μm) and analyzed with a fluorescence

detector (Waters Corporation, Milford, MA) using excitation wavelength 560 nm and emission wavelength of 580 nm. The uptake of SynthoPlate in the organs was determined using an RhB fluorescence-based calibration curve for SynthoPlate and reported as the

percent of injected dose as per calibration curve.

5.2.5 Statistics

The primary objective for this study was to determine the efficacy of SynthoPlate transfusion to reduce blood loss in acute, hemorrhagic shock compared with control

particles and normal saline (0.9%). Secondary endpoints were hemodynamic analysis and

biodistribution. Based on pilot data, to detect a decrease in blood loss of 20% resulting in

a treatment effect size of 1.5, at 80% power, and a significance level of 0.05, we determined

that eight mice per group would be needed for the primary endpoint analysis. All data are

presented as mean ± SD for n ≥ 3 unless stated otherwise in the figure legends. Statistical

significance was determined with the two-tailed Student's t test or one-way analysis of variance with Bonferroni post hoc test using Graph Pad Prism software (GraphPad). A p value of less than 0.05 was considered significant.

5.3 Results

5.3.1 SynthoPlate Transfusion as a Pretreatment Results in Decreased Blood Loss and

Improved Hemodynamics During Acute, Severe Hemorrhagic Shock

We evaluated the capability of SynthoPlate to improve hemostasis, as determined by blood

loss, in a severe model of trauma and hemorrhagic shock, using a validated murine model

194 of uncontrolled hemorrhage induced by liver laceration (Figure 5-1).19 As demonstrated in

Figure 5-2A, mice pretreated 30 minutes before liver laceration with SynthoPlate showed a significant decrease in acute blood loss following liver laceration, compared with mice treated with control particles or normal saline (0.86 ± 0.16 g control particle [CP] vs. 0.84

± 0.13 g normal saline [NS] vs. 0.68 ± 0.09 g SynthoPlate). To ensure the differences in blood loss between treatment groups was due to SynthoPlate

Figure 5-2 Analysis of SynthoPlate pretreatment on blood loss and hemodynamics during hemorrhagic shock. Mice were pretreated 30 minutes before liver laceration with either CP, SynthoPlate, or 0.9% NS. Pretreatment with SynthoPlate results in a significant reduction in blood loss acutely (20 minutes) after liver laceration (n = 12 CP, n = 14 SP and NS) (A). As a measure of control for degree of hemorrhage between the different treatment groups the weight of the resected liver was weight to ensure the same degree of liver injury, as demonstrated in (B) no significant difference between the liver injury between any treatment group was identified. All mice underwent arterial cannulation and real-time hemodynamic monitoring, shown in (C) mice that received SynthoPlate took significantly longer to develop hypotension (defined as <40 mm Hg)

195 compared with mice treated with control particles or saline (n = 8 per group). Representative tracings of the MAP between mice treated with SynthoPlate particles compared to control particles demonstrating the improved hemodynamics and delay to develop hypotension (D). No difference in time to hypotension was noted between SynthoPlate or control particle treated mice (n = 11 CP, n = 12 SP, n = 11 NS) (F). *p < 0.005 **p < 0.01 ***p < 0.05. MAP, mean arterial pressure.

transfusion and not secondary to differences in levels of severity of liver injury, the weight of the segment of liver resected during the laceration was determined. There were no significant differences between treatment groups with regard to the weight of the resected liver (Figure 5-2B, 0.32 ± 0.05 g CP vs. 0.33 ± 0.03 g NS vs. 0.30 ± 0.04 g SynthoPlate).

All animals had femoral arterial lines placed for real-time hemodynamic monitoring, and as a secondary endpoint, we compared the time elapsed following liver laceration to development of hypotension, defined as a mean arterial pressure less than 40 mm Hg. As shown in Figure 5-2C-D, the time to develop hypotension is significantly prolonged in mice pretreated with SynthoPlate, and hypotension was prevented in some mice pretreated with SynthoPlate compared with mice treated with control particles or normal saline (168.3

± 106.6 seconds CP vs. 137 ± 58 seconds NS vs. 546.7 ± 329.8 seconds SynthoPlate).

5.3.2 Systemic Biodistribution SynthoPlate Particles Following Liver Laceration

We sought to evaluate the biodistribution-organ distribution, particle clearance, and off- target sequestration of SynthoPlate particles in the setting of our hemorrhagic shock model.

Ultra high-performance liquid chromatography was performed to determine the percent of injected SynthoPlate particle extracted from an organ harvested following liver laceration.

We found that in the acute hemorrhagic shock setting, SynthoPlate particles are distributed most frequently in the lung, liver, and in the injured liver site (Figure 5-3), although the overall sequestration of these particles in uninjured sites was low. As inferred from Figure

196 5-3, immediately following hemorrhage, only approximately 25% of the injected

SynthoPlate particles are cleared into the various organs, suggesting that most of the transfused SynthoPlate particles remain in circulation and available for continued assistance in hemorrhage control over time, if needed.

Figure 5-3 In vivo biodistribution of SynthoPlate particles. The systemic biodistribution of SynthoPlate particles following liver laceration and hemorrhagic shock was assessed using two different methods. SynthoPlate particle concentration in each organ was determined with UPLC and then expressed as a percent of the overall dose that was injected before liver laceration. As shown, the liver is most common organ the SynthoPlate particles are found, which is consistent with prior data the demonstrated in uninjured mice the liver and spleen were the major organs for SynthoPlate clearance (n = 6).

197 5.3.3 SynthoPlate Particles Reduce Blood Loss and Improve Hemodynamics in Mice in

Hemorrhagic Shock

After determining efficacy in reducing blood loss and improving hemodynamics in a pretreatment setting, we next sought to evaluate efficacy in a rescue model where

SynthoPlate transfusion was performed after liver laceration to more accurately model traumatic hemorrhage (Figure 5-1). Mice transfused with SynthoPlate following liver laceration experienced significantly less blood loss compared to either CP- or NS-treated mice (Figure 5-4A, 0.89 ± 0.17 g CP vs. 0.92 ± 0.19 g NS vs. 0.69 ± 0.18 g SynthoPlate).

The weights of the resected livers were determined to control for model severity.

Consistent with the findings in the pretreatment experiments, there were no significant differences in the weight of the resected liver between the treatment groups (Figure 5-4B,

0.33 ± 0.05 g CP vs. 0.31 ± 0.05 g NS vs. 0.30 ± 0.06 g SynthoPlate). In contrast to the pretreatment experiments, SynthoPlate transfusion following liver laceration did not result in any hemodynamic differences between the SynthoPlate group compared to control particles or normal saline (Figure 5-4C, 543 ± 208 seconds CP vs. 486 ± 127 seconds NS vs. 510 ± 202 SynthoPlate).

5.4 Discussion

We present here the potential applicability of an intravenous synthetic platelet nanotechnology (SynthoPlate) as a primary resuscitative agent in mitigating traumatic hemorrhage in a murine model of uncontrolled hemorrhage. Our data demonstrate the ability of SynthoPlate particles to assist in primary hemorrhage control in a setting of severe bleeding, in a pretreatment setting as well as a rescue strategy. These data are

198

Figure 5-4 Analysis of SynthoPlate transfusion on blood loss and hemodynamics as a rescue strategy in acute, hemorrhagic shock. Mice underwent liver laceration and were transfused either CP, SynthoPlate, or 0.9% NS following the liver laceration. Administration of SynthoPlate following liver laceration led a significant reduction in blood loss compared to control particles or saline (n = 11 CP, n = 12 SP, n = 11 NS) (A). As a control for the degree of hemorrhage, the weight of the resected liver was determined and as demonstrated in (B) there were no significant differences between the treatment groups. Finally, hemodynamic monitoring did not reveal any significant differences in the hemodynamic profiles between the treatment groups, as assessed by time to develop hypotension (C). *p < 0.05.

particularly striking when it is considered that the SynthoPlate particles are administered as a single-dose injection without other sources of fluid resuscitation during the hemorrhage period.

SynthoPlate particles can improve hemostasis by direct mimicry and amplification of platelet's primary hemostatic mechanisms of injury site-specific adhesion and aggregation

199 resulting in “in effect” enhancement of available platelet procoagulant surface to also

amplify secondary hemostasis (Figure 5-1). While SynthoPlate has been evaluated

previously, both via in vitro characterization and in vivo via tail vein bleeding assays, we

sought to evaluate the hemostatic capabilities in a more severe model of bleeding with

resultant shock physiology. We initially chose to evaluate the effects of SynthoPlate

transfusion in a pretreatment setting in the severe hemorrhage model in mice because our previous studies had shown hemostatic benefit of SynthoPlate pretreatment in murine models of subacute bleeding (tail transection), and we hypothesized that this benefit is conserved when bleeding became more acute (hemorrhagic trauma)7,8. After seeing

promising results of hemostatic and hemodynamic benefit in pretreatment evaluation, we

moved to evaluation in a model more reflective of rescuing from traumatic hemorrhage,

that is, transfusion after the liver injury was performed. Again, in this setting, we noted

improved blood loss in mice transfused with SynthoPlate.

SynthoPlate transfusion in the pretreatment setting resulted in improved hemodynamics,

as measure by the time elapsed from liver laceration to a drop to a mean arterial pressure

below 40 mm Hg. In contrast to pretreatment transfusion, the hemodynamic effects of

SynthoPlate transfusion following liver laceration did not result in a different

hemodynamic profile compared to either control particles or saline. In fact, the

hemodynamic profiles between the two experimental strategies are quite different as all

treatment groups in rescue experiments took longer to develop hypotension. We rationalize

this difference between experiments may be due to delivery of the particles. To ensure the

particles, either SynthoPlate or control, are delivered and completely removed from the

200 catheter, a saline flush is performed immediately after injecting the particle dose. This is

in effect a secondary fluid bolus delivered to the mouse which may alter the time to

hypotension as compared to the pretreatment strategy.

We also emphasize here that several “synthetic platelet” designs have been developed and

reported in the past few decades, but these designs have been very rarely evaluated in an

in vivo model of traumatic uncontrolled hemorrhage. The majority of past studies with

synthetic platelet designs have involved particles that only bear the aggregatory function of platelets (e.g., liposomal, albumin or polymeric particles coated with fibrinogen or fibrinogen-relevant Arginylglycylaspartic acid peptides), and these have been evaluated

mostly as pretreatment in platelet dysfunction (e.g., thrombocytopenia) models using tail-

bleeding and ear-punch injuries in mice, rats, or rabbits12,13. Only one specific design made from polymeric nanoparticles decorated with a generic RGD peptide has been studied in a more severe injury (femoral artery bleed) as a pretreatment (administered 5 minutes before injury) and a rescue treatment in blast trauma (administered 5 minutes after blast injury) in mice, showing promising results14,15. Therefore, our current studies with SynthoPlate

administered intravenously as pretreatment (30 minutes before injury), as well as a rescue

treatment in a mouse model of uncontrolled liver hemorrhage, are essentially one of the

very few to show the feasibility and potential of such a technology in mitigating severe

hemorrhage in both prophylactic (e.g., in elective surgery) and emergency (e.g., trauma) framework.

201 Previous work has demonstrated that intravenous injection of SynthoPlate particles in

uninjured mice does not lead to systemic pro-thrombotic risks, as evidenced by a lack of

generation of pro-thrombin fragments or D-dimer fragments in circulation8. Also, 2 hours following injection in uninjured mice, clearance of SynthoPlate particles from circulation was found to be approximately 15%, with the liver and spleen being primarily responsible for clearance8. Analysis of systemic biodistribution of SynthoPlate particles during

hemorrhage showed the liver as the major clearing site for circulating SynthoPlate particles

consistent with the prior studies in uninjured mice. Finally, we quantified the clearance of

SynthoPlate from circulation in the setting of acute hemorrhage and found that only

approximately 25% of injected particle dose was cleared at the time of sacrifice. This

suggests that most the SynthoPlate transfused remains in circulation and can provide

hemostatic assistance in areas with ongoing bleeding.

While our findings certainly present evidence for the beneficial effects of SynthoPlate

transfusion on reducing blood loss in severe hemorrhagic shock, our work is not without

limitations. First, our study was powered to detect differences in blood loss and therefore

other important endpoints in hemorrhagic shock such as survival could not be adequately

addressed in this current study. Second, our comparison groups were control particles and

normal saline. More evidence is suggesting early administration of blood products,

including packed red blood cells, fresh frozen plasma, and platelets have benefits over

crystalloid administration. Future studies that compare SynthoPlate transfusion against

platelet transfusion and in conjunction with platelet transfusion are needed and a planned

area of work.

202 Despite these limitations, we present evidence for the effectiveness in two different strategies, pretreatment and rescue strategies. Pretreatment effectiveness presents the potential for use of SynthoPlate in the setting of thrombocytopenia and platelet dysfunction

(e.g., uremia) to reduce blood loss during planned procedures. Rescue therapy effectives highlights an exciting potential for the use of SynthoPlate as part of a resuscitation protocol, in conjunction with platelets or as a substitute, in acute bleeding.

In conclusion, these data show great promise for the use of SynthoPlate as a viable intravenous resuscitative product for resuscitative mitigation of hemorrhagic shock by decreasing blood loss and stabilizing blood pressure.

5.5 Authorship

M.R.D., D.H., N.L., S.H., C.P., A.S.G., and M.D.N. designed the research study. D.H.,

N.L., C.P., and A.S.G. fabricated the SynthoPlate particles. M.R.D., D.H., and S.H. conducted the experiments. M.R.D., D.H., S.H., and P.L. acquired data. M.R.D., D.H.,

P.L., A.S.G., and M.D.N. analyzed the data. M.R.D., A.S.G., and M.D.N. wrote the article.

5.6 Acknowledgement

We would like to thank Ms. Lauryn Kohut for her expertise and assistance with development of the liver laceration model.

5.7 Disclosure

M.D.N. has the following financial relationships to disclose: Consultant, External

Scientific Advisor for Anticoagulation Science for Janssen Pharmaceuticals (Johnson &

203 Johnson), Trauma Advisory Board, CSL Behring, and US Patent 9,072,760 TLR4 inhibitors for the treatment of human infectious and inflammatory disorders (issued to

Neal, Wipf, Hackam, Sodhi). ‘SynthoPlate’ technology is associated with the US Patent

9107845 and the SynthoPlate trade mark is registered currently with USPTO (ASG, CP).

This work is supported by U.S. National Institutes of Health grants 1 R35 GM119526-01

and UM1HL120877-01, support from the Vascular Medicine Institute at the University of

Pittsburgh, the Hemophilia Center of Western Pennsylvania, and the Institute for

Transfusion Medicine, as well as the American Association for the Surgery for Trauma

Research award (all to M.D.N.), U.S. National Institutes of Health grants 5R01 HL121212

(to A.S.G.) and U.S. National Institutes of Health grant 1S10OD019973-01 (Dr. Simon C.

Watkins).

5.8 References

1. Spinella PC, Cap AP. Prehospital hemostatic resuscitation to achieve zero

preventable deaths after traumatic injury. 2017 Nov;24(6):529–535. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/28832355 PMID: 28832355

2. Etchill EW, Myers SP, Raval JS, Hassoune A, SenGupta A, Neal MD. Platelet

Transfusion in Critical Care and Surgery. 2017 May;47(5):537–549. Available

from: http://insights.ovid.com/crossref?an=00024382-201705000-00002

3. Modery-Pawlowski CL, Tian LL, Pan V, McCrae KR, Mitragotri S, Sen Gupta A.

Approaches to synthetic platelet analogs. Elsevier Ltd; 2013;34(2):526–541.

Available from: http://linkinghub.elsevier.com/retrieve/pii/S0142961212010976

4. Sim X, Poncz M, Gadue P, French D. Understanding platelet generation from

204 megakaryocytes: implications for in vitro-derived platelets. 2016;127(10):1227–

1233. Available from:

http://www.bloodjournal.org/content/early/2016/01/19/blood-2015-08-

607929.abstract

5. Lambert MP, Sullivan SK, Fuentes R, French DL, Poncz M. Challenges and

promises for the development of donor-independent platelet transfusions.

American Society of Hematology; 2013 Apr 25;121(17):3319–24. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/23321255 PMID: 23321255

6. Sen Gupta A, Ravikumar M, Modery‐Pawlowski C. Synthetic platelets [Internet].

US; US9107845B2, 2012. Available from:

https://patents.google.com/patent/US9107845?oq=9107845%2C+2015

7. Modery-Pawlowski CL, Tian LL, Ravikumar M, Wong TL, Gupta A Sen. In vitro

and in vivo hemostatic capabilities of a functionally integrated platelet-mimetic

liposomal nanoconstruct. Elsevier Ltd; 2013;34(12):3031–3041. Available from:

http://dx.doi.org/10.1016/j.biomaterials.2012.12.045 PMID: 23357371

8. Shukla M, Sekhon UDS, Betapudi V, Li W, Hickman DA, Pawlowski CL, Dyer

MR, Neal MD, McCrae KR, Sen Gupta A. In vitro characterization of

SynthoPlateTM (synthetic platelet) technology and its in vivo evaluation in severely

thrombocytopenic mice. 2017;15(2):375–387. PMID: 27925685

9. Dyer M, Haldeman S, Gutierrez A, Kohut L, Sen Gupta A, Neal MD.

Uncontrolled Hemorrhagic Shock Modeled via Liver Laceration in Mice with Real

Time Hemodynamic Monitoring. MyJoVE Corporation; 2017;(123). Available

from: http://www.ncbi.nlm.nih.gov/pubmed/28570538 PMID: 28570538

205 10. Elliott JT, Prestwich GD. Maleimide-Functionalized Lipids that Anchor

Polypeptides to Lipid Bilayers and Membranes. 2000 Nov;11(6):832–841.

Available from: http://pubs.acs.org/doi/abs/10.1021/bc000022a

11. Presolski SI, Hong VP, Finn MG. Copper-Catalyzed Azide-Alkyne Click

Chemistry for Bioconjugation. Hoboken, NJ, USA: John Wiley & Sons, Inc.;

2011. Available from: http://doi.wiley.com/10.1002/9780470559277.ch110148

12. Kutcher ME, Kornblith LZ, Narayan R, Curd V, Daley AT, Redick BJ, Nelson

MF, Fiebig EW, Cohen MJ. A Paradigm Shift in Trauma Resuscitation. 2013 Sep

1;148(9):834. Available from: http://www.ncbi.nlm.nih.gov/pubmed/23864019

PMID: 23864019

13. Hickman DA, Pawlowski CL, Sekhon UDS, Marks J, Gupta A Sen. Biomaterials

and Advanced Technologies for Hemostatic Management of Bleeding.

2017;1700859:1–40. PMID: 29164804

14. Bertram JP, Williams C a, Robinson R, Segal SS, Flynn NT, Lavik EB.

Intravenous hemostat: nanotechnology to halt bleeding. 2009;1(11):11ra22. PMID:

20371456

15. Lashof-Sullivan MM, Shoffstall E, Atkins KT, Keane N, Bir C, VandeVord P,

Lavik EB. Intravenously administered nanoparticles increase survival following

blast trauma. National Academy of Sciences; 2014 Jul 15;111(28):10293–10298.

Available from: http://www.pnas.org/cgi/doi/10.1073/pnas.1406979111

206 Chapter 6: Intravenous synthetic platelet (SynthoPlateTM) nanoconstructs reduce

bleeding and improve ‘golden hour’ survival in a porcine model of traumatic

arterial hemorrhage

Content based on: Hickman DA, Pawlowski CL, Shevitz A, Luc NF, Kim A, Girish A, Marks J, Ganjoo S, Huang S, Niedoba E, Sekhon UDS, Sun M, Dyer M, Neal MD, Kashyap V, Sen Gupta A. Intravenous synthetic platelet (SynthoPlate) nanoconstructs reduce bleeding and improve “golden hour” survival in a porcine model of traumatic arterial hemorrhage. Sci Rep. 2018;(February):1–14. doi: 10.1038/s41598-018-21384-z. PMCID: PMC5814434

6.1 Introduction

In remote civilian locations and austere battlefield conditions, uncontrolled traumatic

hemorrhage remains one of the leading causes of mortality1–7. In such scenarios,

transfusion of whole blood or blood components (RBCs, plasma and platelets) can

significantly improve survival8–10. Extensive trauma resuscitation studies have indicated substantial benefits of early platelet transfusion to treat hemorrhagic shock and enhance survival possibilities8–13, however, the limited availability and portability, need for blood

type matching, special storage requirements, high risks of bacterial contamination at room

temperature and very short shelf-life (3-5 days at room temperature) of platelets, present severe logistical challenges for their applicability in pre-hospital scenarios14–24. These

issues have led to robust research efforts in improving the storage stability, portability and

availability of platelets, e.g. via lyophilization and cold-storage, as well as, developing

pathogen reduction technologies to reduce contamination25–32. However, these approaches

have only marginally improved the pre-hospital applicability of platelets, and the research

continues to find alternative platelet-based options, e.g. Isolated Platelet Membrane (IPM

207 CyplexTM) and Thrombosome® 33–35. These technologies involve isolation, purification,

sterilization, temperature-stabilization and lyophilization of outdated platelets or platelet-

derived membrane to utilize their residual hemostatic bioactivity. However, such

processing can lead to significant ‘loss-in-function’ of hemostatically relevant glycoprotein on the platelet membrane surface36,37, which can potentially lead to batch-to-batch

functional inconsistencies. Due to such challenges with natural platelet-based products, a

parallel area of research has focused on the development of synthetic platelet surrogates38–

40. The design requirements for such platelet surrogates are: (i) large scale in vitro

manufacturability, (ii) sterilizability without compromising biofunctional properties, (iii)

long-term storage as suspension or lyophilized powder, (iv) easy portability, (v)

intravenously administrable ‘on demand’ in pre-hospital scenarios, (vi) post-administration circulation safety without systemic risks, (vii) ability to mimic and amplify endogenous hemostatic mechanisms selectively at the bleeding site, and (viii) biodegradation and safe elimination from the body. Also, such synthetic platelet surrogates can potentially avoid the need for blood type matching and can be used more universally, since they would not bear blood antigens.

As comprehensively reviewed in chapters 2 and 3 and in several review articles39–41,

several ‘synthetic platelet’ designs have been reported in the past that primarily mimic

platelet’s aggregatory capability only via surface- decoration of polymeric, lipidic or albumin particles with fibrinogen (Fg) or Fg-relevant peptides. While these designs have shown some promise as intravenous hemostats, the majority of them have only been reportedly evaluated in small animal bleeding models and only in a pre-treatment

208 framework, i.e. particle administration before bleeding injury. No synthetic platelet design

Figure 6-1 Schematic representation of SynthoPlateTM design and mechanism, showing nanoconstructs heteromutlivalently decorated with VBP, CBP and FMP motifs to render platelet-inspired interactions with vWF, collagen and active platelet integrin GPIIb-IIIa respectively, and thus amplify platelet-mediated primary hemostatic mechanisms at the injury site. Illustration by Erika Woodrum.

has yet been reported regarding evaluation in a large animal model, that is also in an

emergency/rescue framework (i.e. particle administration after bleeding injury) in

traumatic bleeding. We have developed a synthetic platelet surrogate, SynthoPlateTM, by leveraging a clinically relevant biocompatible liposomal platform that is heteromultivalently surface-decorated with three types of peptides: von Willebrand Factor

(vWF)-binding peptide (VBP) and collagen-binding peptide (CBP) to mimic platelet’s bleeding site-selective adhesion mechanisms and active platelet integrin GPIIb-

209

Figure 6-2 Lipid-to-peptide bioconjugation schemes (along with corresponding chemical structures) for synthesizing DSPE-PEG-VBP, DSPE-PEG-CBP and DSPE-PEG-CMP molecules, that were used to manufacture the SynthoPlateTM nanoconstruct.

IIIa-binding fibrinogen-mimetic peptide (FMP) to mimic activated platelet aggregation

mechanism (Figure 6-1)42–44. This integrative biomimetic approach to combine platelets’

adhesion and aggregation functions on a single particle platform is highly unique and can

also be adapted to other particle platforms beyond the SynthoPlateTM’s liposomal

platform45. The chemical components and manufacturing details for SynthoPlateTM are provided in the Methods section and the relevant chemical synthesis schemes are shown in

Figure 6-2. We have previously reported the detailed biofunctional and mechanistic characterization of SynthoPlateTM nanoconstructs46 and have also demonstrated that the

heteromultivalent integration of platelet-inspired adhesion and aggregation mechanisms

210 on the same particle platform results in higher hemostatic capability compared to designs

bearing ‘adhesion only’ or ‘aggregation only’ functions43,45. We further demonstrated that

prophylactic administration of SynthoPlateTM in normal as well as thrombocytopenic mice

could augment hemostatic capability in a non-trauma bleeding model (bleeding time reduction in murine tail transection)43,45. Building on these studies, here we sought to

investigate the translational promise and hemostatic capability of SynthoPlateTM in

emergency management of traumatic exsanguinating hemorrhage in a pilot scale study in

large animal (pig) model. For this, we first characterized the effect of long-term storage (in

saline) on SynthoPlateTM stability, as well as, the effect of sterilization on SynthoPlateTM

stability and biofunction in vitro. Next, for in vivo studies, we adapted a femoral artery

uncontrolled hemorrhage model in pigs, that has been utilized for evaluation of hemostatic

dressings47,48. In traumatic exsanguinating hemorrhage, significant number of patients succumb within the first 1-2 hours, and hemostatic intervention within this window (known

as the ‘golden hour’) can significantly improve survival10,49,50. Our model results in the loss

of ~1 liter of blood within 15-20 minutes, resulting in 75% mortality within first 45 min.

Therefore, we utilized this model to test the SynthoPlateTM technology I.V.-administered

immediately following injury, and the animals were observed for up to 2 hours to monitor

blood loss, blood pressure, heart rate and survival. The systemic safety and biodistribution

of SynthoPlateTM was also characterized. Administrations of saline or of unmodified

211 (control) particles were used as comparison groups.

Figure 6-3 [A] SynthoPlateTM diameter (size distribution) analysis over a 6-month period demonstrated that the particles retained their size within 20% of their starting diameter when stored in a 0.9% NaCl solution at 25oC, indicating long term stability; [B] After sterilization with filtration or E-beam, SynthoPlateTM showed minimal alteration in size (diameter) compared to fresh-made (unsterilized) samples, indicating that the sterilization did not affect particle stability; [C] Representative fluorescent images and quantitative analysis of surface-averaged fluorescence intensity of adhered SynthoPlateTM showing that sterilized SynthoPlateTM (red) maintained its ability to adhere to ‘collagen + vWF’-coated surface at levels similar to fresh-made (unsterilized) SynthoPlateTM, indicating that sterilization did not affect the biofunctional properties of VBP and CBP; [D] Aggregometry analysis showed that neither fresh-made (unsterilized) or sterilized SynthoPlateTM induced spontaneous aggregation of platelets without agonist (w/o ADP), but both unsterilized and sterilized SynthoPlateTM markedly increased platelet aggregation in the presence of ADP, indicating that sterilization did not affect the biofunctional property of FMP on and also SynthoPlateTM did not have any thrombotic risk towards resting platelets.

6.2 Results

6.2.1 SynthoPlateTM stability in storage

Liposomal nanoconstruct instability is reflected by large alterations in size (diameter), stemming from particle disassembly, aggregation and clustering. Dynamic Light Scattering

(DLS) analysis was used to monitor SynthoPlateTM stability in saline suspension over a 6- month period. Fresh-made liposome batches show a size (diameter) variability of +/- 15% in laboratory scale manufacture. As shown in Figure 6-3A, our analysis showed that when

212 stored as a suspension in 0.9% NaCl at room

temperature, SynthoPlateTM particles showed

additional size variability over time of about +/- 5%,

thus maintaining their size within an overall

variability of +/- 20% of their starting diameter for

up to 6 months. Representative DLS data at Month 1,

Month 3 and Month 6 are shown in Figure 6-4. Also,

by visual inspection, there were no signs of particle

aggregation or settlement out of solution. This

suggests that SynthoPlateTM is amenable to long- Figure 6-4 Representative Dynamic Light term storage under these conditions and indicates Scattering (DLS) data of SynthoPlateTM their potential for pre-hospital availability as a small size distribution characterization over a 6- volume suspension without compromising stability. month period; data is shown for 1 month,

3 month and 6 month time points,

demonstrating that there is minimal 6.2.2 SynthoPlateTM sterilization and its effects alteration in particle diameter and SynthoPlateTM sterilization studies by filtration (0.2 therefore indicating particle stability in µm filter) as well as E-beam exposure (25 and 40 storage (saline suspension at 25oC). kGy) were carried out with 6 different challenge organisms, namely Staphylococcus aureus, Kocuria rhizophila, Clostridium sporogenes,

Bacillus subtilis spizizenii, Candida albicans and Aspergillus brasiliensis, as described in

the Methods section. For all organisms tested, SynthoPlateTM itself did not have any bacteriostatic or fungistatic activity, and this data is shown in Figure 6-5. SynthoPlateTM

solutions sterilized by either filtration or by E-beam exposure showed no visible growth of

213 microorganisms, suggesting effective sterilization by all methods tested. The endotoxin

Figure 6-5 Bacteriostatic and fungistatic analysis of SynthoPlateTM suspension exposed to E-beam sterilization and challenged with 6 different organisms, namely Staphylococcus aureus, Kocuria rhizophila, Clostridium sporogenes, Bacillus subtilis spizizenii, Candida albicans and Aspergillus brasiliensis, that were allowed to grow in appropriate conditions for up to 5 days. T= test, C= control, N = No growth, P = positive growth and N/A = positive growth established.

bioburden of the concentrated SynthoPlateTM formulation (1 x 105 moles lipid per ml) was

measured to be 3.5-5 EU/ml as per a standard chromogenic limulous ameobocyte lysate

(LAL) assay (see Methods for details). When diluted in sterile saline to the in vivo dosing level of 5 x 104 moles lipid per ml, this equates to a dose of ~0.83 EU/kg, which is

substantially less than the endotoxin safety threshold of ≤ 5 EU/kg for in vivo usage51.

Therefore, our SynthoPlateTM preparation can be considered safe for in vivo usage.

Furthermore, after sterilization by either filtration or E-beam, SynthoPlateTM particles showed statistically insignificant variability in size compared to fresh-made (unsterilized)

SynthoPlateTM (Figure 6-3B), suggesting that the sterilization methods did not affect

particle integrity, size and stability.

To test whether sterilization affects the platelet-inspired adhesive function of CBP and

VBP moieties on SynthoPlateTM, Rhodamine B-labeled (red fluorescent) SynthoPlateTM

214 particles in saline suspension, unsterilized or sterilized, were incubated on ‘vWF +

Figure 6-6 Representative Lumi-Aggregometry raw data from studies on SynthoPlateTM (fresh unsterilized versus sterilized) interaction with platelets (resting or agonist-activated) in platelet-rich-plasma (PRP); results indicate that neither unsterilized nor sterilized SynthoPlateTM has any activating and aggregatory effect on resting platelets (top group of traces w/o ADP), while both unsterilized and sterilized SynthoPlateTM are capable of enhancing the aggregation of activated platelets (bottom group of traces with ADP) above that seen for PRP only. These results also demonstrate that sterilization does not affect the pro-aggregatory function of SynthoPlateTM on activated platelets.

collagen’ coated glass coverslips in a 12-well plate with stirring at 120 rpm, following

which the coverslips were washed and imaged (experimental details provided in Methods

section). Representative images of Rhodamine B (red) fluorescent adhered SynthoPlateTM,

as well as, statistical data of fluorescence intensity analysis from multiple images per

condition (5 images per condition), are shown in Figure 6-3C. As evident from the results,

sterilized SynthoPlateTM particles retained their capability to adhere to ‘collagen + vWF’-

coated surfaces at levels similar to freshly prepared SynthoPlateTM particles, indicating that

sterilization by filtration or E-beam does not affect the platelet-inspired adhesive activity

of VBP and CBP moieties on SynthoPlateTM. To test whether sterilization affects the ability

215 of FMP moieties on SynthoPlateTM to interact with active platelet integrin GPIIb-IIIa for

amplifying platelet aggregation, lumi-aggregometry assays were performed using platelet

rich plasma (PRP) isolated from citrated human whole blood. These assays were performed

with 100% PRP or PRP diluted 50% (v/v) with platelet poor plasma (PPP) supplemented

with fresh made (unsterilized) or sterilized (filtration and E-beam) SynthoPlateTM particles,

with or without addition of platelet agonist ADP (experimental details in Methods section).

In these studies, neither unsterilized nor sterilized SynthoPlateTM particles were found to

induce any spontaneous platelet aggregation in the absence of ADP, but both were able to

markedly enhance platelet aggregation in the presence of ADP, as shown in Figure 6-3D.

Representative raw data from aggregometry studies are shown in Figure 6-6. Therefore,

these results indicate that sterilization by filtration or E-beam does not affect the platelet- inspired aggregatory activity of FMP moieties on SynthoPlateTM. Hence, altogether these

results establish that SynthoPlateTM

can be effectively sterilized without

affecting its platelet-inspired

bioactivity, which is important in

the context of its future translation

through large-scale manufacturing

and sterilization without

compromising its hemostatic

Figure 6-7 MTT-based metabolic activity analysis data of function. human 3T3 fibroblasts in culture incubated with 0.5 nM SynthoPlateTM (dose relevant to in vivo studies) or 1 nM SynthoPlateTM (double of in vivo dose) showed no statistical difference compared to cell metabolic activity in culture media only, indicating that SynthoPlateTM is not cytotoxic at the dose used.

216 6.2.3 SynthoPlateTM cytotoxicity analysis

SynthoPlateTM was incubated with 3T3 human fibroblasts (see Methods for culture

conditions and viability analysis) at doses relevant to the pig dose used in the current

studies (0.5 nM) and double the dose (1 nM). Cell viability was analyzed by standard MTT

based metabolic assay and compared to cell cultures without SynthoPlateTM incubation (i.e. culture media only). As shown in Figure 6-1. the cell metabolic activity was not affected by SynthoPlateTM incubation (i.e. no statistical difference), indicating that there is no

cytotoxicity associated with SynthoPlateTM at the dose (as well as double the dose) used in vivo in the current studies.

6.2.4 In vivo Safety and Biodistribution Studies in Pigs

The systemic safety of SynthoPlateTM in pigs within the ‘golden hour’ (60 min) time-frame, was tested by administering the SynthoPlateTM dose in un-injured pigs intravenously through the jugular vein. Dose was calculated by

5 10 ÷ 50 = 1 10 / , 12 11 where∗ a typical𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 unit𝑝𝑝𝑝𝑝𝑝𝑝 of platelets𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 for 𝑚𝑚human𝑚𝑚 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 transfusion𝑠𝑠𝑠𝑠 𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 𝑣𝑣 𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣contains ≥ 3∗ x 1011𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 platelets. Since𝑚𝑚𝑚𝑚

actual platelets are ~10 times larger (in diameter) than SynthoPlate, we decided to scale the transfusion of SynthoPlate (or control particles) by 1 order of magnitude, to 5 x 1012

particles per dose. Administrations of saline or unmodified (control) particle dose were

used as comparison. Two animals per treatment group (i.e. 6 total) were used for these

studies. The effects of the treatments on vitals, blood chemistry, complement (C3 to C3a)

activation, platelet function (aggregation) and blood clot viscoelastic properties (using

217 ROTEM) were examined for the 60-minute post-administration period (experimental

Figure 6-8 [A] Average vitals over the course of the experiment of uninjured pigs and [B] average blood lab values over the course of the experiment of uninjured pigs upon administration of unmodified particles or SynthoPlateTM, showed that there are no statistically significant differences in values compared to saline administration, indicating that SynthoPlateTM administration has no detrimental physiological effect; [C] Analysis of C3:C3a plasma concentration ratios in blood drawn from pigs administered with unmodified particles or SynthoPlateTM showed no statistically significant alterations of this ratio compared to saline administration groups, indicating that in the dose and administration protocol used, SynthoPlateTM did not have complement activation (and CARPA) risk; [D] Biodistribution analysis from harvested organs indicated a similar distribution profile for unmodified particles and SynthoPlateTM, with a majority of particles cleared through the liver.

details in Methods section). Sixty minutes post treatment administration, pigs were euthanized and major organs (lungs, heart, liver, kidney, spleen) were harvested for histological (using H&E stain) and biodistribution (using UPLC on tissue homogenate) analysis. As shown in Figure 6-8A, following administration of unmodified particles or

SynthoPlateTM in uninjured pigs, none of the animals showed any substantial alteration in their vitals compared to saline. Ex vivo analysis of arterial blood collected from the pigs post treatment administration showed no significant changes in the blood analysis

218 parameters between saline and the two particle treatment groups (Figure 6-8B), indicating

that the SynthoPlateTM dose did not affect physiological parameters of blood in vivo. Also,

as shown in Figure 6-8C, no significant alterations in the C3:C3a plasma concentration

ratios were found in animals administered with unmodified particles or SynthoPlateTM. It

is important to note here that complement ratio alterations of 4-fold or greater are reported

to be a risk indicator for drastic activation of complement system and deleterious reactions

such as complement activation-related pseudoallergy (CARPA)52–54. SynthoPlateTM (as well as control particles) did not show any such effect compared to saline. Therefore, altogether these results suggest that within our experimental dose and animal observation window, administration of SynthoPlateTM did not cause any alarming systemic risks. The

biodistribution analysis of SynthoPlateTM (as well as unmodified particles) in uninjured

pigs is shown in Figure 6-8D, demonstrating that both types of particles have similar organ

distribution profile during the 1-hour circulation period. Within this window, most particles

of either type were found to be cleared to the liver, possibly by the reticulo-endothelial

system (RES), as this mechanism is known for clearance of PEG-ylated liposomes55 (base

particle for the design of SynthoPlateTM nanoconstructs). Interestingly ~30 % of the

injected dose of particles was found to still remain in the systemic circulation (blood),

suggesting that this amount can be potentially still available after the 1-hour period to keep localizing at the injury site for hemostatic action. Hematoxylin and eosin staining-based

histopathological analysis of harvested organs from SynthoPlateTM-injected un-injured animals showed no signs of off-target thrombi in the tissue samples (Figure 6-9).

219 Combining these findings with our previous report that SynthoPlateTM does not activate

(and aggregate) resting circulating platelets and also does not spontaneously trigger

Figure 6-9 Representative hematoxylin and eosin (H&E) stained histology images (32x magnification) of organ samples from pigs treated with saline, unmodified particles or SynthoPlateTM particles, show that SynthoPlateTM (as well as unmodified particles) does not cause any thrombotic risks in organs; scale bar of 200 μm shown in bottom row kidney image is applicable to all histology images.

220 thrombin (and fibrin) generation in plasma43,45, we rationalize that our SynthoPlateTM dose was safe in pigs, with minimal systemic pro-thrombotic, pro-coagulant and complement activation risks.

Figure 6-10 Schematic representation of pig femoral artery hemorrhage model setup. A CO2 sensor was placed at the end of the endotracheal tube and mechanical ventilation was provided, EKG electrodes were placed on the pig’s limbs, a pulse-oximeter probe was placed on the pig’s mouth, an esophageal temperature probe was placed to measure core temperature, an angiocatheter was placed in the carotid artery to acquire invasive blood pressure and also withdraw blood samples for ex vivo analysis, an angiocatheter was placed in the internal jugular vein to deliver saline (or nanoparticle treatments) via an infusion pump.

6.2.5 Hemostatic Efficacy Evaluation in Femoral Artery Bleeding Model

The pig femoral artery hemorrhage model is described in detail in the Methods section and an experimental schematic is shown in Figure 6-10. Briefly, an acute hemorrhagic injury

(near transection) was caused in the femoral artery with a 3.5 mm arterial punch and 1-

221 minute following injury, saline, unmodified (control) particles or SynthoPlateTM particles

Figure 6-11 Hemostatic efficacy analysis in injured pigs shows that [A] pigs administered with SynthoPlateTM had a reduced blood loss rate (blue) compared to those treated with saline (red) and unmodified particles (green), especially within first 30 min post treatment administration, where blood loss rate in SynthoPlateTM treated pig was significantly lower than those treated with unmodified (control) particles and saline (** p<0.01); [B] SynthoPlateTM administration in pigs also resulted in significantly lower total blood loss compared to saline administration (* p<0.05); [C] SynthoPlateTM administration in pigs resulted in maintenance and stabilization of a higher average mean arterial blood pressure (MAP) over time (data points shown for every 10 min); [D] SynthoPlateTM administration in pigs resulted in a significant enhancement of survival, with 100% surviving the first 60 min (golden hour) and 75% surviving the additional 60 min (p < 0.05), compared to administration of saline (25% survival by 60 min and 0% survival by ~90 min) or unmodified particles (50% survival by 90 min, 25% survival at 120 min).

were administered as a bolus dose through the jugular vein. Four animals per treatment group (i.e. 12 total) were used for these studies. Vitals, blood loss, hemodynamic

222 parameters and survival were monitored in real time for 60 minutes (golden hour) and an

additional 60 minutes (i.e. total observation for 120 min). During these observations, blood was also drawn through a carotid angiocatheter to run platelet aggregometry and ROTEM

analysis ex vivo. At the end of experiments, pigs were euthanized (unless they had already

succumbed earlier from their injury), and organs and clot were harvested for histological

and fluorescence-based distribution analysis (experimental details in Methods section). As shown in Figure 6-11A, bolus administration of SynthoPlateTM immediately following

femoral artery injury, resulted in substantial reduction of the blood loss rate over time. It is

important to note here that the condition of the saline-treated pigs rapidly declined after

injury as indicated by their mean arterial pressure (Figure 6-11C) and 75% of them died

before 60 minutes (Figure 6-11D); therefore, blood loss rate for this group appears to approach zero after this time point (Figure 6-11A). Meanwhile, pigs injected with unmodified (control) particle dose continued to undergo low rates of bleeding due to a delay in destabilization of mean arterial pressure (MAP) (Figure 6-11B), but ultimately one of them succumbed while actively bleeding around 60 minutes and two other died around 100 minutes (Figure 6-11D), though bleeding stopped in both of them at around

60 minutes possibly due to hemorrhage-induced hypotension (Figure 6-11A and C).

SynthoPlateTM –treated pigs exhibited significantly lower blood loss rates than comparison

groups during the first 30 min and achieved full hemostasis (at 15 min, 18 min, 30 min and

80 min for the 4 pigs respectively) such that the blood loss rate for this group also became

zero by the 80-minute time point (Figure 6-11A). Furthermore, these animals exhibited

stabilized MAP (Figure 6-11C) throughout the 120-minute period. Total blood loss for

SynthoPlateTM-treated pigs (normalized to body weight and time survived) was also found

223 to be significantly lower to comparison treatment (Figure 6-11B). As shown in Figure

6-11D, SynthoPlateTM-treated animals showed a significantly improved survival, with

Figure 6-12 Schematic representation and representative images of hemostasized injury site in the femoral artery of pigs treated with SynthoPlateTM or saline or control particles: [A] Representative hematoxylin and eosin (H&E) stained histology image (32x magnification) and [B] representative bright field image (10x magnification) of the site of injury (transected artery) with the injured vessel components (EC: endothelial cell, SMC: smooth muscle cell) and hemostatic clot in view for injured pig treated with SynthoPlateTM; [C1] representative fluorescence image (10x magnification) of FITC-anti-CD42b (green fluorescence) stained platelets, [C2] Rhodamine B-labeled (red fluorescence) SynthoPlateTM particles and [C3] overlay of C1 and C2, in the same field of view as the brightfield image B demonstrating that red fluorescent SynthoPlateTM is co-localized with green fluorescent platelets at the site of injury within the hemostatic clot; [D1, D2 and D3] are similar representative images for injured pig treated with saline, and [E1, E2 and E3] are similar representative images for injured pig treated with unmodified (control) particles showing that some platelets indeed localize at the injury site to promote hemostasis but presence of control particle is minimal at the site and therefore control particles do not have the ability to augment hemostasis; scale bar of 50 μm shown in C1 is applicable to all fluorescent images.

100% surviving the golden hour (first 60 minutes) and 75% maintaining survival at the end of 120 minutes. In comparison, 75% of the saline-treated animals died within the first 60 min and 75% of the unmodified (control) particle-treated animals progressively died during the 60-120 min period. Immunohistochemistry and fluorescence imaging analysis of harvested clots from the hemostasized femoral artery of SynthoPlateTM-treated animals revealed significant incorporation of SynthoPlateTM in the platelet-rich clot at the site of

224 injury (yellow overlay of green

platelets and red SynthoPlateTM), but

no such red fluorescent particle

presence was found at the arterial

injury site for unmodified (control)

particle-treated pigs (representative

fluorescence images for the three

treatments shown in Figure 6-12).

Also, similar to the safety experiments

in uninjured pigs, during the golden

hour (60 min) there were no drastic

changes in the C3:C3a plasma

concentration for SynthoPlateTM-

treated or unmodified particle-treated

animals compared to saline-treated

animals (Figure 6-13). Furthermore, Figure 6-13[A] Complement (C3:C3a) analysis data on the biodistribution of the unmodified drawn blood from pigs and [B] biodistribution data from

TM pigs subjected to femoral artery injury and administered particles and SynthoPlate (analyzed

intravenously with unmodified particles or SynthoPlate at the point of pig death or euthanasia) nanoconstructs. were similar (Figure 6-13).

Aggregometry studies showed that pigs treated with SynthoPlateTM had a stable platelet aggregation profile while pigs treated with unmodified (control) particle demonstrated a trend for decreasing platelet aggregation

225 and pigs treated with saline showed a trend for platelet hyper-aggregation (Figure 6-14A).

ROTEM analysis of blood revealed no presence of coagulopathy in SynthoPlateTM or saline

treated animals, as indicated by clotting time (CT), maximum clot firmness (MCF), and

amplitude 10 minutes after CT (A10) values from EXTEM and FIBTEM assays that all

(Figure 6-14B)56. Unmodified particle treated fall within ‘normal’ reference range, based

on criteria established in previous studies animals did begin to show signs of coagulopathy

at the 120-minute time point as indicated by MCF and A10 EXTEM and FIBTEM values

that are below the reference range (Figure 6-14B). Altogether, the data suggest that

SynthoPlateTM has the ability to reduce blood loss rate as well as total blood loss (Figure

6-11A and B), stabilize MAP (Figure 6-11C) and increase overall survival (Figure 6-11D)

after traumatic hemorrhage injury, due to the ability of SynthoPlateTM to enhance active platelet recruitment and aggregation to form hemostatic clot at the injury site and prevent coagulopathy.

6.3 Discussion

Traumatic exsanguinating hemorrhage continues to be a major cause of military and civilian fatalities in pre-hospital settings, where blood components have limited availability15,16. In such scenarios, volume resuscitation with saline or plasma expanders

shows only limited benefit57. Currently, studies are being directed at improving the

survivability in such conditions by anti-fibrinolytic agents like tranexamic acid (TXA) as

well as by improving blood component storage, portability and availability through cold-

storage and lyophilization27,58–63. Time is critical in mitigation of exsanguinating

226

Figure 6-14 [A] Representative ex vivo Lumi-aggregometry analysis data and [B] ROTEM analysis data (CT, MCF and A10 parameters) of blood samples drawn from pigs after being subjected to femoral artery injury and administered intravenously with saline or unmodified particles or SynthoPlate nanoconstructs.

hemorrhage since significant mortalities occur within 1-2 hours of injury, and hemorrhage control within this time can significantly improve survival49,50. Therefore, intravenously

227 transfusable synthetic RBC and platelet surrogates that can be administered early at point- of-injury can have tremendous benefit in emergency mitigation of hemorrhagic shock39–41.

In the area of synthetic platelet surrogates, past approaches have focused mostly on

mimicking only the aggregatory mechanism of platelets by decorating micro- and

nanoparticles with fibrinogen or fibrinogen-relevant peptides, essentially creating a ‘super-

fibrinogen’ construct39,41. However, platelet’s native hemostatic action is a co-operative

concert of injury site-selective adhesion and aggregation mechanisms64. Hence in our

design, we hypothesized that integrating these two mechanisms on a single particle

platform will allow a superior mimicry of platelet’s hemostatic mechanisms selectively at

the bleeding site while minimizing off-target effects. We previously tested this hypothesis

via heteromultivalent decoration of liposomal as well as albumin-based nanoparticles with

VBP, CBP and FMP motifs and successfully validated the superior hemostatic efficacy of this integrative design compared to ‘adhesion only’ and ‘aggregation only’ designs43,45.

SynthoPlateTM is a refined liposome-based version stemming from this integrative design44, and we have recently demonstrated its hemostatic efficacy in a non-trauma

bleeding model in severely thrombocytopenic mice46. Building on this, here we

investigated the translational promise of the SynthoPlateTM technology in vitro

(sterilization effects and storage stability) and in vivo (large animal hemorrhage model).

Our studies demonstrate that sterilization of SynthoPlateTM by either filtration or E-beam

does not affect the product’s stability and platelet-inspired biofunctionalities. Our studies

also indicate that long-term storage in suspension does not affect SynthoPlateTM stability.

228 Altogether, these data suggest that sterile suspensions of SynthoPlateTM can potentially

allow long-term storage and availability for hemorrhage mitigation in pre-hospital settings.

Our in vitro and in vivo studies clearly demonstrate that the SynthoPlateTM dose did not

have any cytotoxicity and systemic risks in the experimental conditions tested. Future

studies will be directed at assessing dose escalation-based pharmacology and toxicology

analysis, as well as, evaluating the potential of multiple doses. Following arterial injury

and bolus administration of SynthoPlateTM (or control) particles, our results show that

compared to saline (0% survival), SynthoPlateTM treatment renders 100% survival during

the ‘golden hour’. Interestingly, unmodified (control) particles showed ~50% survival

during this first hour, however, survival for this treatment group fell to 25% during the

second hour of the experimental window. This can be attributed to the fact that the

unmodified particles seemed to transiently improve MAP, possibly due the colloidal nature

of injected liposome suspension transiently increasing blood colloidal osmotic pressure

(BCOP). However, since unmodified particles were incapable of biofunctionally

enhancing hemostasis, these animals continued to bleed at a higher rate and succumbed. In

contrast, a combination of hemodynamic stabilization and hemostatic enhancement

significantly improved the survival of the SynthoPlateTM-treated pigs within and beyond

the ‘golden hour’. It is to be noted that, because of MAP stabilization and longer survival

time, the blood loss rate data Figure 6-11A) apparently suggests that the SynthoPlateTM- treated animals continued to bleed (albeit at a very low rate) up to and beyond the ‘golden hour’. However, this is a ‘survival bias’, since majority of injured animals treated with saline or unmodified particles essentially could not recover from being rapidly

229 hypotensive, which resulted in ‘apparent reduction’ of blood loss rate and ultimately death.

Interestingly, several ‘externally applicable’ hemostatic dressings were recently evaluated

in a pig femoral artery hemorrhage model, that showed overall survival to be 50-60% in

the 1-hour period47. Considering that in our studies, administering only a single bolus

intravenous dose of SynthoPlateTM, with no additional dressings, resulted in 100% survival

of the pigs for 60 minutes (and 75% for 120 min), we rationalize that the superior

hemostatic capability of SynthoPlateTM can be potentially combined with externally

applied hemostatic dressings, to further improve the hemostatic outcomes in pre-hospital

scenarios. We also note that in the current studies the treatments were administered within

1 minute after injury and we have not yet evaluated the ‘administration time window’ for

SynthoPlateTM following injury to determine at what point of delayed administration its

hemostatic benefit may fail to increase survival. Our future studies will further evaluate

these possibilities in porcine trauma models. Our current findings therefore strongly

demonstrate the potential of SynthoPlateTM as a viable synthetic platelet surrogate for

point-of-injury management of traumatic hemorrhagic.

6.4 Materials and Methods

6.4.1 Materials

The lipid components, namely, 1,2-Distearoyl-sn-glycero-3-phosphocholine (DSPC), 1,2-

Distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-1000]

(DSPE-mPEG1000), 1,2-Distearoyl-sn-glycero-3-phosphoethanolamine-N-

[maleimide(polyethylene glycol)-2000] (DSPE-PEG2000-Mal) and 1,2-distearoyl-sn-

glycero-3-phosphoethanolamine-N-[azido(polyethylene glycol)-2000] (DSPE-PEG2000-

230 Azide) were purchased from Avanti Polar Lipids (Alabaster, AL, USA). Rhodamine B- dihexadecanoyl-sn-glycero-3-phosphoethanolamine (DHPE-RhB) was purchased from

Invitrogen (Carlsbad, CA, USA). The peptides CTRYLRIHPQSWVHQI (VBP), C[GPO]7

(CBP) and cyclo-{Pra}CNPRGD{Tyr(OEt)}RC (FMP) were custom-synthesized by

Genscript (Piscataway, NJ, USA). VBP and CBP were conjugated to DSPE-PEG2000-Mal

65 via thiol-maleimide coupling and FMP was conjugated to DSPE-PEG2000-azide via

copper-catalyzed alkyne-azide cycloaddition (CuCAAC)66 (schematics shown in

Supplementary Figure S.1), conjugates were purified by dialysis and characterized by

MALDI-TOF mass spectrometry. Sterile normal saline solution (0.9% NaCl) was

purchased from Baxter (Deerfield, IL, USA). Cholesterol, rat tail type I collagen, copper(II)

sulfate (CuSO4), Tris(3-hydroxypropyltriazolylmethyl)amine (THPTA), sodium

ascorbate, decaethylene glycol monododecyl (C12E10) and the Mammalian Cell Lysis Kit

were from Sigma-Aldrich (Saint Louis, MO, USA). Ethylenediamine Tetraacetic Acid

(EDTA), cellulose dialysis tubing (MWCO 2k and 3.5k), phosphate buffered saline (PBS),

chloroform, methanol, LAL Chromogenic Endotoxin Quantification kit, Penicillin-

Streptomycin (PS), DMEM without phenol red and Vybrant® MTT Cell Proliferation

Assay Kit were purchased from Fisher Scientific (Pittsburgh, PA, USA). Citrated human

whole blood was drawn from healthy, aspirin-refraining donors via venipuncture according

to IRB approved protocols. Adenosine di-phosphate (ADP) and soluble calf skin type I

collagen were from Bio/Data Corporation (Horsham, PA, USA), and human vWF (FVIII-

free) was from Hematologic Technologies (Essex Junction, VT, USA). For rotational

thromboelastometry (ROTEM) analysis, all reagents were purchased from ROTEM

(Munich, Germany). For in vivo studies on pigs, Yorkshire Farm Pigs were purchased from

231 Shoup Investments Ltd. Telazol, isoflurane and pentobarbital were obtained from Patterson

Veterinary (Greely, CO, USA). The Beadbug Mircotube Homogenizer was purchased from

Benchmark Scientific (Edison, NJ, USA). The Guinea Pig Complement C3 ELISA kit and

Human Complement C3a des Arg ELISA kit were from Abcam (Cambridge, MA, USA).

Mouse-derived embryonic fibroblast cell line 3T3 was provided graciously by the

laboratory of Eben Alsberg, Case Western Reserve University. High glucose Dulbecco’s

Modified Eagle Medium (DMEM) and Fetal Bovine Serum (FBS) were from American

Type Culture Collection (Manassas, VA, USA).

6.4.2 Manufacture of SynthoPlateTM

SynthoPlateTM was manufactured using ‘film rehydration and extrusion’ technique as

42,44 described previously . Briefly, DSPC, cholesterol, DSPE-PEG2000-VBP, DSPE-

PEG2000-CBP, DSPE-PEG2000-FMP and DSPE-mPEG1000 were homogeneously mixed at

0.475, 0.45, 0.0125, 0.0125, 0.025 and 0.025 mole fractions, respectively, in 1:1 chloroform:methanol. Solvent was removed via rotary evaporation, and the thin lipid film was rehydrated with normal saline solution (0.9% NaCl) at a concentration of 1 x 105 moles lipid per mL. This lipid suspension was subjected to 10 freeze/thaw cycles and subsequent extrusion through 200nm pore diameter polycarbonate membrane using a pneumatic extruder (Northern Lipids, Burnaby, Canada) to create heteromultivalently decorated

SynthoPlateTM vesicles. Dynamic light scattering (DLS) and electron microscopy

characterization indicated fresh-made vesicles were ~200nm in diameter.

232 6.4.3 Long term Storage-in-Suspension Studies

In order to evaluate the stability of SynthoPlateTM under storage, particles were

manufactured as described above, packaged into polypropylene tubes, purged with

nitrogen gas and stored for up to 6 months at room temperature (~25°C). Particle size

distribution was measured weekly via Dynamic Light Scattering (DLS). This data was

considered as reflective of particle stability over time, since large variations in size would

be indicative of particle instability.

6.4.4 Studies on SynthoPlateTM sterilization and its biofunctional effects

SynthoPlateTM samples were evaluated at Lexamed (Toledo, OH, USA) for sterilization

via filtration and E-beam exposure. First, it was necessary to validate the sterilization

methods by demonstrating that SynthoPlateTM itself does not inhibit the growth of common contaminating organisms and hence does not interfere with the standard sterilization assays. For this, SynthoPlateTM samples were inoculated with 6 different challenge

organisms, namely Staphylococcus aureus, Kocuria rhizophila, Clostridium sporogenes,

Bacillus subtilis spizizenii, Candida albicans and Aspergillus brasiliensis, and allowed to

grow in appropriate conditions for up to 5 days. Any differences in growth profiles between

samples with or without SynthoPlateTM over the 5-day period were recorded. Next,

SynthoPlateTM samples were sterilized by filtration in 0.2 µm filter and by E-beam

exposure at both 25 and 40 kGy doses. Sterility was assessed by observation of the samples

for any visible growth of microorganisms over the course of 7 days. Furthermore,

endotoxin burden in the formulation was evaluated using a standard chromogenic limulous

ameobocyte lysate (LAL) assay. Briefly, nanoparticle samples were incubated with 0.5%

233 (v/v) of 1.5% C12E10 detergent for 20 minutes at 65°C followed by 1 hour at room

temperature to solubilize the lipid components of SynthoPlateTM and release any masked

endotoxin67. Endotoxin burden in the detergent-treated SynthoPlateTM samples was

quantified using the LAL assay.

To assess whether sterilization affects SynthoPlateTM particle stability, post-sterilization

particle size distribution was measured by DLS and compared to fresh-made (unsterilized) particles. Furthermore, in order to measure whether sterilization affects the platelet- inspired biofunctional abilities of SynthoPlateTM, specific adhesion and aggregation assays

were performed as adapted from our previous reports43,46. To test whether sterilization

affects the platelet-inspired adhesive function of CBP and VBP moieties on SynthoPlateTM

(to collagen and vWF, respectively), acid washed glass coverslips were coated with a

solution of 1:1 vWF:collagen, and 200µl RhB-labeled SynthoPlateTM suspensions (0.4

mg/ml), un-sterilized or sterilized, were incubated on these coverslips in a 12-well plate

for 30 minutes at 37oC on a shaker set at 100 rpm. Following this, SynthoPlateTM

suspension was removed, and loosely bound particles were further washed away with PBS.

The coverslips were then mounted onto microscope slides and imaged (using Zeiss

AxioObserver inverted fluorescence microscope) for adhered SynthoPlateTM (Rhodamine

B fluorescence, λex=560 nm, λem=580 nm). To test whether sterilization affects the ability

of FMP moieties to bind to active platelet integrin GPIIb-IIIa to amplify platelet aggregation, platelet lumi-aggregometry assays were performed using a PAP-8E (Bio/Data

Corp, Horsham, PA, USA) instrument. For this, human whole blood was collected in

citrated tubes from healthy consenting donors in accordance with protocols and guidelines

234 approved by the Institutional Review Board (IRB) of University Hospitals Cleveland

Medical Center (UHCMC, IRB Number 12-16-06). Informed consent was obtained from

all subjects as per this IRB-approved guideline. The whole blood was centrifuged at 150

x g for 15 minutes to obtain platelet rich plasma (PRP). Some of the PRP was further

centrifuged at 2500 x g for 25 minutes to obtain platelet poor plasma (PPP). Aggregometry

studies were carried out at 37°C with stirring at 1200 rpm. ADP-mediated aggregation was

assessed with 225µl of 100% PRP or PRP diluted 50% (v/v) with PPP supplemented with

fresh-made (un-sterilized) or sterilized (filtration and E-beam) SynthoPlateTM particles

with a particle concentration of 5 x 1010/mL, with or without adding agonist (2 x 10-5 M

ADP). These adhesion and aggregation results of SynthoPlateTM post-sterilization were

compared to that for fresh-made (unsterilized) SynthoPlateTM.

6.4.5 In vitro cytotoxicity assay with SynthoPlateTM

3T3 cells were grown in DMEM supplemented with 10% FBS and 1% PS. Cultures were

maintained in a humidified atmosphere of 5% CO2 at 37°C. 3T3 cells were seeded into 96

well plates at 5000 cells/well. Within 24 hours, media was removed and seeded wells were

treated with media containing SynthoPlateTM particles versus media without the particles.

SynthoPlateTM particles were applied at 0.5 nM and 1 nM concentrations, which

correspond to the dose administered to pigs and double the dose administered to the pigs respectively. All treatments were suspended in phenol red free DMEM and an n=8 was maintained for all test groups. The treated cells were incubated at normal culture conditions

for 24 hours. After 24 hours, a standard MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-

diphenyltetrazolium bromide) assay was used to assess metabolic viability of the cells. In

235 this assay, metabolic activity (i.e. cell viability) reduces the MTT dye to insoluble formazan

which has a purple color, and the intensity of this can be assessed by colorimetric (i.e. UV-

Visible spectrometry) technique to reflect cell viability. For this, post incubation with or

without SynthoPlateTM, media was removed and fresh phenol red free media and MTT

stock solution were applied to all wells. The cells were incubated at normal culture

conditions for 4 hours. After 4 hours, a solution of SDS-HCl was mixed into all wells. and

incubated for a final period of 4 hours. Finally, the formazan intensity (i.e. cell metabolic viability) was quantified with UV-visible spectrometry, measuring absorbance at 570 nm.

6.4.6 In vivo Safety and Biodistribution Studies in Pigs

All in vivo porcine studies were carried out in accordance with relevant guidelines and

regulations approved by Case Western Reserve University Institutional Animal Care and

Use Committee (IACUC, Protocol Number 2015-0135). Yorkshire Farm Pigs 25-36kg

were acclimated to the laboratory space for 48 hours. Pigs were fasted for 12 hours before

experiments but had free access to water. Pigs were sedated with Telazol (6-8mg/kg)

injected intramuscularly, intubated with an endotracheal tube and anesthetized with

isoflurane (1-5% to effect). A CO2 sensor was placed at the end of the endotracheal tube.

Mechanical ventilation was provided to keep end-tidal CO2 and respiration rate initially at

normal values. EKG electrodes were placed on the pig’s limbs. A pulse-oximeter probe

was placed on the pig’s mouth (cheek or tongue). An esophageal temperature probe was

placed to measure core temperature, which was maintained between 36oC-38oC with a water-filled warming blanket. An angiocatheter was placed in the carotid artery to acquire invasive blood pressure and withdraw blood samples for ex vivo testing. Another

236 angiocatheter was placed in the internal jugular vein to deliver saline or particle treatments.

The 50ml dose of SynthoPlateTM (or control particles) to be administered in the pigs was

1X1011 particles/ml (per calculation shown in Supplementary Section S.4) followed by a

450ml saline infusion. Invasive arterial blood pressure, CO2, SpO2, temperature, and heart

rate were recorded every 30 seconds for the first 10 minutes, every minute for the next 20

minutes, and every 5 minutes thereafter for a total of 60 minutes. Arterial blood was drawn

via the carotid artery angiocatheter at baseline, 30 minutes, and 60 minutes. At the end of

experiments, pigs were euthanized with an I.V. overdose of pentobarbital (0.22ml/kg).

In order to evaluate the systemic safety of SynthoPlateTM in pigs within the ‘golden hour’

time-frame, without any confounding effects from injury, treatments were administered (at

the bolus dose stated previously) in a pilot group of non-injured animals (total n=6; saline:

n=2, unmodified particles: n=2, SynthoPlateTM: n=2), and the animals were observed for 1 hour. The surgeons were blinded to the administered treatments, and blinding was ensured by a researcher who assigned a randomly generated code to each treatment. The effects of the administered treatments on vitals, blood chemistry, platelet function (aggregation), and blood clot viscoelastic properties were examined. It has been reported that nanoparticles

(including liposomes) can sometimes lead to a hypersensitivity reaction and complement activation (known as complement activation-related pseudoallergy or CARPA)52.

Therefore, the risk for such reactions was assessed ex vivo. For this, specific complement

activation marker (C3 to C3a) was measured in the blood drawn from the pig at baseline

as well as 30 and 60 minutes post administration of saline, unmodified particles or

SynthoPlateTM. In complement activation, C3 is cleaved into C3a and C3b leading to a

237 decrease in plasma C3 levels and an increase in plasma C3a levels, and thus the C3:C3a

ratio can indicate such risks. For these studies, platelet poor plasma was isolated from pig

blood samples by centrifuging for 25 minutes at 2500 x g and measurements of C3 and

C3a were carried out using Guinea Pig Complement C3 ELISA kit and Human

Complement C3a des Arg ELISA kit respectively (Abcam, Cambridge, MA, USA). Pig

whole blood samples were also analyzed for pH, calculated base excess, carbon dioxide

(CO2), calculated bicarbonate, partial pressure of oxygen (PaO2), oxygen saturation (SaO2), white blood cell count (WBC), hematocrit (Hct), hemoglobin (Hg), red blood cell count

(RBC), platelet count (Plt), mean corpuscular volume in femtoliters (MCV fL), mean corpuscular hemoglobin concentration (MCHC), red blood cell distribution width-

coefficient of variation (RDW-CV), prothrombin time (PT), international normalized ratio

(INR), and, activated partial thromboplastin time (APTT) compared between blood of

saline-treated, unmodified particle-treated and SynthoPlateTM-treated pigs. Furthermore,

platelet aggregometry was conducted on PRP isolated from pig blood samples, using

methods described previously. In addition, effect of the various treatment groups on clot

viscoelastic properties was assessed by Rotational Thromboelastometry (ROTEM) via

analysis of drawn blood samples.

At the 1-hour time point pigs were euthanized and major organs (lungs, heart, liver, kidney,

spleen) were harvested, weighed and fixed in formalin. Three random samples 200-350mg

were removed from each pig organ and placed into pre-weighed homogenizing tubes, and the wet weight of each organ sample was recorded. The samples were then vacuum dried and their dry weights recorded. The dry organs were homogenized with a BeadBug

238 Microtube Homogenizer using a Mammalian Cell Lysis kit and then centrifuged (20

minutes, 12,000 x g) to separate the organ tissue from the supernatant. The supernatant was

added to 1:1 methanol:chloroform to dis-assemble the Rhodamine B labeled lipids of

SynthoPlateTM in the samples. These samples were analyzed with Ultra Performance

Liquid Chromatography (UPLC) using a fluorescence detector (λex = 560nm, λem = 580nm)

to assess Rhodamine B fluorescence. Biodistribution of the SynthoPlateTM particles was

determined by calculating the nanogram (ng) of particles per milliliter (ml) of supernatant

utilizing an appropriate calibration curve that correlates RhB fluorescence intensity to

particle concentration. Samples from each organ were also processed for histological

analysis. For this, samples were embedded in paraffin and sectioned and stained with

hematoxylin and eosin (H & E stain). The resultant slides where imaged on a Zeiss

Axiophot microscope and images were obtained with Zeiss Axiovision camera and

software.

6.4.7 Femoral Artery Bleeding Model in Pigs

Figure 5 shows a representative experimental set-up of the model for these studies. Pigs

(total n=12; Saline: n=4, unmodified particles: n=4, SynthoPlateTM: n=4) were prepared as

described previously. The skin at the inguinal area of the thigh was incised ~10cm, the

femoral artery was exposed and injured (near transection) using a 3.5mm aortic punch

(Scanlan International, St Paul, Minnesota, USA). Saline or particle treatments were administered 1 minute after injury via the jugular at a rate of 50ml/min. Blood loss from femoral artery injury was measured by carefully suctioning the shed blood without disturbing the injury site, into a container that was weighed every 30 seconds for the first

239 10 minutes, every minute for the next 20 minutes, and every 5 minutes thereafter until

bleeding stopped due to death or hemostasis. Blood mass was correlated to blood volume

by approximating blood density to be 1000 kg/m3. With saline administration, which is a

standard volume resuscitation strategy for treating hemorrhage in pre-hospital scenarios, this injury resulted in an average of 31.5 mL/kg blood loss in 24 minutes, and showed only

25% survival within the first 60 min (i.e. within the ‘golden hour’ period). At the end of

experiments, pigs were euthanized with an I.V. overdose of pentobarbital (0.22ml/kg).

6.4.8 Hemostatic Efficacy Evaluation in Femoral Artery Bleeding Model

The femoral artery bleeding model was performed as described above. Induction of the

injury was designated as time-point zero (0). Saline or particle (unmodified control or

SynthoPlateTM) treatments were administered 1 minute after injury via the jugular, (50 ml

particle concentration 1 x 1011 particles/ml followed by 450 ml saline) at 50 ml per minute.

Blood loss was measured as described previously. Arterial blood was drawn via the carotid artery angiocatheter at baseline, and 15, 30, 60 and 120 minutes, to run ex vivo analysis.

At the end of experiments, pigs were euthanized, major organs (lungs, heart, liver, kidney, spleen) and clot were harvested, weighed and fixed in formalin. Organ samples were prepared and analyzed for SynthoPlateTM (or control unmodified particle) biodistribution

at time of death as described previously for uninjured animals. Sections of the femoral

artery injury site from SynthoPlateTM-treated animals were deparaffinized and rehydrated

by washing with xylene 2 times for 2 minutes, 1:1 xylene: ethanol for 3 minutes, ethanol 2

times for 3 minutes, 95% ethanol for 3 minutes, 70% ethanol for 3 minutes and finally 50%

ethanol for 3 minutes. Antigen retrieval was performed by incubating in Tris-EDTA buffer

240 in a water bath at 60oC overnight. Slides were washed, blocked in 10% serum with 1%BSA

in TBS for 2 hours, incubated with the FITC-labeled (green fluorescent) CD42b antibody

(for staining platelet glycoprotein GPIbα), washed, protected with a cover-slip and then imaged (using a Zeiss AxioObserver inverted fluorescence microscope). Brightfield,

SynthoPlateTM fluorescence (red, Rhodamine B) and platelet fluorescence (green, FITC)

were captured for the same field of view.

6.4.9 Statistical Analysis

Statistical analysis of aggregometry, biodistribution, blood loss data, and complement data

was done using two-way ANOVA with a Bonferroni post-test. Vitals data, blood analysis data and ROTEM data were analyzed using one-way ANOVA with a Tukey post-test.

Survival data was analyzed using a Log-rank test. In all analyses, significance was considered to be p < 0.05.

6.5 Acknowledgements

This research was supported in part by funding from the National Institutes of Health

(NIH), specifically National Heart Lung and Blood Institute (NHLBI Grant R01

HL121212, PI: Sen Gupta) and National Institute of General Medical Sciences (NIGMS,

Grant R35 GM119526, PI: Neal; Co-I: Sen Gupta), as well as, funding from Case Coulter

Translational Research Partnership (Grant 419, PI: Sen Gupta), National Center for

Accelerated Innovation – Cleveland Clinic (NIH Grant U54HL119810, PI: Sen Gupta),

and Ohio Third Frontier Technology Validation and Start-up Funds (TECG20160125, PI:

Sen Gupta). The authors also acknowledge NIH National Center for Research Resources

241 (Grant1 C06 RR12463-01, PI: Kutina) for supporting the laboratory facilities for Sen Gupta in the Wickenden Building at Case Western Reserve University. The authors acknowledge

Steve Schomisch and Cassie Cipriano for their assistance with animal studies. The content of this publication is solely the responsibility of the authors and does not represent the official views of the NIH and other funding entities.

6.6 Author Contributions

A. Sen Gupta, V. Kashyap and M. Neal directed research. E. Niedoba, C. L. Pawlowski, and D. A. Hickman prepared all particles, and dynamic light scattering characterizations.

D. A. Hickman, C. L. Pawlowski, N. F. Luc, U. D. S. Sekhon, J. Marks, and M. Sun carried out ex vivo aggregometry and ROTEM studies. A. Girish carried out in vitro cytotoxicity studies with SynthoPlateTM. S. Ganjoo and S. Huang carried out in vivo complement studies. D. A. Hickman and C. L. Pawlowski carried out histology and immunohistochemistry studies. D. A. Hickman, C. L. Pawlowski, A. Kim, A. Shevitz, S.

Ganjoo, S. Huang, E. Niedoba, N. F. Luc, U. D. S. Sekhon, J. Marks, and M. Sun carried out in vivo pig studies. D.A. Hickman, C.L. Pawlowski, M. Dyer, M. Neal and A. Sen

Gupta carried out data analysis, interpretation and wrote the paper.

6.7 Competing Financial Interests

A. Sen Gupta is a co-inventor on patents related to SynthoPlateTM technology: US 9107845 and US 9,6363,383. The SynthoPlateTM trademark is currently recorded with USPTO (U.S.

Serial Number: 86-829,160). The other authors state that they have no conflict of interest.

242 6.8 References

1. Kauvar DS, Lefering R, Wade CE. Impact of Hemorrhage on Trauma Outcome:

An Overview of Epidemiology, Clinical Presentations, and Therapeutic

Considerations. 2006 Jun;60(Supplement):S3–S11. Available from:

https://insights.ovid.com/crossref?an=00005373-200606001-00002

2. Holcomb JB, McMullin NR, Pearse L, Caruso J, Wade CE, Oetjen-Gerdes L,

Champion HR, Lawnick M, Farr W, Rodriguez S, Butler FK. Causes of Death in

U.S. Special Operations Forces in the Global War on Terrorism. 2007;245(6):986–

991. PMID: 17522526

3. Blackbourne LH, Baer DG, Eastridge BJ, Kheirabadi B, Kragh JF, Cap AP,

Dubick M a., Morrison JJ, Midwinter MJ, Butler FK, Kotwal RS, Holcomb JB.

Military medical revolution. 2012;73:S372–S377. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201212005-00002 PMID: 23192060

4. Cohen MJ, Kutcher M, Redick B, Nelson M, Call M, Knudson MM, Schreiber

MA, Bulger EM, Muskat P, Alarcon LH, Myers JG, Rahbar MH, Brasel KJ,

Phelan HA, del Junco DJ, Fox EE, Wade CE, Holcomb JB, Cotton BA, Matijevic

N, PROMMTT Study Group on behalf of the P study. Clinical and mechanistic

drivers of acute traumatic coagulopathy. NIH Public Access; 2013 Jul;75(1 Suppl

1):S40-7. Available from: http://www.ncbi.nlm.nih.gov/pubmed/23778510 PMID:

23778510

5. Dorlac WC, DeBakey ME, Holcomb JB, Fagan SP, Kwong KL, Dorlac GR,

Schreiber MA, Persse DE, Moore FA, Mattox KL. Mortality from isolated civilian

243 penetrating extremity injury. 2005 Jul;59(1):217–22. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/16096567 PMID: 16096567

6. Smith ER, Shapiro G, Sarani B. The profile of wounding in civilian public mass

shooting fatalities. 2016 Jul;81(1):86–92. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/26958801 PMID: 26958801

7. van Oostendorp SE, Tan ECTH, Geeraedts LMG. Prehospital control of life-

threatening truncal and junctional haemorrhage is the ultimate challenge in

optimizing trauma care; a review of treatment options and their applicability in the

civilian trauma setting. 2016 Dec 13;24(1):110. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/27623805 PMID: 27623805

8. Holcomb JB, Tilley BC, Baraniuk S, Fox EE, Wade CE, Podbielski JM, del Junco

DJ, Brasel KJ, Bulger EM, Callcut R a, Cohen MJ, Cotton B a, Fabian TC, Inaba

K, Kerby JD, Muskat P, O’Keeffe T, Rizoli S, Robinson BRH, Scalea TM,

Schreiber M a, Stein DM, Weinberg J a, Callum JL, Hess JR, Matijevic N, Miller

CN, Pittet J, Hoyt DB, Pearson GD, Leroux B, van Belle G. Transfusion of

Plasma, Platelets, and Red Blood Cells in a 1:1:1 vs a 1:1:2 Ratio and Mortality in

Patients With Severe Trauma. 2015;313:471. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/25647203 PMID: 25647203

9. Holcomb JB, del Junco DJ, Fox EE, Wade CE, Cohen MJ, Schreiber MA, Alarcon

LH, Bai Y, Brasel KJ, Bulger EM, Cotton BA, Matijevic N, Muskat P, Myers JG,

Phelan HA, White CE, Zhang J, Rahbar MH, PROMMTT Study Group. The

prospective, observational, multicenter, major trauma transfusion (PROMMTT)

study: comparative effectiveness of a time-varying treatment with competing risks.

244 NIH Public Access; 2013 Feb;148(2):127–36. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/23560283 PMID: 23560283

10. Holcomb JB, Jenkins D, Rhee P, Johannigman J, Mahoney P, Mehta S, Cox ED,

Gehrke MJ, Beilman GJ, Schreiber M, Flaherty SF, Grathwohl KW, Spinella PC,

Perkins JG, Beekley AC, McMullin NR, Park MS, Gonzalez EA, Wade CE,

Dubick MA, Schwab CW, Moore FA, Champion HR, Hoyt DB, Hess JR. Damage

Control Resuscitation: Directly Addressing the Early Coagulopathy of Trauma.

2007 Feb;62(2):307–310. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17297317 PMID: 17297317

11. Borgman MA, Spinella PC, Perkins JG, Grathwohl KW, Repine T, Beekley AC,

Sebesta J, Jenkins D, Wade CE, Holcomb JB. The Ratio of Blood Products

Transfused Affects Mortality in Patients Receiving Massive Transfusions at a

Combat Support Hospital. 2007 Oct;63(4):805–813. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/18090009 PMID: 18090009

12. Davenport R. Haemorrhage control of the pre-hospital trauma patient. BioMed

Central; 2014;22(Suppl 1):A4. Available from:

http://sjtrem.biomedcentral.com/articles/10.1186/1757-7241-22-S1-A4

13. Ishikura H, Kitamura T. Trauma-induced coagulopathy and critical bleeding: the

role of plasma and platelet transfusion. BioMed Central; 2017 Dec 20;5(1):2.

Available from: http://jintensivecare.biomedcentral.com/articles/10.1186/s40560-

016-0203-y

14. Carmen R. The selection of plastic materials for blood bags. 1993 Jan;7(1):1–10.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/8431654 PMID: 8431654

245 15. Lambert MP, Sullivan SK, Fuentes R, French DL, Poncz M. Challenges and

promises for the development of donor-independent platelet transfusions.

American Society of Hematology; 2013 Apr 25;121(17):3319–24. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/23321255 PMID: 23321255

16. Spinella PC, Dunne J, Beilman GJ, O’Connell RJ, Borgman MA, Cap AP, Rentas

F. Constant challenges and evolution of US military transfusion medicine and

blood operations in combat. Wiley/Blackwell (10.1111); 2012 May 1;52(5):1146–

1153. Available from: http://doi.wiley.com/10.1111/j.1537-2995.2012.03594.x

17. Moroff G, Holme S. Concepts about current conditions for the preparation and

storage of platelets. 1991 Jan;5(1):48–59. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/1802276 PMID: 1802276

18. Devine D V., Serrano K. The Platelet Storage Lesion. 2010 Jun;30(2):475–487.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/20513565 PMID: 20513565

19. Bordin JO, Heddle NM, Blajchman MA. Biologic effects of leukocytes present in

transfused cellular blood products. 1994 Sep 15;84(6):1703–21. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/8080981 PMID: 8080981

20. Heddle NM, Klama LN, Griffith L, Roberts R, Shukla G, Kelton JG. A

prospective study to identify the risk factors associated with acute reactions to

platelet and red cell transfusions. 1993 Oct;33(10):794–7. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/8236418 PMID: 8236418

21. Blajchman MA. Bacterial contamination and proliferation during the storage of

cellular blood products. 1998;74 Suppl 2:155–9. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/9704439 PMID: 9704439

246 22. Schreiber GB, Busch MP, Kleinman SH, Korelitz JJ. The risk of transfusion-

transmitted viral infections. The Retrovirus Epidemiology Donor Study. 1996 Jun

27;334(26):1685–90. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/8637512 PMID: 8637512

23. Corash L. Inactivation of viruses, bacteria, protozoa, and leukocytes in platelet

concentrates: Current research perspectives. W.B. Saunders; 1999 Jan 1;13(1):18–

30. Available from:

https://www.sciencedirect.com/science/article/pii/S0887796399800856

24. Milford EM, Reade MC. Comprehensive review of platelet storage methods for

use in the treatment of active hemorrhage. 2016 Apr;56 Suppl 2:S140-8. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/27100750 PMID: 27100750

25. Hoffmeister KM, Josefsson EC, Isaac NA, Clausen H, Hartwig JH, Stossel TP.

Glycosylation restores survival of chilled blood platelets. American Association

for the Advancement of Science; 2003 Sep 12;301(5639):1531–4. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/12970565 PMID: 12970565

26. Pidcoke HF, McFaul SJ, Ramasubramanian AK, Parida BK, Mora AG, Fedyk CG,

Valdez-Delgado KK, Montgomery RK, Reddoch KM, Rodriguez AC, Aden JK,

Jones JA, Bryant RS, Scherer MR, Reddy HL, Goodrich RP, Cap AP. Primary

hemostatic capacity of whole blood: a comprehensive analysis of pathogen

reduction and refrigeration effects over time. Wiley/Blackwell (10.1111); 2013

Jan;53:137S–149S. Available from: http://doi.wiley.com/10.1111/trf.12048

27. Reddoch KM, Pidcoke HF, Montgomery RK, Fedyk CG, Aden JK,

Ramasubramanian AK, Cap AP. Hemostatic Function of Apheresis Platelets

247 Stored at 4°C and 22°C. 2014 May;41:54–61. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=00

024382-201405001-00012

28. Seghatchian J, de Sousa G. Pathogen-reduction systems for blood components: the

current position and future trends. 2006 Dec;35(3):189–96. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17110168 PMID: 17110168

29. Reddy HL, Dayan AD, Cavagnaro J, Gad S, Li J, Goodrich RP. Toxicity testing of

a novel riboflavin-based technology for pathogen reduction and white blood cell

inactivation. 2008 Apr;22(2):133–53. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/18353253 PMID: 18353253

30. Janetzko K, Hinz K, Marschner S, Goodrich R, Klüter H. Pathogen reduction

technology (Mirasol®) treated single-donor platelets resuspended in a mixture of

autologous plasma and PAS. Wiley/Blackwell (10.1111); 2009 Oct 1;97(3):234–

239. Available from: http://doi.wiley.com/10.1111/j.1423-0410.2009.01193.x

31. Noorman F, van Dongen TTCF, Plat M-CJ, Badloe JF, Hess JR, Hoencamp R.

Transfusion: -80°C Frozen Blood Products Are Safe and Effective in Military

Casualty Care. 2016 Dec 13;11(12):e0168401. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/27959967 PMID: 27959967

32. Acker JP, Marks DC, Sheffield WP. Quality Assessment of Established and

Emerging Blood Components for Transfusion. Hindawi; 2016 Dec 14;2016:1–28.

Available from: https://www.hindawi.com/archive/2016/4860284/

33. Nasiri S. Infusible platelet membrane as a platelet substitute for transfusion: an

overview. SIMTI Servizi; 2013 Jul;11(3):337–42. Available from:

248 http://www.ncbi.nlm.nih.gov/pubmed/23736926 PMID: 23736926

34. Fitzpatrick GM, Cliff R, Tandon N. Thrombosomes: a platelet-derived hemostatic

agent for control of noncompressible hemorrhage. 2013 Jan;53 Suppl 1:100S–

106S. Available from: http://www.ncbi.nlm.nih.gov/pubmed/23301961 PMID:

23301961

35. Sum R, Hager S, Pietramaggiori G, Orgill DP, Dee J, Rudolph A, Orser C,

Fitzpatrick GM, Ho D. Wound-healing properties of trehalose-stabilized freeze-

dried outdated platelets. 2007 Apr;47(4):672–9. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17381626 PMID: 17381626

36. Chao FC, Kim BK, Houranieh AM, Liang FH, Konrad MW, Swisher SN, Tullis

JL. Infusible platelet membrane microvesicles: a potential transfusion substitute

for platelets. 1996 Jun;36(6):536–42. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/8669086 PMID: 8669086

37. Biró E, Akkerman JWN, Hoek FJ, Gorter G, Pronk LM, Sturk A, Nieuwland R.

The phospholipid composition and cholesterol content of platelet-derived

microparticles: a comparison with platelet membrane fractions. 2005

Dec;3(12):2754–63. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/16359513 PMID: 16359513

38. Blajchman MA. Substitutes for success. Nature Publishing Group; 1999 Jan

1;5(1):17–18. Available from: http://www.nature.com/doifinder/10.1038/4694

39. Modery-Pawlowski CL, Tian LL, Pan V, McCrae KR, Mitragotri S, Sen Gupta A.

Approaches to synthetic platelet analogs. Elsevier Ltd; 2013;34(2):526–541.

Available from: http://linkinghub.elsevier.com/retrieve/pii/S0142961212010976

249 40. Sen Gupta A. Bio-inspired Nanomedicine Strategies for Synthetic Blood

Substitutes. 2017;54(4):1–9. Available from:

http://ndt.oxfordjournals.org.ep.fjernadgang.kb.dk/content/28/4/803%5Cnhttp://cja

sn.asnjournals.org/cgi/doi/10.2215/CJN.00480115%5Cnhttps://www.jstage.jst.go.j

p/article/internalmedicine/54/11/54_54.3815/_article%5Cnhttp://onlinelibrary.wile

y.com/doi/10.10 PMID: 22747472

41. Hickman DA, Pawlowski CL, Sekhon UDS, Marks J, Gupta A Sen. Biomaterials

and Advanced Technologies for Hemostatic Management of Bleeding.

2017;1700859:1–40. PMID: 29164804

42. Ravikumar M, Modery CL, Wong TL, Sen Gupta A. Peptide-Decorated

Liposomes Promote Arrest and Aggregation of Activated Platelets under Flow on

Vascular Injury Relevant Protein Surfaces in Vitro. American Chemical Society;

2012 May 14;13(5):1495–1502. Available from:

http://pubs.acs.org/doi/10.1021/bm300192t

43. Modery-Pawlowski CL, Tian LL, Ravikumar M, Wong TL, Gupta A Sen. In vitro

and in vivo hemostatic capabilities of a functionally integrated platelet-mimetic

liposomal nanoconstruct. Elsevier Ltd; 2013;34(12):3031–3041. Available from:

http://dx.doi.org/10.1016/j.biomaterials.2012.12.045 PMID: 23357371

44. Sen Gupta A, Ravikumar M, Modery‐Pawlowski C. Synthetic platelets [Internet].

US; US9107845B2, 2012. Available from:

https://patents.google.com/patent/US9107845?oq=9107845%2C+2015

45. Anselmo AC, Modery-Pawlowski CL, Menegatti S, Kumar S, Vogus DR, Tian

LL, Chen M, Squires TM, Sen Gupta A, Mitragotri S. Platelet-like Nanoparticles:

250 Mimicking Shape, Flexibility, and Surface Biology of Platelets To Target Vascular

Injuries. American Chemical Society; 2014 Nov 25;8(11):11243–11253. Available

from: http://pubs.acs.org/doi/10.1021/nn503732m

46. Shukla M, Sekhon UDS, Betapudi V, Li W, Hickman DA, Pawlowski CL, Dyer

MR, Neal MD, McCrae KR, Sen Gupta A. In vitro characterization of

SynthoPlateTM (synthetic platelet) technology and its in vivo evaluation in severely

thrombocytopenic mice. 2017;15(2):375–387. PMID: 27925685

47. Rall JM, Cox JM, Songer AG, Cestero RF, Ross JD. Comparison of novel

hemostatic dressings with QuikClot combat gauze in a standardized swine model

of uncontrolled hemorrhage. 2013 Aug;75:S150–S156. Available from:

http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=01

586154-201308002-00009

48. Fülöp A, Turóczi Z, Garbaisz D, Harsányi L, Szijártó A. Experimental models of

hemorrhagic shock: A review. 2013;50(2):57–70. PMID: 23615606

49. Hess JR. Resuscitation of trauma-induced coagulopathy. 2013;664–7. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/24319249 PMID: 24319249

50. Cowley R. A total emergency medical system for the State of Maryland.

1975;24(7):37–45. Available from:

https://www.ncbi.nlm.nih.gov/pubmed/1142842 PMID: 1142842

51. Li Y, Boraschi D. Endotoxin contamination: a key element in the interpretation of

nanosafety studies. Future Medicine Ltd London, UK ; 2016 Feb 20;11(3):269–

287. Available from: https://www.futuremedicine.com/doi/10.2217/nnm.15.196

52. Szebeni J, Fontana J, Wassef N, Circulation PM-, 1999 undefined. Hemodynamic

251 changes induced by liposomes and liposome-encapsulated hemoglobin in pigs: a

model for pseudoallergic cardiopulmonary reactions to. Available from:

https://scholar.google.com/scholar?hl=en&as_sdt=0%2C36&q=Hemodynamic+ch

anges+induced+by+liposomes+and+liposome-

encapsulated+hemoglobin+in+pigs&btnG=

53. Szebeni J, Muggia F, Gabizon A, Barenholz Y. Activation of complement by

therapeutic liposomes and other lipid excipient-based therapeutic products:

Prediction and prevention. Elsevier B.V.; 2011;63(12):1020–1030. Available

from: http://linkinghub.elsevier.com/retrieve/pii/S0169409X11001943

54. Urbanics R, Bedőcs P, Szebeni J. Lessons learned from the porcine CARPA

model: constant and variable responses to different nanomedicines and

administration protocols. De Gruyter; 2015 Jan 1;7(3):219–231. Available from:

https://www.degruyter.com/view/j/ejnm.2015.7.issue-3/ejnm-2015-0011/ejnm-

2015-0011.xml

55. Immordino ML, Dosio F, Cattel L. Stealth liposomes: review of the basic science,

rationale, and clinical applications, existing and potential. Dove Press;

2006;1(3):297–315. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/17717971 PMID: 17717971

56. Hunt H, Stanworth S, Curry N, Woolley T, Cooper C, Ukoumunne O, Zhelev Z,

Hyde C. Thromboelastography (TEG) and rotational thromboelastometry

(ROTEM) for trauma-induced coagulopathy in adult trauma patients with

bleeding. 2015;(2). Available from:

http://doi.wiley.com/10.1002/14651858.CD010438.pub2 PMID: 25686465

252 57. Coats TJ, Brazil E, Heron M. The effects of commonly used resuscitation fluids on

whole blood coagulation. 2006 Jul 1;23(7):546–549. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/16794099 PMID: 16794099

58. CRASH-2 trial collaborators H, Shakur H, Roberts I, Bautista R, Caballero J,

Coats T, Dewan Y, El-Sayed H, Gogichaishvili T, Gupta S, Herrera J, Hunt B,

Iribhogbe P, Izurieta M, Khamis H, Komolafe E, Marrero M-A, Mejía-Mantilla J,

Miranda J, Morales C, Olaomi O, Olldashi F, Perel P, Peto R, Ramana P V, Ravi

RR, Yutthakasemsunt S. Effects of tranexamic acid on death, vascular occlusive

events, and blood transfusion in trauma patients with significant haemorrhage

(CRASH-2): a randomised, placebo-controlled trial. 2010 Jul 3;376(9734):23–32.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/20554319 PMID: 20554319

59. Brown JB, Neal MD, Guyette FX, Peitzman AB, Billiar TR, Zuckerbraun BS,

Sperry JL. Design of the Study of Tranexamic Acid during Air Medical

Prehospital Transport (STAAMP) Trial: Addressing the Knowledge Gaps. NIH

Public Access; 2015;19(1):79–86. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/25076119 PMID: 25076119

60. Cap AP, Perkins JG. Lyophilized Platelets: Challenges and Opportunities. 2011

May;70:S59–S60. Available from:

https://insights.ovid.com/crossref?an=00005373-201105001-00022

61. Apelseth TO, Cap AP, Spinella PC, Hervig T, Strandenes G. Cold stored platelets

in treatment of bleeding. Wiley/Blackwell (10.1111); 2017 Nov 1;12(4):488–495.

Available from: http://doi.wiley.com/10.1111/voxs.12380

62. Cap AP, Spinella PC. Just chill-it’s worth it! 2017 Dec;57(12):2817–2820.

253 Available from: http://www.ncbi.nlm.nih.gov/pubmed/29226373 PMID: 29226373

63. Snyder EL, Rinder HM. Platelet Storage — Time to Come in from the Cold? 2003

May 15;348(20):2032–2033. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/12748320 PMID: 12748320

64. Hoffman M, Monroe DM 3rd. A cell-based model of hemostasis. 2001;85(6):958–

965. PMID: 11434702

65. Elliott JT, Prestwich GD. Maleimide-Functionalized Lipids that Anchor

Polypeptides to Lipid Bilayers and Membranes. 2000 Nov;11(6):832–841.

Available from: http://pubs.acs.org/doi/abs/10.1021/bc000022a

66. Presolski SI, Hong VP, Finn MG. Copper-Catalyzed Azide-Alkyne Click

Chemistry for Bioconjugation. Hoboken, NJ, USA: John Wiley & Sons, Inc.;

2011. Available from: http://doi.wiley.com/10.1002/9780470559277.ch110148

67. Harmon P, Cabral-Lilly D, Reed RA, Maurio FP, Franklin JC, Janoff A. The

Release and Detection of Endotoxin from Liposomes. 1997 Aug 1;250(2):139–

146. Available from: http://www.ncbi.nlm.nih.gov/pubmed/9245430 PMID:

9245430

254 Chapter 7: Trauma-targeted Delivery of Tranexamic Acid for Augmenting

Hemostasis

7.1 Abstract

Trauma-associated non-compressible hemorrhage remains a leading cause of global mortality. Rapid stoppage of bleeding (hemostasis) is critical to save lives in trauma.

However, traumatic hemorrhage and coagulopathy often lead to a hyperfibrinolytic phenotype where hemostatic clots are incapable of stabilizing due to upregulated tPA and plasmin activity. Tranexamic Acid (TXA), a synthetic lysine analog that can inhibit tPA, plasminogen and plasmin, has emerged as a promising drug to mitigate fibrinolysis in hemorrhage. Tranexamic Acid is currently FDA-approved for treating heavy menstrual and post-partum bleeding, and has shown clinical promise for trauma treatment. However,

TXA at high or repeat dose causes thrombo-embolic complications, neuropathy, systemic coagulopathy etc., possibly due to off-target action. We hypothesized that encapsulation of

TXA within injury site-targeted nanoparticles can enable its delivery and clot-stabilizing action selectively at the trauma site, to improve hemostasis and survival while avoiding systemic off-target effects. To test this hypothesis, we used liposomes as a model delivery vehicle, decorated its surface with a fibrinogen-mimetic peptide (FMP) for anchorage to activated platelets within trauma-associated developing clot, and encapsulated TXA within them at clinically effective doses. These TXA-loaded trauma-targeted nanovesicles

(TTNVs), ~170 nm in diameter, were evaluated in vitro for their clot stabilizing effect in rat blood, and then in vivo for systemic safety, hemostatic efficacy and survival benefit in a liver hemorrhage injury model in rats. Treatment with TXA-loaded control (untargeted)

255 nanovesicles (TCNVs), free TXA or saline only (no drug) were used as comparison. Our studies show that in vitro, the TTNVs were capable of resisting lysis in tPA-spiked rat blood, at levels comparable to free TXA. In vivo, I.V.-administration of TTNVs did not cause any systemic negative effects in un-injured rats. Furthermore, in the rat liver hemorrhage model, post-injury administration of TTNVs showed substantial reduction in blood loss and improvement of short (1 hr) and long term (72 hr) survival. Histopathologic and immunofluorescence evaluation of excised tissue from euthanized rats confirmed systemic safety and trauma site-targeted action of the TTNVs. Overall, these studies establish the potential of the TTNV system for safe injury site-selective enhancement of hemostasis to improve survival outcomes in trauma.

Key words: Trauma, Hemorrhage, Hyperfibrinolysis, Tranexamic Acid, Targeted

Delivery, Rat Model

256 7.2 Introduction

Trauma-associated non-compressible hemorrhage, shock and coagulopathy remain among

the leading causes of mortality in both civilian and military populations, especially in the

age group 1-501–3. Recent clinical studies, e.g. the seminal Pragmatic, Randomized

Optimal Platelet and Plasma Ratios (PROPPR) trial and the Pre-Hospital Air Medical

Plasma (PAMPer) trial, have established that in such trauma hemorrhage scenarios,

transfusion of whole blood or blood components (e.g. 1:1:1 RBC: plasma: platelets) can

significantly improve mortality outcomes4–7. While the role of RBCs is to restore tissue oxygenation, it is the platelets and the coagulation components in plasma that work together to form a fibrin enmeshed hemostatic clot at the injury site that seals the vascular breach and stops the bleeding.

The cellular and molecular mechanisms of this hemostatic clot formation is shown in

Figure 7-1. In this process, as per the current consensus on the cell-based model of hemostasis8,9, vascular injury exposes tissue factor (TF) bearing cells (e.g. vascular smooth

muscle cells, adventitial fibroblasts etc.) to blood plasma such that the interaction of

coagulation factor FVII with TF and subsequent formation of the activated FVIIa-TF

complex leads to the generation of an initial amount of thrombin (FIIa). This thrombin then

activates other coagulation factors, e.g. FIX to FIXa and FX to FXa, as well as activates

blood platelets in the locality of the vascular damage. The vascular injury also exposes sub-

endothelial collagen, as well as induces secretion and deposition of von Willebrand Factor

(vWF) from the vascular endothelium and activated platelets10–15. Platelets can adhere to

these exposed proteins (e.g. to vWF by platelet GPIbα and to collagen by platelet GPIa/IIa)

257 and this adhesion further activates platelets which aggregate at the injury site principally

by bridging of platelet surface integrin GPIIb-IIIa by fibrinogen (Fg)

Figure 7-1. Cellular and molecular components of hemostasis regulation mediated by multifactorial mechanisms involving Tissue Factor (TF) bearing cells, adherent active platelets expressing negatively charged phosphatidylserine (PS), co-localized coagulation factors, thrombin generation, thrombin-induced conversion of fibrinogen to fibrin, crosslinking of fibrin to stabilize clot and degradation of crosslinked fibrin over time; in trauma-associated hyperfibrinolysis the fibrin degradation processes are upregulated while inhibitors for the processes are downregulated.

to form the primary hemostatic plug16,17. The surface of these aggregated active platelets

expose a high extent of negatively charged phospholipids like phosphatidylserine (PS)18,19

which allows for co-localization of the various coagulation factors, especially forming the

‘tenase’ complex (FVIIIa-FIXa-FX) in presence of co-factor calcium (Ca++) to generate

activated FXa. This in turn forms the ‘prothrombinase’ complex (FXa-Fva-FII) on the PS- rich platelet surface in presence of Ca++, leading to amplified generation of thrombin (FIIa).

The thrombin converts local Fg into fibrin, which via self-assembly as well as

transglutaminase FXIIIa-induced reaction forms a highly crosslinked biopolymeric

mesh20,21 to secure the platelet plug and other blood components (e.g. RBCs) to stop

bleeding (i.e. hemostasis).

258 Therefore, crosslinked fibrin formation and stabilization is central to the establishment of

hemostatic clots. As the injury environment transitions from the ‘hemostatic’ phase to

‘healing’ phase, natural well-regulated mechanisms contribute to clot (fibrin) degradation.

The most pertinent process is the secretion of tissue plasminogen activator (tPA) by

vascular endothelium which converts plasminogen (Plg) into plasmin, predominantly

amplified when Plg is bound to fibrin, and this plasmin degrades fibrin (e.g. into fibrin D- dimers)22,23. This process remains well-regulated by plasminogen activator inhibitor (PAI-

1), also secreted by the endothelium, which inhibits tPA locally24 to optimize the balancing

kinetics of fibrin formation (by thrombin) and fibrin degradation (by plasmin). Other

processes, e.g. the thrombin-thrombomodulin mediated activated Protein C (aPC) pathway

to deactivate FVa and FVIIIa also regulate thrombin (and hence fibrin) generation

kinetics25. However, in traumatic hemorrhage these mechanisms are severely dysregulated, often resulting in a unique coagulopathic scenario termed ‘hyperfibrinolysis’26–29. In this

hemorrhagic scenario, the lethal combination of plasma depletion and dilution,

hypothermia and hypoxic tissue acidosis has been reported to amplify the aPC pathway,

reduce fibrin deposition and crosslinking, upregulate tPA production and downregulate

PAI-1 activity. As a result, fibrinolytic activity offsets fibrin generation activity and the

clot becomes unstable, leading to impaired hemostasis and subsequent mortalities. Clinical

observation of this pathophysiology in trauma patients has prompted robust research into

the use of anti-fibrinolytic agents for clot stabilization and hemostatic benefit. Such agents

include, tranexamic acid or trans-4-(aminomethyl) cyclohexanecarboxylic acid

(abbreviated as TXA) and ε-aminocaproic acid (abbreviated as EACA), synthetic Lysine

analogs that inhibit tPA activation of Plg, Plg conversion to plasmin as well as action of

259 plasmin directly30–32. We have decided to utilize TXA over EACA due to it being 10 times

more potent in vitro and binding more strongly to both the weak and strong receptor sites

on the plasminogen molecule than EACA33. Additionally, a summary of the cost associated

with TXA compared to EACA show that the cost per course of treatment of TXA is slightly

cheaper than EACA ($3.70 vs $3.80)34. Several recent clinical trials have established a beneficial role of TXA in mitigating hemorrhagic shock in trauma35–37, and TXA is now

listed as an ‘essential medicine’ by the World Health Organization (WHO) for use in

hemostatic management of heavy menstrual and post-partum hemorrhage, traumatic

bleeding and surgery (pre- and post-operative). TXA is available in both oral and

intravenous form, and in traumatic hemorrhage the intravenous form is used

predominantly. While these studies have advanced the intravenous use of TXA in trauma,

several parallel reports have also emphasized potential systemic side-effect risks of TXA

use, especially at high dose or repeat dose, in the context of systemic off-target thrombotic,

thrombo-embolic and seizure risks38–40. Such findings have prompted recent research into

packaging of TXA within various delivery vehicles to explore site-localized controlled

release potential41–45. However, these research reports are majorly in the context of topical

application of TXA-loaded particles directly at the injury site. A formulation for

intravenous administration of TXA for injury site-targeted delivery and action in a setting

of traumatic hemorrhage has not been studied. In this framework, we hypothesized that

encapsulation of TXA within injury site-targeted nanoparticles can enable its delivery and

clot-stabilizing action selectively at the trauma site, to improve hemostasis and survival

while avoiding systemic off-target effects. Here we report our findings from testing this

260 hypothesis using injury site-targeted liposomal nanovesicles as the TXA carrier system in

a traumatic liver hemorrhage model in rats.

7.3 Materials and Methods

7.3.1 Materials

The lipid components, namely, 1,2-Distearoyl-sn-glycero-3-phosphocholine (DSPC), 1,2-

Distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-1000]

(DSPE-mPEG1000), and 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-

[azido(polyethylene glycol)-2000] (DSPE-PEG2000-Azide) were purchased from Avanti

Polar Lipids (Alabaster, AL, USA). Rhodamine B-dihexadecanoyl-sn-glycero-3-

phosphoethanolamine (DHPE-RhB) was purchased from Invitrogen (Carlsbad, CA, USA).

The peptide cyclo-{Pra}CNPRGD{Tyr(OEt)}RC (FMP) was custom-synthesized by

Genscript (Piscataway, NJ, USA). Fibrinogen mimetic peptide was conjugated to DSPE-

46 PEG2000-azide via copper-catalyzed alkyne-azide cycloaddition (CuCAAC) and was

purified by dialysis and characterized by MALDI-TOF mass spectrometry. Sterile normal

saline solution (0.9% NaCl) was purchased from Baxter (Deerfield, IL, USA). Cellulose

dialysis tubing (MWCO 2k and 3k), phosphate buffered saline (PBS),

(CaCl2), chloroform, methanol, and Dimethyl sulfoxide (DMSO) were purchased from

Fisher Scientific (Pittsburgh, PA, USA). L-Ascorbic acid 99% (C6H8O6) for the 2%

Ascorbic Acid solution was purchased from Sigma-Aldrich (Saint Louis, MO, USA). Tris

Hydrochloride (tris(hydroxymethyl)aminomethane hydrochloride) was purchased from

Promega Corporation (Madison, WI, USA). For rotational thromboelastometry (ROTEM)

analysis, all reagents were purchased from ROTEM (Munich, Germany). For in vivo

261 studies on pigs, SAS Sprague Dawley Rats were purchased from Charles River. Isoflurane,

buprenorphine, pentobarbital, lidocaine, ketamine, and xylazine were obtained from

Patterson Veterinary (Greely, CO, USA). The Beadbug Mircotube Homogenizer was

purchased from Benchmark Scientific (Edison, NJ, USA). The Rat D-dimer (D2D) ELISA

Kit was purchased from Biomatik (Wilmington, DE, USA). CD41 antibody FITC for

immunohistochemistry was purchased from Biorbyt (San Francisco, CA, USA).

7.3.2 TXA-loaded targeted nanovesicle (TTNV) development and characterization

For targeted delivery of TXA to hemorrhagic injury site, we utilized a liposomal vehicle

decorated on the surface by a cyclic RGD peptide, namely cyclo-CNPRGDY(OEt)RC that

has high affinity for the activated form of platelet surface integrin GPIIb-IIIa47,48. The

binding of this peptide to integrin GPIIb-IIIa mimic the binding of fibrinogen alpha chain

to that integrin, and thus we designate this peptide as fibrinogen mimetic peptide (FMP).

Compared to various integrin-binding ubiquitous RGD sequences (e.g. GRGDS), the FMP

sequence has high specificity towards the stimulated form of platelet integrin GPIIb-IIIa47,

and thus provides an ideal way to direct nanoparticle platforms to sites of active platelet aggregation and clot formation in hemorrhagic injury, as demonstrated in our previous

49,50 studies . This peptide was conjugated to the lipid DSPE-PEG2k-Azide by reacting with

Propargylglycine-terminated FMP via copper-catalyzed cycloaddition (i.e. ‘click’ chemistry) as shown in the schematic of Figure 7-5A. The resultant DSPE-PEG2K-FMP

conjugate was combined at 2.5 mol% with DSPC (46.5 mol %), cholesterol (45 mol%),

DSPE-mPEG(2k)(2.5 mol%), DSPE-mPEG(1k)(2.5 mol%), and DHPE-Rhodamine (1

mol%) to form FMP-decorated liposomal nanovesicles by film rehydration and extrusion

262 technique. To confirm the ability of the FMP-decorated nanovesicles to bind to clot-

associated activated platelets, the FMP-decorated nanovesicles were incubated with

platelet-rich plasma (PRP) in well plates in presence of thrombin to induce clot formation.

In these incubation samples, platelets were pre-stained with calcein and the plasma was spiked with AlexaFluor 647-labeled fibrinogen. Post 30 min incubation, the clots were harvested and analyzed by an Olympus FV 1000 confocal microscope, to detect platelet, fibrin and nanovesicle fluorescence in the clot.

The TXA was loaded within the liposomal vesicles during the lipid hydration step of the process, since TXA is highly soluble in water and allows for lipid hydration with an aqueous solution of TXA (167mg/ml). The TXA dose to be encapsulated was chosen as

93mg/kg in relevance to reported clinical doses37. The scaling of the dose from human to

rat was carried out per allometric scaling51. The dose for trauma associated hemorrhage in

humans is ~15mg/kg. This is converted to a rat dose using the allometric scaling factor of

6.2 to obtain the rat dose of 93mg/kg. Resultant TXA-loaded FMP-decorated trauma- targeted liposomal nanovesicles (TTNVs, schematic shown in Figure 7-5A) were characterized for size by dynamic light scattering (DLS). The encapsulation efficiency and release kinetics of TXA from were characterized by utilizing a colorimetric assay by conjugating TXA to ascorbic acid and analyzing with absorbance spectrometry (reaction schematic and calibration curve shown in Figure 7-2). FMP particles encapsulating TXA were synthesized, sealed in a 3k MWCO dialysis tube and placed in a beaker of 250 mL

5mM Tris-HCl 1mM CaCl2 buffer (pH 8.0) and incubated at 37°C for 3 hours. At each time

point (0 and every 10 minutes for 3 hours) 450uL of the buffer solution,

263

Figure 7-2. Ascorbic Acid-based reaction for assaying TXA, and representative calibration curve using this assay, based on which TXA release kinetics from nanovesicles was estimated.

which contained the released TXA, was removed from the 250mL reaction vessel and

replaced with 450uL of fresh 5mM Tris-HCl 1mM CaCl2 buffer (pH 8.0). After 3 hours, a

colorimetric assay was performed by conjugating TXA to ascorbic acid. 2mL of Ascorbic

acid stock solution (2% w/v in DMSO) was added to each collected sample and heated at

100°C for 45 minutes. The optical density of each sample was quantified by measuring the

absorbance at 390nm and the amount of TXA released at each time point and thus TXA

release kinetics was determined by utilizing an appropriate calibration curve that correlates

TXA-ascorbic acid absorbance intensity to TXA concentration.

7.3.3 In vitro analysis of the effect of free TXA and TTNVs on tPA-induced fibrinolysis

Rotational thromboelastometry (ROTEM) of rat whole blood was used to characterize and confirm the effect of TXA as well as TXA-loaded liposomal vesicles on improving clot stability in fibrinolytic conditions. ROTEM profiles provide effective characterization of

264

Figure 7-3. Representative instrument set-up, characteristic profile, actual instrument and relevant measured parameters in Rotational Thromboelastometry (ROTEM) analysis of whole blood (WB). For ROTEM analysis, whole blood sample is placed into a cuvette warmed at 37oC and a cylindrical pin is immersed in the blood. A gap of 1 mm remains between the pin and the cup wall, with blood in that space. The pin is rotated by a spring to the right and the left at 4.75o. When the blood is liquid, this movement is unrestricted. As the blood starts clotting, the developing clot increasingly restricts the rotation of the pin with rising clot firmness. Over time the clot can naturally lyse or undergo lysis due to action of extraneously added agents (e.g. tPA). The kinetics of this clot development, growth, stabilization and lysis processes are detected mechanically and processed digitally provide typical ROTEM profiles (TEMogram) from which various numerical parameters are calculated as shown above. In the specific studies described in the current manuscript, the parameter of interest were the lysis parameters (LI 30, LI 60 and ML) since the focus was to analyze the effect of TXA (free or nanovesicle-loaded) on clot stability.

clotting and clot formation time (CT and CFT), maximum clot firmness (MCF) and clot lysis parameters in whole blood52. The relevance of various parameters to clot characteristics is described in the schematic in Figure 7-3. As evident in comparing A1 and A2 in Figure 7-6A, a highly fibrinolytic effect could be created by adding 0.75ug/ml tPA to human blood in ROTEM assay, signified by the rapid onset of lysis (Lysis Index

265 after 30 minutes, LI30) and maximum lysis (ML) compared to that for healthy human

blood. Saline (control), or TXA at 4.9mg/ml as free TXA or loaded within liposomal

vesicles were added in these tPA-treated human blood assays and studied by ROTEM to

characterize clot properties, with the rationale that the TXA will inhibit the tPA effect and

thereby ‘rescue’ some of the clot stability (evidenced by improvement in MCF, LI30

profiles and delay in lysis). The TXA-loaded vesicles used in these studies did not have

FMP decorations since these are in vitro studies with blood in pooled conditions and thus an effect of ‘targeting’ the injury site is not relevant. The studies were mainly focused on establishing that TXA loaded within vesicles had comparable anti-fibrinolytic activity to free TXA, such that in subsequent in vivo studies the clot-targeted delivery of TXA-loaded vesicles is rationalized.

7.3.4 In vivo safety of FMP-decorated empty nanovesicles (TNVs) and TXA-loaded

TTNVs

In these studies, we designate FMP-decorated empty nanovesicles as ‘targeted nanovesicles’ (TNVs) and TXA-loaded versions of these targeted nanovesicles as TTNVs.

All in vivo rat studies were carried out in accordance with relevant guidelines and regulations approved by Case Western Reserve University Institutional Animal Care and

Use Committee (IACUC, Protocol Number 2017-0102). SAS Sprague Dawley Rats 200-

320g were acclimated to the laboratory space for 48 hours. Rats were anesthetized with isoflurane (1-5% to effect). Following induction of , ophthalmic ointment is applied, a rectal thermometer, pulse oximeter, blood pressure cuff and tail vein catheter are placed. For studying in vivo safety of TNVs as well as TTNVs, rats were administered with

266 nanovesicle dose of 8.8mg/kg. For evaluating TXA-loaded nanovesicles the TXA dose was kept at 93mg/kg. All injections were given via tail vein catheter at a 0.5ml volume. The vitals were monitored every 1 minute for the first 30 minutes, every 5 minutes for the next

30 minutes, then every 8 hours for 24 hours for anomalies that could be indicative of systemic effects. 24 hours after treatment administration, rats were euthanized with an

overdose of pentobarbital. Blood was drawn at baseline and pre-euthanasia endpoints to

analyze systemic effects. Post-euthanasia, clearance organs (heart, lungs, liver, spleen, kidney) were harvested for histology.

7.3.5 Development of rat liver hemorrhage traumatic injury model

TXA-loaded targeted nanovesicles (TTNVs) were evaluated in a rat liver injury model adapted from Morgan et al53. Expected 1hr mortality is 30-40% in untreated (saline only,

no drug) animals. Figure 7-8A shows a representative experimental set-up of the model

for these studies. Rats were prepared as previously described. The abdomen was shaved, a

subcutaneous buprenorphine administration and intramuscular injection of Lidocaine at the

incision site were administered. The abdomen was scrubbed, the peritoneal cavity opened

and the liver exposed. The peritoneal cavity was packed with pre-weighed absorbent gauze

and the liver was cut sharply 1.3 cm from the superior vena cava. The abdomen was

immediately closed, and treatment was administered 1 minute after injury. The vitals were

monitored every 1 minute for the first 30 minutes, every 5 minutes for the next 30 minutes,

then every 8 hours for 72 hours. One hour after injury, the gauze was removed from the

abdomen and weighed to determine blood loss, the abdomen was sutured closed and the

wound area was injected with lidocaine. 72 hours post-injury, rats were euthanized with an

267 overdose of pentobarbital. Blood was drawn at baseline and pre-euthanasia endpoints to analyze systemic effects. After euthanasia, the major organs (heart, lungs, liver, spleen, kidneys, brain, and blood) were harvested for histological and microscopic evaluation of off-target effects (if any).

7.3.6 Evaluation of TTNVs in rat liver hemorrhage model

Rats (total: n=48; saline: n=12, Free TXA n:=12, TXA-loaded control (unmodified) nanovesicles (TCNV): n=12, and TXA-loaded targeted (FMP-decorated) nanovesicles

(TTNV): n=12) were prepared and injured as previously described. TXA at a dose of

93mg/kg was administered 1 minute after injury via tail vein injection in rats at 0.5ml volume as free TXA, or loaded within nanoparticles (control or targeted). The vitals were monitored as previously described for anomalies that could be indicative of systemic effects. One hour after injury, blood loss was determined. 72 hours post-injury, rats were euthanized with an overdose of pentobarbital. Blood was drawn at baseline and pre- euthanasia endpoints to analyze systemic effects. After euthanasia, the major clearance organs (heart, lungs, liver, spleen, kidney) were harvested for histological and microscopic evaluation of off-target effects (if any). Organs were fixed in formalin, processed and slides were made for H&E and immunostaining. Sections of the liver traumatic injury site were deparaffinized and rehydrated by washing with xylene 2 times for 2 minutes, 1:1 xylene: ethanol for 3 minutes, ethanol 2 times for 3 minutes, 95% ethanol for 3 minutes, 70% ethanol for 3 minutes and finally 50% ethanol for 3 minutes. Antigen retrieval was performed by incubating in Tris-EDTA buffer in a water bath at 60 °C overnight. Slides were washed, blocked in 10% serum with 1%BSA in TBS for 2 hours, incubated with the

268 FITC-labeled (green fluorescent) CD41a antibody (for staining platelet integrin GPIIb),

washed, protected with a cover-slip and then imaged (using an Olympus FV 1000

fluorescence confocal microscope). Cell Nucleus (blue, DAPI), nanovesicle fluorescence

(red, Rhodamine B) and platelet fluorescence (green, FITC) were captured for the same

field of view.

7.3.7 Statistical analysis

Vitals data, blood loss data and ROTEM data were analyzed using one-way ANOVA with

a Tukey post-test. Survival data was analyzed using a Log-rank test. In all analyses,

significance was considered to be p < 0.05.

Figure 7-4. Representative confocal images showing red fluorescent FMP-decorated nanovesicles (designated in the article as trauma-targeted nanovesicle or TNV) can bind and localize with activated platelets (pseudocolored blue) within crosslinked fibrin-rich (pseudocolored green) clots formed from platelet-rich plasma (PRP) in presence of thrombin; these simulated studies confirm the ability of TNVs to incorporate within developing clots at the trauma site upon injury.

269 7.4 Results

7.4.1 Characterization of TTNVs and TXA release kinetics

Figure 7-4 shows representative confocal microscopy images of clots where FMP- decorated nanovesicles were incubated with PRP in presence of thrombin. As evident from the images, RhB-labeled nanovesicles (red) were capable of specifically binding to calcein- stained activated platelets (pseudocolored blue) within AlexaFluor 647 stained fibrin-rich

clots (pseudocolored green). Figure 7-5B shows a schematic for envisioned clot-localized delivery of TXA from injury site-targeted liposomal nanovesicles. Figure 7-5C shows representative results of DLS based analysis of size-distribution of TXA-loaded liposomal vesicles, demonstrating that the nanovesicle formulations can be consistently prepared at an average diameter of ~ 170 nm. Figure 7-5D shows representative results of encapsulation efficiency (EE) of TXA in such vesicles, demonstrating a consistent encapsulation efficiency of ~ 65%. These TXA-loaded nanovesicles were put in 3,000

MWCO dialysis tubing and incubated in a beaker of 250 mL 5mM Tris-HCl 1mM CaCl2

buffer (pH 8.0) at 37°C for 3 hours. Samples (450ul) were removed every 10 minutes,

replaced with fresh buffer, and underwent an ascorbic acid detection assay to determine

TXA concentration. As shown in Figure 7-5E, the release of TXA from these vesicles

followed a first order kinetic profile. It is important to note here that such release kinetics

are in agreement with profiles shown for other water-soluble drugs like streptokinase (SK)

from similar clot-targeted nanovesicles in fibrinolysis studies54. Furthermore, in those

studies the release of the drug was demonstrated to be a combination of diffusional release

as well as release due to degradation of DSPC-enriched vesicles by phospholipase enzymes

(e.g. sPLA2) that are reportedly upregulated at the site of clot formation (secreted from

270 stimulated platelets and macrophages)55,56. Since the lipid composition of the TXA- encapsulating nanovesicles incorporate the same mole% of DSPC, a similar release mechanism for the encapsulated payload can be rationalized for these systems.

Figure 7-5. [A] Schematic of the manufacturing process of TTNVs using DSPE-PEG2K-FMP, DSPC, DHPE- RhB lipids and incorporating TXA in the nanovesicle core; [B] Envisioned targeted delivery of TXA at the trauma site from TTNVs anchored to active platelets in the developing clot, to allow local inhibition of tPA, plasminogen and plasmin to site-specifically reduce hyperfibrinolysis; [C] Representative DLS characterization results from multiple batches of TTNVs manufactured for the studies show an approximate diameter of ~170 nm; [D] Representative encapsulation efficiency (EE) of TXA in TTNVs for multiple batches show an approximate encapsulation of ~ 60%; [E] Representative release profile of TXA from TTNVs at 37oC in PBS shows a first order release kinetics.

7.4.2 ROTEM analysis of the effect of TXA-loaded liposomes in human blood

In order to confirm that the TXA delivered via such nanovesicles is capable of resisting hyperfibrinolysis in rat blood clots (correlation to subsequent in vivo studies), ROTEM studies were carried out with rat whole blood spiked with tPA. Figure 7-6A shows representative

271

Figure 7-6. ROTEM analysis of clot characteristics of whole blood spiked with tPA and treated with saline (no TXA), free TXA or TXA loaded within nanovesicles; [A1] Normal ROTEM profile of rat blood; [A2] ROTEM profile of rat blood spiked with 0.75 mg/ml tPA showing rapid onset of lysis; Treatment with [A3] free TXA as well as [A4] nanovesicle-loaded TXA is capable of conserving clot characteristics and delay lysis onset possibly by inhibiting tPA, plasminogen and plasmin in the milieu; [B] shows ROTEM parameters from these studies, demonstrating that treatment with saline only (no TXA) is incapable of inhibiting lysis as the LI30 goes down (i.e. more lysis compared to MCF) and maximum lysis (ML) significantly increases compared to free TXA and TXA-loaded nanovesicle (TNV) groups; [C] shows Lysis 60 values, corroborating that treatment of tPA-spiked blood with free TXA or TNV allows for conservation of clot mechanics and reduce lysis to keep clot stability comparable to normal whole blood, while treatment with saline is incapable of providing such resistance to lysis.

272 ROTEM profiles of normal (A1), tPA-spiked (A2), tPA-spiked free TXA treated (A3) and tPA-spiked TXA-loaded nanovesicle treated (A4) rat whole blood. Spiking with 0.75 ug/ml of tPA resulted in a significant hyperfibrinolytic effect in rat whole blood clots, as evident from the ROTEM profile comparison between Figure 7-6A, A1 and A2, where MCF in

A2 was significantly reduced and clot lysis set in rapidly. Treatment of this tPA-spiked blood with free TXA enabled resistance to this ‘lytic’ state, as evident from Figure 7-6A3 where the MCF was demonstrably conserved and lysis onset was significantly delayed.

Treatment of tPA-spiked blood with TXA-loaded nanovesicles (TNV) allowed for similar resistance to rapid lysis, as evident from Figure 7-6A4. The table in Figure 7-6B shows

ROTEM parameters for the three treatment groups, namely saline (no TXA), free TXA and

TXA-loaded nanovesicles (TNVs), demonstrating that treatment with TNVs is capable of conserving MCF, resisting LI30 and reducing ML, at levels comparable to treatment with free TXA. Figure 7-6C shows quantitative analysis of lysis at 60 minutes (Lysis 60) in these studies, confirming that the treatment of tPA-spiked blood (i.e. induced hyperfibrinolytic state) with TXA-loaded nanovesicles can render substantial resistance to lysis at levels comparable to free TXA. Altogether these results suggest that if TXA is encapsulated within nanovesicles that allow targeted delivery selectively at a clot forming site post-injury, it can render localized inhibition of tPA (and plasmin) to resist clot lysis, while avoiding potential off-target effects associated with systemic delivery of free TXA.

Therefore, this strategy was further evaluated in vivo, in the rat liver hemorrhage model.

273 7.4.3 In vivo safety evaluation of various doses of FMP-decorated liposomes

To enable delivery of TXA-loaded nanovesicles specifically to the injury-associated clot

forming site, the nanovesicle surface was decorated with fibrinogen-mimetic peptides

(FMP) that bind to

Figure 7-7. In vivo safety evaluation with TNVs as well as TTNVs show that intravenous administration of these vesicles at dose of 8.8 mg/kg (TTNVs incorporate TXA dose of 93 mg/Kg) does not affect vitals over 24-hour period (A and B); Representative heart rate and SpO2 traces over the first 60 min period are shown in C and D. Total 6 animals (3 per group) were used for these studies.

the stimulated conformation of integrin GPIIb-IIIa on active platelets (a major component of clots). To ensure that these FMP-decorated nanovesicles themselves (TNVs) as well as loaded with TXA (TTNVs) do not pose any systemic negative effect in rats, safety studies were first conducted by dosing a pilot group of un-injured rats with these nanovesicles and monitoring vitals and reactions for 24 hours. Figure 7-7, A and B shows results from these studies, where administration of TNVs (empty trauma-targeted nanovesicles) or TTNVs

(TXA-loaded trauma-targeted nanovesicles) showed no negative effects on heart rate (HR),

274 oxygen saturation (SpO2) and body temperature over the 24-hour period following injection. Figure 7-7, C and D show representative traces of heart rate and SpO2 respectively over 60 min periods for rats administered with TNVs and TTNVs, showing stable vitals and such vitals were conserved over the 24-hour observation period.

Altogether, these results confirmed that intravenous administration of the TNVs and

TTNVs did not cause any drastic systemic effects in rats and all rats survived with stable vitals, until euthanized. Therefore, in subsequent studies TTNVs were administered to injured (liver hemorrhage) rats and evaluated for blood loss and survival, compared against administration of saline (no TXA), free TXA and TXA-loaded control (untargeted) nanovesicles (TCNVs).

Figure 7-8. [A] Schematic of rat liver injury hemorrhage model set-up (details described in Methods section); [B] Post-injury vitals of animals administered intravenously with saline, free TXA, TCNV or TTNV show that hemorrhagic injury causes higher heart rate while SpO2 stays somewhat consistent due to efficient use of ventilator; [C] Blood loss analysis post-injury and treatment administration shows that rats treated with TCNVs and TTNVs undergo substantially lower blood loss compared to rats treated with saline or free TXA and TTNV-treated rats exhibit lower blood loss than TCNV-treated rats; [D] 1-hr survival analysis post- injury and treatment administration shows that TTNV-treated rats exhibit substantially improved survival from the other treatment groups, and [E] this trend is conserved over 72 hours although mortality in all groups increase to some extent during that period. Total 48 animals (12 per group) were used for these studies.

275 7.4.4 Effect of clot-targeted delivery of TXA-loaded nanovesicles in rat liver injury model

The rat liver injury hemorrhage model was described in detail in the Methods section. In

this model, there is significant bleeding in the peritoneal cavity from the injured liver,

resulting in ~ 30 % fatality within 1 hour if no treatment is given (i.e. saline group in Figure

7-8D). It is important to note here that in the ‘saline’ group, saline was not transfused at

high volumes as a resuscitative transfusion strategy to treat hypovolemia from blood loss,

but administered only at volumes of 0.5ml that corresponds to the injection volume of other

treatments (free TXA and TXA-loaded untargeted and targeted nanovesicles). As shown in Figure 7-8B, post-injury during the first hour, rats without any TXA treatment (saline group) as well as with TXA treatment (as free drug as well as loaded within TCNVs and

TTNVs) showed higher heart rate compared to un-injured rats (Figure 7-7) since the hemorrhagic loss forces the heart to beat faster to compensate and maintain blood pressure.

The SpO2 in un-injured and injured rats remained comparable since they are all under

proper ventilation. More importantly, administration of free TXA was incapable of

reducing blood loss in the injured rats, whereas TXA delivery in TCNVs and TTNVs both

reduced blood loss substantially, with TTNV-treated rats showing reduced blood loss compared to TCNV-treated rats (Figure 7-8C). Furthermore, administration of free TXA as well as TXA in TCNVs were incapable of substantially improving 1-hr survival in rats

(survival level was comparable to ‘saline’ group) as shown in Figure 7-8D. In contrast, administration of TXA in TTNVs substantially improved survival during the 1-hr period.

This is very promising, considering that fact that most preventable mortalities resulting from traumatic hemorrhage occurs within 1-2 hours of injury (termed the ‘golden hour’).

Therefore, the results suggest that TXA nanoformulation in TTNVs can be a promising

276 technology for point-of-injury rapid administration in the hemorrhaging patient to improve

survival. Additionally, monitoring the survival over 72 hrs, TTNV-administered rats

continued to maintain improved survival compared to those administered with saline, free

TXA or TCNV (Figure 7-8E). Interestingly, administration of free TXA performed worse

than saline in maintaining survival over the 72-hr period, suggesting possible off-target

effects or reduced at the injury site so as to be incapable of augmenting (or

maintaining) clot stability.

Figure 7-9. Representative images from histopathologic analysis (H & E staining) of excised tissue sections from liver injured rats administered with various treatments show that TTNVs do not have any off-target effects on clearance organs and free TXA can have some off-target effects in lungs.

277 At the end of the 72-hr observation period, rats were euthanized and clearance organs as

well as injured liver tissue were excised for histologic and immunofluorescence evaluation.

Figure 7-9 shows representative images of histologic evaluation (H & E staining) of tissue

sections from the various organs excised from rats administered with the various treatment

groups. The images show that all animals underwent liver injury as noted by the hydropic

changes seen in the hepatocytes near the blood vessels. Of particular note is the prominent

pulmonary congestion observed in the rats treated with free TXA signified by the red blood

cells and exudate in the alveolar space. This could be indicative of pulmonary emboli. In

fact, the lungs in the rats treated with TTNVs showed the least presence of such effects

compared to all other treatment groups (saline, free TXA and TCNV). There were drastic

effects seen in the heart, spleen and kidney for the TCNV and TTNV treated animals,

compared to saline and free TXA treated animals. Overall, these results indicate that

TTNVs do not cause any major systemic effects in injured animals and that free TXA may

cause off-target unwanted effects systemically (e.g. in the lung).

Figure 7-10 shows representative immunofluorescence images of the liver hemorrhage injury site tissue, showing nucleated cells stained by DAPI (blue), platelets in the

hemostatic clot stained by FITC-antiCD41a (green) and TXA-loaded nanovesicles stained

by Rhodamine B (red). As evident from the figure, administration of saline (essentially no

treatment) was incapable of augmenting formation of an effective hemostatic clot (low

levels of green platelet fluorescence). Administration of free TXA improved the hemostatic

clot formation and stability, as demonstrated by the enhanced presence of green fluorescent

platelets in the clot. Administration of TXA in control nanovesicles (TCNV) showed partial

278 incorporation of the vesicles into hemostatic clots, as evident from the overlay of green platelets and red nanovesicles. However, the overlay suggests only low levels of incorporation, possibly because the vesicles did not have surface decorations with a clot site-targeting ligand (i.e. FMP motifs). In contrast, administration of TXA in targeted nanovesicles (TTNV) showed significant incorporation of vesicles (red) into the clot (green platelets) as evident from the overlay, indicating that the vesicle surface-decoration with

FMP motifs enables active anchorage of the vesicles within the clots by virtue of binding to active platelet surface integrin GPIIb-IIIa. Furthermore, the clot structure in TTNV- treated rat liver injury site seemed larger in size and more compact, as evident from the green fluorescence of the platelets. This is possibly due to the fact that localized release of

TXA from clot-incorporated TTNVs inhibits clot lysis and stabilizes the fibrin network of the clots which results in retention of the activated platelets within the clots, thereby enhancing the overall clot composition and stability. Additionally, by virtue of bridging with the active platelets in the developing clot, the TTNVs may amplify the overall aggregation of platelets in an injury site-selective manner. These results corroborate the findings from the in vitro ROTEM analysis shown previously in Figure 7-6 in the context of clot stability (i.e. conserved MCF and reduced lysis). The results further suggest that hemostatic clot stabilization by injury site-targeted delivery of TXA using TTNVs can result in improved short term and long-term survival in hemorrhagic trauma.

279

Figure 7-10. Representative confocal microscopy images of immunostained tissue sections from injured liver site of rats administered with the various treatment groups show that rats treated with saline, free TXA and TCNVs form sub-optimal clot structure in liver vasculature, while rats treated with TTNVs exhibit substantial co-localization of TTNVs with platelets in the hemostasized liver and the clot structure for TTNV-treated rats appeared more compact.

7.5 Discussion

Robust efforts are being directed at improving the survival outcome in traumatic hemorrhage and coagulopathy, especially in resource-limited pre-hospital an in-hospital

280 scenarios. This is evident in continued research and development in the areas of making

whole blood as well blood components more portable, storage stable and available outside

of large trauma centers and blood banks via improving pathogen reduction technologies,

reducing contamination risks, enhancing storage through reduced temperature processing

(e.g. cold-stored as well as cryo-preserved platelets and plasma), engineering donor-

independent biologic and synthetic blood cell mimics (e.g. stem cell derived as well as

synthetic RBCs and platelets) etc.57. Along with such strategic improvements in making

blood-derived products available and accessible for transfusion-based hemorrhage control, significant research continues to be directed at developing materials and technologies that allow augmentation of hemostasis in conjunction as well as absence of blood products, including recombinant coagulation factors, tourniquet systems, biomaterials based topical, intracavitary and injectable hemostat technologies, and pharmaceutical agents that may improve platelet production, fibrin generation and clot stability58. One such agent, TXA,

has generated significant clinical interest due to its ability to inhibit plasminogen activator,

plasminogen itself as well as plasmin, thus allowing a robust reduction in fibrinolytic

processes that are often associated in destabilizing hemostatic clots in trauma. Due to this

capability, an oral formulation of TXA (brand name Lysteda) was approved by the Food

and Drug Administration (FDA) in 2009 to treat heavy menstrual bleeding, and the drug

has been clinically evaluated more extensively in recent years for the hemostatic

management of bleeding complications in trauma, surgery and post-partum hemorrhage.

Tranexamic Acid is currently clinically available at 100 mg/ml injectable formulation

(Cyklokapron) for intravenous administration in the treatment of traumatic hemorrhage and

is currently listed in the ‘Essential Medicine’ list of World Health Organization (WHO).

281 To date, the CRASH 2 trial has been the largest clinical evaluation of TXA in trauma

(20,211 patients over 40 countries) and a detailed look at the trial outcomes show that TXA given within 3 hrs of injury reduced mortality but given after 3 hrs of injury increased bleeding and mortality. This can possibly be due to systemic off-target antiplasmin effects where body’s natural regulation of clot formation and clot dissolution may be affected by

TXA. Furthermore, the optimal dosing of TXA remains a topic of continued research, since

TXA use especially at high doses have been associated with complications like seizures, vascular thrombo-embolism etc. as well as mortality in patients with physiological fibrinolysis. Such effects can be rationalized to be possibly stemming from off-target non- specific action of TXA, especially when administered at high doses. Furthermore, such observations emphasize the continued research on optimizing TXA-based anti-fibrinolytic strategies for broad spectrum application towards hemorrhage control.

In this study, we investigated the potential of encapsulating and delivering TXA in a clot- targeted nanovesicle formulation, that ensures injury site-selective delivery and release of the drug for localized action. For this purpose, we utilized unilamellar liposomes as a model drug delivery system and decorated the liposome surface with peptide motifs that allow active anchorage to the stimulated conformation of integrin GPIIb-IIIa on activated platelets involved in clot formation at the injury site. Our in vitro studies demonstrate that the nanovesicles can effectively bind to platelets within clots and TXA released from the nanovesicles can effectively stabilize the clot and inhibit clot lysis. The TXA dose in these formulations were kept at 93mg/kg to maintain relevance to the clinical dosing. One should note here that loading of TXA is not limited to liposomal vesicles, since a few recent studies

282 have demonstrated the ability to load and deliver TXA in other particulate systems as stated

previously. For intravenously administrable nanoformulation of TXA, we chose liposomes

as a model vehicle since it is a clinically well-studied drug delivery platform59.

Furthermore, developing such injury site-targeted nanovesicle systems provide the

possibility of intravenously delivering multiple types of hemostasis augmenting agents in

combination with TXA as well as stand-alone, to maintain high availability of the agents at therapeutically active concentration at the target site, while minimizing off-target

distribution and effects. Our in vivo studies indicate that the injury site-targeted

nanovesicles are systemically safe at doses of 8.8mg/kg in rats, and delivery of TXA using

these targeted nanovesicles (TTNV) substantially reduced blood loss and improved

survival. We show here that all in vivo studies described here were carried out at pilot scale, where 6 animals were used for in vivo safety studies and 12 animals per treatment group (4 groups = 48 animals) were used for in vivo hemostatic efficacy studies, totaling

54 animals. Histology based evaluation of excised tissue from clearance organs of animals indicate that animals treated with TTNV after liver injury experienced less lung injury complications compared to control groups. Immunofluorescence-based evaluation of injury site tissue sections from the liver indicate that TXA-loaded TTNV could efficiently incorporate within clots at high extent and possibly contributed to robust clot formation and stability for improving hemostasis, reducing blood loss and enhancing survival in the rat liver hemorrhage model.

The current studies are indeed the first ‘proof-of-concept’ experiments and analyses of injury site-targeted intravenous delivery of TXA in a model nanoparticle system for post- injury administration in an established traumatic hemorrhage model in rats. Based on the

283 promise of the current studies, we anticipate future studies from our group and others

directed at evaluating such TXA delivery in the context of varying the ‘time of

administration’ of the TXA nanoformulation post-injury to explore whether such encapsulation and targeted delivery can improve the survival outcome in traumatic hemorrhage when administered at longer time-points post injury. It is also important to note here that the mechanistic elucidation of whether TXA can maintain its anti-fibrinolytic capabilities upon long time periods in vivo, can be efficiently studied by developing specific animal models that present persistent levels of trauma-associated hyperfibrinolysis. While in the current studies we have used an acute liver injury model in normal wild-type rats, future studies can be directed at developing genetically engineered murine and rodent models that inherently present hyperfibrinolytic phenotype. Associated studies should also be directed towards analyzing pharmacology/toxicology profiles and potential immune response attributes in utilizing such TXA-loaded nanoformulation intravenously, especially to find the maximum tolerated dose (MTD) level and the potential of repeat dosing in severely injured exsanguinating patients.

7.6 Conclusion

In this study, we present proof-of-concept studies of an injury site-targeted TXA nanoformulation to treat traumatic hemorrhage and improve hemostasis and survival. For this, we utilized active platelet integrin GPIIb-IIIa-anchoring liposomes as the injury site- targeted delivery system and encapsulated TXA in the aqueous core. The TXA could be released via first order kinetics to potentially allow local inhibitory action on tPA and plasmin to reduce fibrinolysis. Our in vitro ROTEM studies demonstrated the ability of

284 this TXA-nanoformulation in inhibiting lysis and maintaining clot stability. Our in vivo

studies demonstrated the systemic safety of the nanovesicles, as well as, the ability of the

TXA-nanoformulation in reducing blood loss and improving survival over 1 hr to 3-day

observation period in a rat liver injury traumatic hemorrhage model. Post-euthanasia evaluation of excised tissues indicated the enhanced ability of the targeted TXA nanovesicles (TTNV) to get incorporated within developing clots at the injury site and improve clot size and quality. Histopathological evaluation of the TTNVs did not demonstrate systemic off-target thrombotic risks in clearance organs (liver, lung, kidney

and spleen).

7.7 Acknowledgement

This research was supported in part by funding from the National Institutes of Health

(NIH), specifically National Heart Lung and Blood Institute (NHLBI Grant R01

HL121212, PI: A. Sen Gupta). D.A. Hickman was supported by AHA Pre-doctoral

Fellowship 17CPRE33670016, NIH grant T32 GM007250 and NIH grant TL1 TR000441.

The authors also acknowledge NIH National Center for Research Resources (Grant1 C06

RR12463-01, PI: Kutina) for supporting the laboratory facilities for Sen Gupta in the

Wickenden Building at Case Western Reserve University. The content of this publication

is solely the responsibility of the authors and does not represent the official views of the

NIH and other funding entities.

285 7.8 References 1. Krug EG, Sharma GK, Lozano R. The global burden of injuries. American Public

Health Association; 2000 Apr;90(4):523–6. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/10754963 PMID: 10754963

2. Cap AP, Spinella PC. Just chill-it’s worth it! 2017 Dec;57(12):2817–2820.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/29226373 PMID: 29226373

3. Cannon JW. Hemorrhagic Shock. Massachusetts Medical Society; 2018 Jan

25;378(4):370–379. Available from:

http://www.nejm.org/doi/10.1056/NEJMra1705649

4. Holcomb JB, Tilley BC, Baraniuk S, Fox EE, Wade CE, Podbielski JM, del Junco

DJ, Brasel KJ, Bulger EM, Callcut R a, Cohen MJ, Cotton B a, Fabian TC, Inaba

K, Kerby JD, Muskat P, O’Keeffe T, Rizoli S, Robinson BRH, Scalea TM,

Schreiber M a, Stein DM, Weinberg J a, Callum JL, Hess JR, Matijevic N, Miller

CN, Pittet J, Hoyt DB, Pearson GD, Leroux B, van Belle G. Transfusion of

Plasma, Platelets, and Red Blood Cells in a 1:1:1 vs a 1:1:2 Ratio and Mortality in

Patients With Severe Trauma. 2015;313:471. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/25647203 PMID: 25647203

5. Sperry JL, Guyette FX, Brown JB, Yazer MH, Triulzi DJ, Early-Young BJ, Adams

PW, Daley BJ, Miller RS, Harbrecht BG, Claridge JA, Phelan HA, Witham WR,

Putnam AT, Duane TM, Alarcon LH, Callaway CW, Zuckerbraun BS, Neal MD,

Rosengart MR, Forsythe RM, Billiar TR, Yealy DM, Peitzman AB, Zenati MS.

Prehospital Plasma during Air Medical Transport in Trauma Patients at Risk for

Hemorrhagic Shock. 2018;379(4):315–326. Available from:

http://www.nejm.org/doi/10.1056/NEJMoa1802345 PMID: 30044935

286 6. Johansson PI, Stensballe J, Oliveri R, Wade CE, Ostrowski SR, Holcomb JB. How

I treat patients with massive hemorrhage. 2014 Nov 13;124(20):3052–3058.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/25293771 PMID: 25293771

7. Gurney JM, Spinella PC. Blood transfusion management in the severely bleeding

military patient. 2018 Feb;31(2):1. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/29470190 PMID: 29470190

8. Hoffman M, Monroe DM 3rd. A cell-based model of hemostasis. 2001;85(6):958–

965. PMID: 11434702

9. Versteeg HH, Heemskerk JWM, Levi M, Reitsma PH. New Fundamentals in

Hemostasis. American Physiological Society Bethesda, MD; 2013 Jan;93(1):327–

358. Available from: http://www.physiology.org/doi/10.1152/physrev.00016.2011

10. Savage B, Almus-Jacobs F, Ruggeri ZM. Specific Synergy of Multiple Substrate–

Receptor Interactions in Platelet Thrombus Formation under Flow. Cell Press;

1998 Sep 4;94(5):657–666. Available from:

https://www.sciencedirect.com/science/article/pii/S0092867400816074

11. CHEN J, LÓPEZ JA. Interactions of Platelets with Subendothelium and

Endothelium. 2005 Jan;12(3):235–246. Available from:

http://doi.wiley.com/10.1080/10739680590925484

12. Farndale RW, Sixma JJ, Barnes MJ, De Groot PG. The role of collagen in

thrombosis and hemostasis. 2004;2(4):561–573. PMID: 15102010

13. Peyvandi F, Garagiola I, Baronciani L. Role of von Willebrand factor in the

haemostasis. SIMTI Servizi; 2011 May;9 Suppl 2(Suppl 2):s3-8. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/21839029 PMID: 21839029

287 14. De Ceunynck K, De Meyer SF, Vanhoorelbeke K. Unwinding the von Willebrand

factor strings puzzle. 2013 Jan 10;121(2):270–277. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/23093621 PMID: 23093621

15. Herr AB, Farndale RW. Structural insights into the interactions between platelet

receptors and fibrillar collagen. American Society for Biochemistry and Molecular

Biology; 2009 Jul 24;284(30):19781–5. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/19401461 PMID: 19401461

16. Cierniewski CS, Byzova T, Papierak M, Haas TA, Niewiarowska J, Zhang L,

Cieslak M, Plow EF. Peptide ligands can bind to distinct sites in integrin

alphaIIbbeta3 and elicit different functional responses. 1999 Jun

11;274(24):16923–32. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/10358039 PMID: 10358039

17. Bennett JS. Structure and function of the platelet integrin alphaIIbbeta3. American

Society for Clinical Investigation; 2005 Dec;115(12):3363–9. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/16322781 PMID: 16322781

18. Lentz B. Exposure of platelet membrane phosphatidylserine regulates blood

coagulation. 2003; Available from:

https://www.sciencedirect.com/science/article/pii/S0163782703000250

19. HEEMSKERK JWM, MATTHEIJ NJA, COSEMANS JMEM. Platelet-based

coagulation: different populations, different functions. 2013 Jan;11(1):2–16.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/23106920 PMID: 23106920

20. Ariëns R, Lai T, Weisel J, Greenberg C. Role of factor XIII in fibrin clot formation

and effects of genetic polymorphisms. 2002; Available from:

288 http://www.bloodjournal.org/content/100/3/743.short

21. MOSESSON MW. Fibrinogen and fibrin structure and functions. Blackwell

Science Inc; 2005 Aug;3(8):1894–1904. Available from:

http://doi.wiley.com/10.1111/j.1538-7836.2005.01365.x

22. Chapin J, Hajjar K. Fibrinolysis and the control of blood coagulation. 2015;

Available from:

https://www.sciencedirect.com/science/article/pii/S0268960X14000721

23. Kattula S, Byrnes JR, Wolberg AS. Fibrinogen and Fibrin in Hemostasis and

Thrombosis. 2017 Mar 22;37(3):e13–e21. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/28228446 PMID: 28228446

24. Gils A, Declerck PJ. Plasminogen activator inhibitor-1. 2004 Sep;11(17):2323–34.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/15379715 PMID: 15379715

25. Dahlbäck B, Villoutreix BO. The anticoagulant protein C pathway. 2005 Jun

13;579(15):3310–3316. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/15943976 PMID: 15943976

26. White NJ. Mechanisms of trauma-induced coagulopathy. 2013 Dec

1;2013(1):660–663. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/24319248 PMID: 24319248

27. Kolev K, Longstaff C. Bleeding related to disturbed fibrinolysis. Wiley-Blackwell;

2016;175(1):12–23. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/27477022 PMID: 27477022

28. Chang R, Cardenas JC, Wade CE, Holcomb JB. Advances in the understanding of

trauma-induced coagulopathy. American Society of Hematology;

289 2016;128(8):1043–9. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/27381903 PMID: 27381903

29. Davenport RA, Guerreiro M, Frith D, Rourke C, Platton S, Cohen M, Pearse R,

Thiemermann C, Brohi K. Activated Protein C Drives the Hyperfibrinolysis of

Acute Traumatic Coagulopathy. Europe PMC Funders; 2017;126(1):115–127.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/27841821 PMID: 27841821

30. McCormack PL. Tranexamic Acid. 2012 Mar 26;72(5):585–617. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/22397329 PMID: 22397329

31. Roberts I. Tranexamic acid in trauma: how should we use it? 2015 Jun;13:S195–

S199. Available from: http://www.ncbi.nlm.nih.gov/pubmed/26149023 PMID:

26149023

32. Pabinger I, Fries D, Schöchl H, Streif W, Toller W. Tranexamic acid for treatment

and prophylaxis of bleeding and hyperfibrinolysis. Springer; 2017 May;129(9–

10):303–316. Available from: http://www.ncbi.nlm.nih.gov/pubmed/28432428

PMID: 28432428

33. Pharmacia & Upjohn Company. CYKLOKAPRON ® tranexamic acid injection

Antifibrinolytic agent [Internet]. 2000. Available from:

https://www.accessdata.fda.gov/drugsatfda_docs/label/2011/019281s030lbl.pdf

34. Casati V, Guzzon D, Oppizzi M, Cossolini M, Torri G, Calori G, Alfieri O.

Hemostatic effects of aprotinin, tranexamic acid and epsilon-aminocaproic acid in

primary cardiac surgery. 1999 Dec;68(6):2252-6; discussion 2256-7. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/10617012 PMID: 10617012

35. Morrison JJ, Dubose JJ, Rasmussen TE, Midwinter MJ. Military Application of

290 Tranexamic Acid in Trauma Emergency Resuscitation (MATTERs) Study.

American Medical Association; 2012 Feb 1;147(2):113. Available from:

http://archsurg.jamanetwork.com/article.aspx?doi=10.1001/archsurg.2011.287

36. Brown JB, Neal MD, Guyette FX, Peitzman AB, Billiar TR, Zuckerbraun BS,

Sperry JL. Design of the Study of Tranexamic Acid during Air Medical

Prehospital Transport (STAAMP) Trial: Addressing the Knowledge Gaps. NIH

Public Access; 2015;19(1):79–86. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/25076119 PMID: 25076119

37. Roberts I, Shakur H, Coats T, Hunt B, Balogun E, Barnetson L, Cook L, Kawahara

T, Perel P, Prieto-Merino D, Ramos M, Cairns J, Guerriero C. The CRASH-2 trial:

a randomised controlled trial and economic evaluation of the effects of tranexamic

acid on death, vascular occlusive events and transfusion requirement in bleeding

trauma patients. Various; 2013 Mar;17(10):1–79. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/23477634 PMID: 23477634

38. Ker K, Edwards P, Perel P, Shakur H, Roberts I. Effect of tranexamic acid on

surgical bleeding: systematic review and cumulative meta-analysis. BMJ

Publishing Group; 2012 May 17;344:e3054. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/22611164 PMID: 22611164

39. Hunt BJ. The current place of tranexamic acid in the management of bleeding.

2015 Jan;70:50-e18. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/25440395 PMID: 25440395

40. Moore HB, Moore EE, Huebner BR, Stettler GR, Nunns GR, Einersen PM,

Silliman CC, Sauaia A. Tranexamic acid is associated with increased mortality in

291 patients with physiological fibrinolysis. 2017 Dec;220:438–443. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/28755903 PMID: 28755903

41. Li D, Li P, Zang J, Liu J. Enhanced hemostatic performance of tranexamic acid-

loaded chitosan/alginate composite microparticles. 2012;2012. PMID: 23193369

42. Bhattacharya SS, Banerjee S, Chowdhury P, Ghosh A, Hegde RR, Mondal R.

Tranexamic acid loaded gellan gum-based polymeric microbeads for controlled

release: In vitro and in vivo assessment. 2013 Dec 1;112:483–491. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/24183265 PMID: 24183265

43. van Raath MI, Weijer R, Nguyen GH, Choi B, de Kroon AI, Heger M. Tranexamic

Acid-Encapsulating Thermosensitive Liposomes for Site-Specific Pharmaco-Laser

Therapy of Port Wine Stains. NIH Public Access; 2016 Aug;12(8):1617–40.

Available from: http://www.ncbi.nlm.nih.gov/pubmed/29342342 PMID: 29342342

44. Aydemir Sezer U, Kocer Z, Aru B, Demirel GY, Gulmez M, Aktekin A, Ozkara S,

Sezer S. Combination of gelatin and tranexamic acid offers improved haemostasis

and safe use on internal hemorrhage control. Royal Society of Chemistry; 2016

Oct 5;6(97):95189–95198. Available from:

http://xlink.rsc.org/?DOI=C6RA16790J

45. Baylis JR, Chan KYT, Kastrup CJ. Halting hemorrhage with self-propelling

particles and local drug delivery. PMC Canada manuscript submission; 2016

May;141 Suppl 2(Suppl 2):S36-9. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/27207421 PMID: 27207421

46. Presolski SI, Hong VP, Finn MG. Copper-Catalyzed Azide-Alkyne Click

Chemistry for Bioconjugation. Hoboken, NJ, USA: John Wiley & Sons, Inc.;

292 2011. Available from: http://doi.wiley.com/10.1002/9780470559277.ch110148

47. Cheng S, Craig WS, Mullen D, Tschopp JF, Dixon D, Pierschbacher MD. Design

and synthesis of novel cyclic RGD-containing peptides as highly potent and

selective integrin alpha IIb beta 3 antagonists. 1994 Jan 7;37(1):1–8. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/7507165 PMID: 7507165

48. Huang G, Zhou Z, Srinivasan R, Penn MS, Kottke-Marchant K, Marchant RE,

Gupta AS. Affinity manipulation of surface-conjugated RGD peptide to modulate

binding of liposomes to activated platelets. 2008 Apr;29(11):1676–85. Available

from: http://www.ncbi.nlm.nih.gov/pubmed/18192005 PMID: 18192005

49. Shukla M, Sekhon UDS, Betapudi V, Li W, Hickman DA, Pawlowski CL, Dyer

MR, Neal MD, McCrae KR, Sen Gupta A. In vitro characterization of

SynthoPlateTM (synthetic platelet) technology and its in vivo evaluation in severely

thrombocytopenic mice. 2017;15(2):375–387. PMID: 27925685

50. Hickman DA, Pawlowski CL, Shevitz A, Luc NF, Kim A, Girish A, Marks J,

Ganjoo S, Huang S, Niedoba E, Sekhon UDS, Sun M, Dyer M, Neal MD.

Intravenous synthetic platelet ( SynthoPlate ) nanoconstructs reduce bleeding and

improve ‘ golden hour ’ survival in a porcine model of traumatic arterial

hemorrhage. Springer US; 2018;(February):1–14. Available from:

http://dx.doi.org/10.1038/s41598-018-21384-z

51. Nair AB, Jacob S. A simple practice guide for dose conversion between animals

and human. 2016;7(2):27–31. Available from:

/pmc/articles/PMC4804402/?report=abstract PMID: 27057123

52. Tynngård N, Lindahl TL, Ramström S. Assays of different aspects of haemostasis

293 - what do they measure? BioMed Central; 2015;13:8. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/25688179 PMID: 25688179

53. Morgan CE, Prakash VS, Vercammen JM, Pritts T, Kibbe MR. Development and

Validation of 4 Different Rat Models of Uncontrolled Hemorrhage.

2015;150(4):316–324. PMID: 25693160

54. Pawlowski, Christa L; Li, Wei; Sun, Michael; Ravichandran, Kavya; Hickman,

DaShawn; Kos, Clarissa; Kaur, Gurbani; Sen Gupta A. Platelet microparticle-

inspired clot-responsive nanomedicine for targeted fibrinolysis. Elsevier Ltd;

2017;128:94–108. PMID: 28314136

55. Burke JE, Dennis EA. Phospholipase A2 structure/function, mechanism, and

signaling. 2009 Apr;50 Suppl:S237-42. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/19011112 PMID: 19011112

56. Zhu G, Mock JN, Aljuffali I, Cummings BS, Arnold RD. Secretory phospholipase

A responsive liposomes. 2011 Aug;100(8):3146–59. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/21455978₂ PMID: 21455978

57. Sen Gupta A. Bio-inspired Nanomedicine Strategies for Synthetic Blood

Substitutes. 2017;54(4):1–9. Available from:

http://ndt.oxfordjournals.org.ep.fjernadgang.kb.dk/content/28/4/803%5Cnhttp://cja

sn.asnjournals.org/cgi/doi/10.2215/CJN.00480115%5Cnhttps://www.jstage.jst.go.j

p/article/internalmedicine/54/11/54_54.3815/_article%5Cnhttp://onlinelibrary.wile

y.com/doi/10.10 PMID: 22747472

58. Hickman DA, Pawlowski CL, Sekhon UDS, Marks J, Gupta A Sen. Biomaterials

and Advanced Technologies for Hemostatic Management of Bleeding.

294 2017;1700859:1–40. PMID: 29164804

59. Bozzuto G, Molinari A. Liposomes as nanomedical devices. Dove Press;

2015;10:975–99. Available from: http://www.ncbi.nlm.nih.gov/pubmed/25678787

PMID: 25678787

295 Chapter 8: Design and Administration of Nanoparticles to Prevent Pseudoallergic

Responses in Intravenous Hemostatic Management of Bleeding

8.1 Introduction

Traumatic injury is the leading cause of mortality globally for individuals 5-44 years of

age and accounts for 10% of all deaths in the US1. Blood loss is the primary cause of death

at acute time points post injury2,3. Research has demonstrated that immediately halting blood loss is one of the most effective means of minimizing mortality associated with severe trauma. The current standard of care for treating internal bleeding, blood products, has a number of limitations such as low in-hospital and no pre-hospital availability, short shelf life, and risk of bacterial contamination. Our lab has created an I.V.-administrable platelet-inspired nanomedicine, that can mimic and amplify body’s natural hemostatic mechanisms specifically at the bleeding site while overcoming these limitations and maintaining systemic safety.

The hemostatic capability of these platelet-inspired nanomedicines in decreasing blood loss and increasing survival have previously been described in a murine and porcine model of acute hemorrhagic shock and in correcting tail-bleeding time in thrombocytopenic mice4–

6. When these nanomedicines were first used in a porcine femoral artery injury model, a

model that better represents the human cardiovascular system, it was important to explore

methods to eliminate hypersensitivity reactions such as complement activation-related

pseudoallergy (CARPA), which pigs are highly susceptible and has been found to be

present in 23% of humans receiving nanomedicine7. CARPA is a hypersensitivity reaction

296 to nanoparticles and causes over-activation of the complement system and can lead to

symptoms such as arrhythmias, systemic hypotension, skin flushing, bronchospasm,

increased adenosine and is responsible for over 20,000 deaths per year in the US7,8.

Hypersensitivity reactions to nanoparticles have been linked to the over activation of the

complement system9. The complement system acts to distinguish between healthy cells, compromised cells, and foreign bodies and alters its response accordingly10. Often,

nanoparticles are designed, unintentionally, to have the same size and charge as pathogens

that the complement system is primed to combat7. Infusion of billions of these

nanoparticles can lead to a massive response that is not only detrimental to the

nanoparticles but also to the patient. Currently, this pseudoallergy is mitigated clinically

with FDA approved nanoparticle formulations such as Doxil, by infusing the treatment

slowly over many hours thereby avoiding high plasma concentrations of particles11.

Unfortunately, this is not a practical method of nanomedicine administration for someone

with fatal internal hemorrhage.

Alternatively, we could seek to modify the properties of our nanomedicine to eliminate

CARPA. Various characteristics of nanoparticles known to impact complement system

activation, include administration protocol, size, PEGylation, and zeta potential12,13.

Previously, I have contributed to research published by Onwukwe et al., where we were

able to mitigate the complement response by engineering a particle with a more neutral

zeta potential14. In the following studies, we observed the impact of PEGylation and administration protocol on CARPA response.

297 8.2 Materials and Methods

8.2.1 Materials

The lipid components, namely, 1,2-Distearoyl-sn-glycero-3-phosphocholine (DSPC), 1,2-

Distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-1000]

(DSPE-mPEG1000), 1,2-Distearoyl-sn-glycero-3-phosphocholine (DSPC), 1,2-Distearoyl- sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol)-2000] (DSPE- mPEG2000), 1,2-Distearoyl-sn-glycero-3-phosphoethanolamine-N-

[maleimide(polyethylene glycol)-2000] (DSPE-PEG2000-Mal) and 1,2-distearoyl-sn-

glycero-3-phosphoethanolamine-N-[azido(polyethylene glycol)-2000] (DSPE-PEG2000-

Azide) were purchased from Avanti Polar Lipids (Alabaster, AL, USA). Rhodamine B- dihexadecanoyl-sn-glycero-3-phosphoethanolamine (DHPE-RhB) was purchased from

Invitrogen (Carlsbad, CA, USA). The peptides CTRYLRIHPQSWVHQI (VBP), C[GPO]7

(CBP) and cyclo-{Pra}CNPRGD{Tyr(OEt)}RC (FMP) were custom-synthesized by

Genscript (Piscataway, NJ, USA). VBP and CBP were conjugated to DSPE-PEG2000-Mal

15 via thiol-maleimide coupling and FMP was conjugated to DSPE-PEG2000-azide via

copper-catalyzed alkyne-azide cycloaddition (CuCAAC)16, conjugates were purified by

dialysis and characterized by MALDI-TOF mass spectrometry. Sterile normal saline

solution (0.9% NaCl) was purchased from Baxter (Deerfield, IL, USA). Cholesterol, rat

tail type I collagen, copper(II) sulfate (CuSO4), Tris(3-

hydroxypropyltriazolylmethyl)amine (THPTA), sodium ascorbate, decaethylene glycol

monododecyl (C12E10) and the Mammalian Cell Lysis Kit were from Sigma-Aldrich (Saint

Louis, MO, USA). Ethylenediamine Tetraacetic Acid (EDTA), cellulose dialysis tubing

(MWCO 2k and 3.5k), phosphate buffered saline (PBS), chloroform, methanol, For

298 rotational thromboelastometry (ROTEM) analysis, all reagents were purchased from

ROTEM (Munich, Germany). For in vivo studies on pigs, Yorkshire Farm Pigs were

purchased from Shoup Investments Ltd. Telazol, isoflurane and pentobarbital were

obtained from Patterson Veterinary (Greely, CO, USA). The Guinea Pig Complement C3

ELISA kit and Human Complement C3a des Arg ELISA kit were from Abcam

(Cambridge, MA, USA).

Table 8-1. Characteristics of SynthoPlate PEG variants

8.2.2 Manufacture of SynthoPlateTM 1.0

SynthoPlateTM 1.0 was manufactured using ‘film rehydration and extrusion’ technique as

17,18 described previously . Briefly, DSPC, cholesterol, DSPE-PEG2000-VBP, DSPE-

PEG2000-CBP, and DSPE-PEG2000-FMP were homogeneously mixed at 0.475, 0.45,

0.0125, 0.0125, and 0.025 and mole fractions, respectively, in 1:1 chloroform:methanol.

Solvent was removed via rotary evaporation, and the thin lipid film was rehydrated with

normal saline solution (0.9% NaCl) at a concentration of 1 x 105 moles lipid per mL. This

lipid suspension was subjected to 10 freeze/thaw cycles and subsequent extrusion through

200nm pore diameter polycarbonate membrane using a pneumatic extruder (Northern

Lipids, Burnaby, Canada) to create heteromultivalently decorated SynthoPlateTM vesicles.

299 Dynamic light scattering (DLS) and electron microscopy characterization indicated fresh- made vesicles were ~200nm in diameter.

8.2.3 Manufacture of SynthoPlateTM 1.5

SynthoPlateTM 1.5 was manufactured using ‘film rehydration and extrusion’ technique as described above except DSPC, cholesterol, DSPE-PEG2000-VBP, DSPE-PEG2000-CBP,

DSPE-PEG2000-FMP and DSPE-mPEG2000 were homogeneously mixed at 0.475, 0.45,

0.0125, 0.0125, 0.025 and 0.025 mole fractions, respectively, in 1:1 composition of

chloroform : methanol.

8.2.4 Manufacture of SynthoPlateTM 2.0

SynthoPlateTM 1.5 was manufactured using ‘film rehydration and extrusion’ technique as described above except DSPC, cholesterol, DSPE-PEG2000-VBP, DSPE-PEG2000-CBP,

DSPE-PEG2000-FMP and DSPE-mPEG1000 were homogeneously mixed at 0.475, 0.45,

0.0125, 0.0125, 0.025 and 0.025 mole fractions, respectively, in 1:1 chloroform:methanol.

8.2.5 In vivo Safety Studies in Pigs

All in vivo porcine studies were carried out in accordance with relevant guidelines and

regulations approved by Case Western Reserve University Institutional Animal Care and

Use Committee (IACUC, Protocol Number 2015-0135). Yorkshire Farm Pigs 25-36kg

were acclimated to the laboratory space for 48 hours. Pigs were fasted for 12 hours before

experiments but had free access to water. Pigs were sedated with Telazol (6-8mg/kg)

injected intramuscularly, intubated with an endotracheal tube and anesthetized with

300 isoflurane (1-5% to effect). A CO2 sensor was placed at the end of the endotracheal tube.

Mechanical ventilation was provided to keep end-tidal CO2 and respiration rate initially at

normal values. Electrocardiogram electrodes were placed on the pig’s limbs. A pulse-

oximeter probe was placed on the pig’s mouth (cheek or tongue). An esophageal

temperature probe was placed to measure core temperature, which was maintained between

36oC-38oC with a water-filled warming blanket. An angiocatheter was placed in the carotid

artery to acquire invasive blood pressure and withdraw blood samples for ex vivo testing.

Another angiocatheter was placed in the internal jugular vein to deliver saline or particle

treatments. The 500ml dose of SynthoPlateTM to be administered in the pigs was 1X1010

particles/ml, while the 50ml dose of SynthoPlateTM was 1X1011 particles/ml followed by a

450ml saline infusion. Invasive arterial blood pressure, CO2, SpO2, temperature, and heart

rate were recorded every 30 seconds for the first 10 minutes, every minute for the next 20

minutes, and every 5 minutes thereafter for a total of 60 minutes. Arterial blood was drawn

via the carotid artery angiocatheter at baseline, 30 minutes, and 60 minutes. At the end of

experiments, pigs were euthanized with an I.V. overdose of pentobarbital (0.22ml/kg).

In order to evaluate the CARPA response of SynthoPlateTM PEG variants within the

‘golden hour’ time-frame, without any confounding effects from injury, treatments were administered (at the 500ml volume) in a pilot group of non-injured animals (total n=5;

SynthoPlateTM 1.0: n=2, SynthoPlateTM 1.5: n=2, SynthoPlateTM 2.0: n=1), and the animals

were observed for 1 hour. To evaluate the CARPA response of SynthoPlateTM volume

adminstration within the ‘golden hour’ time-frame, without any confounding effects from

injury, SynthoPlate 2.0 was administered at the 500ml volume or the 50ml volume

301 followed by 450ml saline in a pilot group of non-injured animals (total n=3; 50ml: n=2,

500ml: n=1), and the animals were observed for 1 hour. The surgeons were blinded to the

administered treatments, and blinding was ensured by a researcher who assigned a

randomly generated code to each treatment. The effects of the administered treatments on

vitals, skin and blood chemistry were examined. The risk for CARPA reactions was also

assessed ex vivo. For this, specific complement activation marker (C3 to C3a) was

measured in the blood drawn from the pig at baseline as well as 30 and 60 minutes post

administration treatment. In complement activation, C3 is cleaved into C3a and C3b

leading to an increase in plasma C3a levels, and thus the plasma C3a concentration can

indicate such risks. For these studies, platelet poor plasma was isolated from pig blood

samples by centrifuging for 25 minutes at 2500 x g and the measurement of C3a was carried

out using Human Complement C3a des Arg ELISA kit respectively (Abcam, Cambridge,

MA, USA).

302

Figure 8-1 Schematic representation of pig femoral artery hemorrhage model setup. A CO2 sensor was placed at the end of the endotracheal tube and mechanical ventilation was provided, EKG electrodes were placed on the pig’s limbs, a pulse-oximeter probe was placed on the pig’s mouth, an esophageal temperature probe was placed to measure core temperature, an angiocatheter was placed in the carotid artery to acquire invasive blood pressure and also withdraw blood samples for ex vivo analysis, an angiocatheter was placed in the internal jugular vein to deliver saline (or nanoparticle treatments) via an infusion pump.

8.2.6 Femoral Artery Bleeding Model in Pigs

Figure 8-1 shows a representative experimental set-up of the model for these studies. Pigs

(total n=9; 50ml: n=6, 500ml: n=3) were prepared as described previously. The skin at the

inguinal area of the thigh was incised ~10cm, the femoral artery was exposed and injured

(near transection) using a 3.5mm aortic punch (Scanlan International, St Paul, Minnesota,

USA). SynthoPlate 2.0 at the 500ml volume or the 50ml volume followed by 450ml saline

were administered 1 minute after injury via the jugular at a rate of 60ml/min. Blood loss

from femoral artery injury was measured by carefully suctioning the shed blood without

303 disturbing the injury site, into a container that was weighed every 30 seconds for the first

10 minutes, every minute for the next 20 minutes, and every 5 minutes thereafter until

bleeding stopped due to death or hemostasis. Blood mass was correlated to blood volume

by approximating blood density to be 1000 kg/m3. With saline administration, which is a

standard volume resuscitation strategy for treating hemorrhage in pre-hospital scenarios, this injury resulted in an average of 31.5 mL/kg blood loss in 24 minutes, and showed only

25% survival within the first 60 min (i.e. within the ‘golden hour’ period). At the end of

experiments, pigs were euthanized with an I.V. overdose of pentobarbital (0.22ml/kg).

8.2.7 Hemostatic Efficacy Evaluation in Femoral Artery Bleeding Model

The femoral artery bleeding model was performed as described above. Induction of the

injury was designated as time-point zero (0). SynthoPlate 2.0 at the 500ml volume or the

50ml volume followed by 450ml saline treatments were administered 1 minute after injury

via the jugular, (at the concentrations previously described) at 60 ml per minute. Blood loss

was measured as described previously. Arterial blood was drawn via the carotid artery

angiocatheter at baseline, and 15, 30, 60 and 120 minutes, to run ex vivo analysis. At the

end of experiments, pigs were euthanized.

8.2.8 Statistical Analysis

Statistical analysis of blood loss data and complement data was done using two-way

ANOVA with a Bonferroni post-test. Vitals data were analyzed using one-way ANOVA

with a Tukey post-test. Survival data was analyzed using a Log-rank test. In all analyses,

significance was considered to be p < 0.05.

304 8.3 Results

8.3.1 Synthesis of SynthoPlate PEG variants

PEGylation of a nanoparticles has been reported to effect the complement activation- related pseudoallergy response19. This led us to develop and evaluate three iterations of

SynthoPlate (1.0, 1.5, 2.0) with different PEGylation compositions (Table 8-1).

SynthoPlate 1.0 was designed with 5 mol percent 2,000 MW PEG that was all conjugated between DSPE and functional peptide (VBP, CBP or FMP). The 2,000 MW PEG mol percentage was increased to 7.5 for SynthoPlate 1.5 by adding an additional 2.5 mol percent of DSPE conjugated (peptide free) PEG. In the SynthoPlate 2.0 formulation, the additional

2.5 mol percent of DSPE conjugated (peptide free) PEG was replaced with 1,000 MW PEG to help reduce steric interference of the PEG chains. The SynthoPlate PEG variants had a hydrodynamic diameter that was roughly 200nm and zeta potential of about -6mV (Table

8-1). Varying PEG content did not significantly affect other characteristics of these nanoparticles.

8.3.2 Evaluation of SynthoPlate PEG variants

The SynthoPlate PEG variants were evaluated in our in vivo porcine safety model to assess their effect on complement activation-related pseudoallergy (total n=5; SynthoPlateTM 1.0:

n=2, SynthoPlateTM 1.5: n=2, SynthoPlateTM 2.0: n=1). As previously mentioned CARPA

can lead to changes in heart rate, blood pressure and breathing, so these parameters were

assessed by monitoring heart rate, mean arterial pressure and CO2 respectively. After

injection in non-injured pigs, SynthoPlate 1.0 caused a drastic drop in pigs’ heart rate, mean

arterial pressure and CO2 (Figure 8-2A, B, and C respectively). This ultimately led to the

305 pigs administered SynthoPlate 1.0 dying within 30 minutes of particle administration

(Figure 8-2D). This was indicative of a drug induced hypersensitivity reaction. Pigs

administered, SynthoPlate 1.5 had spikes in their heart rate, mean arterial pressure, and

CO2 within 10 minutes of particles administration but stabilized within 40 minutes without

intervention (Figure 8-2A, B, and C respectively). All animals treated with SynthoPlate

1.5 survived 60 minutes (Figure 8-2D). These observations are indicative of a mild drug hypersensitivity reaction. The Pig administered SynthoPlate 2.0 had stable heart rate and

CO2 and a slight spike in mean arterial pressure at 10 minutes that stabilized within 10

minutes without intervention (Figure 8-2A, C, and B respectively). This animal survived

the 60 minutes experiment time period (Figure 8-2D). These observations are likely an

indication of no reaction to the drug. The increase in mean arterial pressure alone is likely

a compensatory mechanism due to the increase in total blood volume from the 500ml

treatment.

306

Figure 8-2 Vital analysis in safety pigs administered SynthoPlate PEG variants show that [A] treatment with SynthoPlate 2.0 had no effect on heart rate while SynthoPlate 1.5 did show minimal effect and SynthoPlate 1.0 demonstrated more pronounced effects. There were similar results seen for [B] Mean Arterial Pressure and [C] CO2. [D] Animals administered SynthoPlate 2.0 and 1.5 had 100% survival rates while those administered SynthoPlate 1.0 had a 0% survival rate.

CARPA can also lead to a skin response. For this reason, the skin was monitored for

changes after SynthoPlate administration. SynthoPlate 1.0 administration lead to a

generalized erythematous rash (Figure 8-3A) that appeared roughly 1 minute after injection and resolved at the end of the injection (5 minutes). Administration of

SynthoPlate 1.5, lead to a marbled erythematous rash (Figure 8-3B) that also appeared roughly 1 minute after injection and resolved at 5 minutes. A milder skin reaction was observed with SynthoPlate 2.0 (Figure 8-3C). We observed generalized small erythematous patches with the same time course as SynthoPlate 1.0 and 1.5. These results support the vitals data.

307

Figure 8-3 Images of pig skin responses after SynthoPlate PEG variant administration. [A] A Generalized erythematous rash is seen after SynthoPlate 1.0 administration. [B] After SynthoPlate 1.5 administration a marbled erythematous rash is observed. [C] A generalized rash with small erythematous patches is observed after 500ml administration of SynthoPlate 2.0. [D] Little to no reaction is observed after 50ml bolus administration of SynthoPlate 2.0.

One of the main risk indicators for drastic activation of complement system and deleterious reactions like CARPA is a more than 4 -fold increase in complement protein20–22. We measured the product of C3 cleavage when the complement system is activated, complement C3a. We did not measure a significant increase in complement system activation in our SynthoPlate PEG variants (Figure 8-4A). There was a slight increase of

308 C3a plasma concentration in animals treated with our SynthoPlate 1.0 variant (Figure

8-4A). Animals treated with SynthoPlate 1.5 had relatively stable C3a plasma

concentrations while animals treated while animals treated with our SynthoPlate 2.0 variant

had a decrease in C3a plasma concentration (Figure 8-4A).

Figure 8-4 Analysis of C3a plasma concentration in blood drawn from pigs administered with [A] SynthoPlate PEG variants showed no significant alterations and [B] 50ml or 500ml SynthoPlate showed no significant alterations.

8.3.3 Evaluation of SynthoPlate Administration Protocol Safety

SynthoPlate 2.0 elicited the mildest reaction in pigs after administration and was therefore chosen to evaluate the effect of administration volume in our porcine safety model. Pigs were evaluated with an infusion speed of 60ml/min with either 50ml (bolus injection- <1

min) followed by 450ml saline infusion or 500ml (infusion- 8.3min) (total n=3; 50ml: n=2,

500ml: n=1). SynthoPlate 2.0. Pigs treated with bolus injections had peaks in heart rate and

CO2 at 20min that began to stabilize shortly after, while their mean arterial pressure was relatively stable (Figure 8-5A, C and B respectively). Pigs infused with SynthoPlate 2.0 had stable heart rate and CO2 but a peak in mean arterial pressure at 10 minutes, likely due

to an increase in intravascular volume (Figure 8-5A, C and B respectively). Both

treatments had a 100% 60-minute survival rate (Figure 8-5D). Pigs treated with bolus

309 injection had little to no skin changes (Figure 8-3D) while pigs infused with SynthoPlate

2.0 experienced generalized small erythematous patches (Figure 8-3C).

Figure 8-5 Vital analysis in safety pigs administered SynthoPlate 2.0 at 50ml (bolus) or 500ml (infusion) show that [A] Heart rate remains stable after infusion while there is a spike at 20min with bolus injection [B] Mean Arterial Pressure remains relatively stable with bolus injection but has a spike with infusion and [C] CO2 remains stable with infusion but has slight variations with bolus injection. [D] All pigs administered SynthoPlate 2.0 via bolus or infusion survived the experiment time (60 minutes).

8.3.4 Evaluation of SynthoPlate Administration Protocol Efficacy

These administration protocols where then evaluated in our porcine femoral artery injury

model (Figure 8-1). SynthoPlate 2.0 was injected by bolus or infusion, 1 minute after injury (total n=9; 50ml: n=6, 500ml: n=3). Pigs receiving a bolus injection had an increased

yet stable heart rate; decreased, yet stable mean arterial pressure; and stable CO2 (Figure

8-6A, B and C respectively). Pigs treated with SynthoPlate 2.0 infusion had a decreased

310

Figure 8-6 Vital analysis in injury pigs administered SynthoPlate 2.0 at 50ml (bolus) or 500ml (infusion) show that [A] Heart rate increases with bolus injection and decreases with infusion [B] Mean Arterial Pressure decreases slightly and stabilized with bolus injection but, decreases significantly and rebounds (in animals that survive) and [C] CO2 remains stable with bolus injection but drops significantly and rebounds with an infusion. [D] Analysis of C3a plasma concentration in blood drawn showed no significant alterations.

yet stable heart rate; a large decrease, then increase, then stabilization of mean arterial pressure; and a large decrease in CO2, before stabilizing around 50 minutes after injury

(Figure 8-6A, B and C respectively). There is a slight increase in blood loss rate and total blood loss in bolus injected animals compared to infusion animals (Figure 8-7A and B), but this is likely a result of bolus injected animals maintaining a higher heart rate and blood

311 pressure during the experiment. Bolus treated animals have an 83%

Figure 8-7 Hemostatic efficacy analysis in injured pigs shows that administration of SynthoPlate 2.0 [A] with bolus injection leads to a slightly higher blood loss rate compared to infusion injection [B] and pigs administered with bolus injection have increased total blood loss rate compared to pigs that received infusion injection, however [C] bolus injection in pigs lead to an increase in survival compared to animals that received infusion injection (83% survival vs. 33% survival).

survival rate compared to a 33% survival rate in animals treated with an infusion of

SynthoPlate 2.0 (Figure 8-7C). Like the safety studies, bolus injected animals had little to

no skin response (Figure 8-3D) while animals receiving an infusion had generalized small

erythematous patches (Figure 8-3C). Neither administration method had a more than 4-

fold increase in complement C3a plasma concentration levels (Figure 8-6D).

8.4 Discussion

Applications of intravenously administered nanoparticles continue to grow for a variety of

. Nanoparticles are small enough that they can be injected into the blood stream

and can be designed to target, treat, and monitor with specificity. The surge of

nanomedicine is being aided by the scientific community’s expanding understanding of the

pathophysiology underlying the diseases we hope to treat and our expertise to engineering

nanoparticles23. With the widespread popularity of nanomedicine comes an increased risk

associated with their use. Adverse drug reactions will occur in 30% of hospitalized patients

312 and 77% of these will be complement mediated. Nanoparticles’ similarities to the pathogens the complement system is primed to combat make them likely to illicit an over activation of the complement system7.

In this study, we examine the effect different components of the drug delivery system can have on a hypersensitivity reaction. We have demonstrated that by just altering the composition (mol percent) of PEGylation of our nanoparticles, we can create a minimal versus a fatal hypersensitivity reaction. Increased PEGylation lead to stabilization of the heart rate and CO2, a mitigated skin response and decreased activation of the complement system. Though there were only slight differences in the hypersensitivity reaction observed when we varied the volume administered, we did see significant differences in efficacy, most importantly in survival. A 50ml injection of our hemostatic nanoparticles was able to increase survival by 50% compared to a 500ml injection. These results demonstrate that mitigating hypersensitivity reactions seen by promising nanomedicines is possible.

A four-fold increase in complement C3a plasma levels, which would be indicative of detrimental complement system activation, was not observed in our study, though the

SynthoPlate 1.0 PEG variant did show an increase in complement C3a levels. Complement

C3a plasma concentration was ideal to monitor for this study because complement C3 is cleaved into C3a in all complement pathways allowing us to study the involvement of the complement system in general. Subsequent studies should be performed to determine which complement pathway is being activated. The classical, alternative and lectin pathways can be studied individually by monitoring plasma levels of complement C1

313 complex, complement Ba (cleaved from C3bB), and the activity of MASP-1 or MASP-2 respectively24,25. It likely that the increases in C3a observed in this study are due to the activation of the alternative pathway, due to the how fast (less than one minute) the reaction to the treatments are observed. Additionally, to determine if the adaptive immune response is playing a role in the responses seen, IgE plasma concentrations should be measured.

Future studies should be targeted towards different characteristics of the nanoparticle design in hopes of understanding not only the pathophysiology of the drug hypersensitivity but also the cellular mechanisms. Our in vivo studies confirm how important each, seemingly minor, aspect of a nanoparticle design can have on hypersensitivity reactions, the success of a technology, and ultimately translating life altering treatments to our patients.

8.5 References

1. Prevention C for DC and. Deaths and Mortality. 2014.

2. Sauaia A, Moore F, Moore E, Moser K. Epidemiology of trauma deaths: a

reassessment. 1995; Available from:

https://journals.lww.com/jtrauma/Abstract/1995/02000/Epidemiology_of_Trauma

_Deaths__A_Reassessment.6.aspx

3. Champion HR, Bellamy RF, Roberts CP, Leppaniemi A. A profile of combat

injury. 2003;54(5 Suppl):S13–S19. PMID: 12768096

4. Shukla M, Sekhon U, Betapudi V, Li W, Hickman D, Modery-Pawlowski C, Dyer

M, Neal M, McCrae K, Sen Gupta A. In vitro characterization and in vivo

314 evaluation of SynthoPlateTM (synthetic platelet) technology in severely

thrombocytopenic mice. 2016;

5. Dyer MR, Hickman D, Luc N, Haldeman S, Loughran P, Pawlwoski C, Gupta A

Sen, Neal MD. Intravenous administration of synthetic platelets (SynthoPlateTM)

in a mouse liver injury model of uncontrolled hemorrhage improves hemostasis.

2018;84(6):1. Available from: http://insights.ovid.com/crossref?an=01586154-

900000000-98752

6. Hickman DA, Pawlowski CL, Shevitz A, Luc NF, Kim A, Girish A, Marks J,

Ganjoo S, Huang S, Niedoba E, Sekhon UDS, Sun M, Dyer M, Neal MD.

Intravenous synthetic platelet ( SynthoPlate ) nanoconstructs reduce bleeding and

improve ‘ golden hour ’ survival in a porcine model of traumatic arterial

hemorrhage. Springer US; 2018;(February):1–14. Available from:

http://dx.doi.org/10.1038/s41598-018-21384-z

7. Szebeni J. Hemocompatibility testing for nanomedicines and biologicals:

predictive assays for complement mediated infusion reactions. 2012;4(1):33–53.

Available from: http://www.degruyter.com/view/j/ejnm.2012.4.issue-1/ejnm-2012-

0002/ejnm-2012-0002.xml

8. Szebeni J, Bedőcs P, Csukás D, Rosivall L, Bünger R, Urbanics R. A porcine

model of complement-mediated infusion reactions to drug carrier nanosystems and

other medicines. 2012;64(15):1706–1716. Available from:

http://linkinghub.elsevier.com/retrieve/pii/S0169409X12002244

9. Szebeni J. Complement activation-related pseudoallergy: A new class of drug-

induced acute immune toxicity. 2005;216(2–3):106–121. Available from:

315 http://linkinghub.elsevier.com/retrieve/pii/S0300483X05003525

10. Ricklin D, Hajishengallis G, Yang K, Lambris JD. Complement: a key system for

immune surveillance and homeostasis. Nature Publishing Group; 2010;11(9):785–

797. Available from: http://www.nature.com/doifinder/10.1038/ni.1923

11. Szebeni J, Bedőcs P, Urbanics R, … RB-J of controlled, 2012 undefined.

Prevention of infusion reactions to PEGylated liposomal via

tachyphylaxis induction by placebo vesicles: a porcine model. Available from:

https://www.sciencedirect.com/science/article/pii/S0168365912001563

12. Szebeni J, Bedocs P, Rozsnyay Z, Weiszhár Z, Urbanics R, Rosivall L, Cohen R,

Garbuzenko O, Báthori G, Tóth M, Bünger R, Barenholz Y. Liposome-induced

complement activation and related cardiopulmonary distress in pigs: Factors

promoting reactogenicity of Doxil and AmBisome. 2012;8(2):176–184. PMID:

21704590

13. Dézsi L, Fülöp T, Mészáros T, … GS-J of controlled, 2014 undefined. Features of

complement activation-related pseudoallergy to liposomes with different surface

charge and PEGylation: comparison of the porcine and rat. Available from:

https://www.sciencedirect.com/science/article/pii/S0168365914005914

14. Onwukwe C, Maisha N, Holland M, Varley M, Groynom R, Hickman D, Uppal N,

Shoffstall A, Ustin J, Lavik E. Engineering Intravenously Administered

Nanoparticles to Reduce Infusion Reaction and Stop Bleeding in a Large Animal

Model of Trauma. 2018;acs.bioconjchem.8b00335. Available from:

http://pubs.acs.org/doi/10.1021/acs.bioconjchem.8b00335

15. Elliott JT, Prestwich GD. Maleimide-Functionalized Lipids that Anchor

316 Polypeptides to Lipid Bilayers and Membranes. 2000 Nov;11(6):832–841.

Available from: http://pubs.acs.org/doi/abs/10.1021/bc000022a

16. Presolski SI, Hong VP, Finn MG. Copper-Catalyzed Azide-Alkyne Click

Chemistry for Bioconjugation. Hoboken, NJ, USA: John Wiley & Sons, Inc.;

2011. Available from: http://doi.wiley.com/10.1002/9780470559277.ch110148

17. Ravikumar M, Modery CL, Wong TL, Gupta a S. Peptide-decorated liposomes

promote arrest and aggregation of activated platelets under flow on vascular injury

relevant protein surfaces in vitro. 2012;13(5):1495–1502. PMID: 22468641

18. Modery-Pawlowski CL, Tian LL, Ravikumar M, Wong TL, Gupta A Sen. In vitro

and in vivo hemostatic capabilities of a functionally integrated platelet-mimetic

liposomal nanoconstruct. Elsevier Ltd; 2013;34(12):3031–3041. Available from:

http://dx.doi.org/10.1016/j.biomaterials.2012.12.045 PMID: 23357371

19. Chanan-Khan a. Complement activation following first exposure to pegylated

liposomal doxorubicin (Doxil(R)): possible role in hypersensitivity reactions.

2003;14(9):1430–1437. Available from:

http://annonc.oupjournals.org/cgi/doi/10.1093/annonc/mdg374

20. Szebeni J, Fontana J, Wassef N, Circulation PM-, 1999 undefined. Hemodynamic

changes induced by liposomes and liposome-encapsulated hemoglobin in pigs: a

model for pseudoallergic cardiopulmonary reactions to. Available from:

https://scholar.google.com/scholar?hl=en&as_sdt=0%2C36&q=Hemodynamic+ch

anges+induced+by+liposomes+and+liposome-

encapsulated+hemoglobin+in+pigs&btnG=

21. Szebeni J, Muggia F, Gabizon A, Barenholz Y. Activation of complement by

317 therapeutic liposomes and other lipid excipient-based therapeutic products:

Prediction and prevention. Elsevier B.V.; 2011;63(12):1020–1030. Available

from: http://linkinghub.elsevier.com/retrieve/pii/S0169409X11001943

22. Urbanics R, Bedőcs P, Szebeni J. Lessons learned from the porcine CARPA

model: constant and variable responses to different nanomedicines and

administration protocols. De Gruyter; 2015 Jan 1;7(3):219–231. Available from:

https://www.degruyter.com/view/j/ejnm.2015.7.issue-3/ejnm-2015-0011/ejnm-

2015-0011.xml

23. Riehemann K, Schneider SW, Luger TA, Godin B, Ferrari M, Fuchs H.

Nanomedicine-Challenge and Perspectives. Wiley-Blackwell; 2009 Jan

19;48(5):872–897. Available from: http://doi.wiley.com/10.1002/anie.200802585

24. Palarasah Y, Nielsen C, Sprogøe U, Christensen ML, Lillevang S, Madsen HO,

Bygum A, Koch C, Skjodt K, Skjoedt M-O. Novel assays to assess the functional

capacity of the classical, the alternative and the lectin pathways of the complement

system. Wiley-Blackwell; 2011 Jun;164(3):388–95. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/21401574 PMID: 21401574

25. Roos A, Bouwman LH, van Gijlswijk-Janssen DJ, Faber-Krol MC, Stahl GL,

Daha MR. Human IgA Activates the Complement System Via the Mannan-

Binding Lectin Pathway. 2001;167(5):2861–2868. Available from:

http://www.jimmunol.org/cgi/doi/10.4049/jimmunol.167.5.2861 PMID: 11509633

318 Chapter 9: Conclusion and Future Directions

9.1 Conclusion

Trauma is a ubiquitous condition that could happen to anyone at any time in any place.

Knowing that it is one of the leading causes of deaths worldwide and that we have the

technology to stabilize patients if we could manage their bleeding until they reach medical

services, makes this an urgent and appealing problem to work towards solving1–3. Current hemostatic management of bleeding complications in trauma have limitations such as no pre-hospital availability, short shelf life, and high risk of contamination that lead to the high rate of preventable traumatic hemorrhage related deaths seen today2,3. These risks are

increased in patients with thrombocytopenia. The overall hypothesis for this work is that

the design of nanoscale delivery systems with platelet-inspired heteromultivalent surface- modifications, can enable injury site-specific biointeractions and drug delivery to enhance platelet-mediated hemostatic processes, leading to decreased blood loss and increased

survival.

Chapters 2 and 3, are a detailed review of current materials and technologies for the

management of traumatic bleeding. Though there are a number of approved and ground-

breaking technologies for managing bleeding, treatment of internal non-compressible

hemorrhage still heavily depends on transfusion of whole blood or blood’s hemostatic

components (platelets, fibrinogen and coagulation factors). The recent clinical trial

PROPPR (Pragmatic, Randomized Optimal Platelet and Plasma Ratios) demonstrated the

importance of platelets, in particular, for the management of traumatic bleeding. The study

319 determined that early platelet administration is associated with improved hemostasis and

reduced mortality in traumatic bleeding patients compared to patients who received blood

products without platelets4.

In Chapter 4, our ability to engineer a synthetic nanoparticle based functional mimic of platelets that can leverage and amplify the body’s physiological clotting mechanisms specifically at the bleeding site was reported. This nanoparticle, SynthoPlateTM, was

evaluated in vitro and found to be a ~150nm unilaminar vesicle that does not activate or

aggregate circulating quiescent platelets, but that co-localizes specifically with activated

platelets in clots enhancing their aggregations and in effect, enhancing fibrin generation.

Also described, was that in a thrombocytopenic mice model, the hemostatic efficacy of

these particles was found to be dose-dependent, with doses comparable to normal murine

platelet count (800-1000 particles per nL) improving bleeding time in thrombocytopenic

mice to levels of normal mice.

Building on these results, Chapter 5 outlines the evaluation of SynthoPlate in the

pretreatment and an emergency application of a mouse liver injury. Blood loss was

significantly reduced and 72hr survival was significantly increased for mice treated with

SynthoPlate in both models. After these promising results, we moved to a large animal

injury model, the pig femoral artery injury, to evaluate the efficacy and safety of

SynthoPlate in Chapter 6. Pigs were chosen because of the similarities between their

cardiovascular system and the human cardiovascular system5. In these studies, SynthoPlate

significantly reduces blood loss rate, stabilizes blood pressure and increases survival during

320 the ‘golden hour’ after traumatic arterial injury compared to control groups. It was also found to be systemically safe at the administered dose.

The development of a platelet-inspired nanovehicle for injury-site targeted delivery of

Tranexamic Acid, TXA, was reported in Chapter 7. We were able to successful encapsulate

TXA in our platelet-targeting nanovehicle and demonstrate in vivo efficacy of drug release.

In vivo, the TXA-loaded targeting nanovesicles were able to safely reduce blood loss and increase 1hr and 72hr survival in a rat liver injury model compared to free TXA and controls.

Chapter 8 investigates ways to refine nanomedicine design to improve systemic safety, by altering surface charge and hydrophilicity, as well as modulating the administration protocol. Here we demonstrate that by altering the composition (mol percent) of

PEGylation of our nanoparticles, we can create a minimal versus a fatal hypersensitivity reaction.

Overall, these studies prove the hypothesis that the design of nanoscale delivery systems with platelet-inspired heteromultivalent surface-modifications, can enable injury site- specific biointeractions and drug delivery to enhance platelet-mediated hemostatic processes, leading to decreased blood loss and increased survival.

321 9.2 Future Directions

9.2.1 Nanovesicle Design

In our attempt to manage traumatic bleeding, we studied the bleeding mechanisms and decided to mimic adhesion and aggregation properties of platelets. Now that we have

successfully mimicked these properties, we could possibly increase the efficacy of our

technology by mimicking even more platelet functions. When platelets become activated

they release a number of molecules from their granules which help with propagation of the

clotting cascade and thrombus formation. In Chapter 7, we have demonstrated that we can

load and deliver payloads with our nanovesicles. Selecting a platelet relevant molecule, or

cocktail of molecules such as polyphosphate, ADP, Factor V, vWF, and fibrinogen to

deliver could help enhance the hemostatic environment. Activated platelets also provide a

highly negative surface for the propagation of the coagulation cascade. We could create a

negatively charged surface on our nanovehicles that is masked, to prevent spontaneous

thrombus formation, until the masking molecules are cleaved by a clot relevant enzyme.

This would allow for our nanovehicle to interact directly with and enhance the actions of

clotting factors.

Additionally, we might consider mimicking the shape and size of platelets. Our lab just

published a study reporting that in the presence of red blood cells, micro-scale non-

spherical particles undergo enhanced ‘margination and adhesion’ compared to nano-scale

spherical particles, resulting in their higher binding6. Though our particles do show efficacy

at reducing blood loss and increasing survival demonstrating their availability to marginate

and adhere in vivo, altering their shape and size may further increase their efficacy.

322 9.2.2 Nanoparticle Evaluation

Our nanovesicle design are evaluated in vitro before they are tested in vivo for practical, ethical and cost reasons. It is our goal to have a system that will model in vivo systems as close as possible. Our current in vitro systems do have their own limitations. For example, in the flow studies, we typically run our samples in platelet rich plasma over vWF and collagen coated surfaces with laminar flow. In the future, we could design a system that allows us to test our designs in more biologically relevant condition such as a design that incorporates endothelial cells, pulsatile flow and whole blood would begin to incorporate some of the variables currently lacking in our system.

Our current in vivo models work to recreate trauma scenarios, but here is currently no standard traumatic hemorrhage animal model. This may be in part because there is no single animal model that can answer all questions. We do our best to utilize the appropriate model for the question at hand. A recent study examining experimental models of coagulopathy found 62 relevant publications describing 27 distinct models of traumatic coagulopathy7. Though pig models were the most common species, other model species included rat, rabbit, sheep, mouse and hamster. Although the overall hemostatic processes may be conserved among many mammals, the response to nanoparticle administration may vary widely from species to species and is important to understand before beginning animal studies. Though we have used mice and rats to model traumatic hemorrhage, our pig model is likely the best model because the pig cardiovascular system is similar to the human cardiovascular system. Unfortunately, the pig model is costly, requires special surgery facilities, special housing, and complex handling which all lead to reduce sample sizes. In

323 our studies, we tested one dose one minute after injury. It is unlikely that patients will receive our treatment one minute after trauma. In future studies it will be important to correlate the treatment efficacy with administration time to better simulate real-world scenarios. Additionally, future studies should be conducted to determine the maximum tolerated dose and minimum effective dose.

9.2.3 Nanoparticle Technology Translation

One of the most challenging yet promising future directions for this research will be its development into an FDA approved drug. Patents have been issued on SynthoPlateTM design (composition) and methods of use, and a company (Haima Therapeutics, LLC) has been established to move SynthoPlateTM to the marketplace. The next steps will include adopting good manufacturing practice so that an industry consistent product is being utilized. The therapeutic window will need to be determined by performing maximum tolerated dose and minimum effective dose studies as well as performing toxicity studies to determine a safety profile. Stability studies in both suspension and lyophilized conditions at varying temperatures will also need to be conducted. Hopefully, the results of this future work will lead to an FDA approved drug that can be utilized for the hemostatic management of bleeding complications in thrombocytopenia and trauma.

9.3 References

1. WHO | Global Health Estimates. World Health Organization; 2018; Available

from: http://www.who.int/healthinfo/global_burden_disease/en/

2. Davis JS, Satahoo SS, Butler FK, Dermer H, Naranjo D, Julien K, Van Haren RM,

324 Namias N, Blackbourne LH, Schulman CI. An analysis of prehospital deaths: Who

can we save? Accreditation Statement AMA PRA Category 1 Creditsi. 2014;

Available from: http://www.aast.org/

3. Teixeira PGR, Inaba K, Hadjizacharia P, Brown C, Salim A, Rhee P, Browder T,

Noguchi TT, Demetriades D. Preventable or Potentially Preventable Mortality at a

Mature Trauma Center. 2007; Available from:

https://insights.ovid.com/pubmed?pmid=18212658

4. Cardenas JC, Zhang X, Fox EE, Cotton BA, Hess JR, Schreiber MA, Wade CE,

Holcomb JB, PROPPR Study Group. Platelet transfusions improve hemostasis and

survival in a substudy of the prospective, randomized PROPPR trial. American

Society of Hematology; 2018 Jul 24;2(14):1696–1704. Available from:

http://www.ncbi.nlm.nih.gov/pubmed/30030268 PMID: 30030268

5. Zaragoza C, Gomez-Guerrero C, Martin-Ventura JL, Blanco-Colio L, Lavin B,

Mallavia B, Tarin C, Mas S, Ortiz A, Egido J. Animal Models of Cardiovascular

Diseases. 2011;2011:1–13. Available from:

http://www.hindawi.com/journals/bmri/2011/497841/ PMID: 21403831

6. Cooley M, Sarode A, Hoore M, Fedosov DA, Mitragotri S, Sen Gupta A.

Influence of particle size and shape on their margination and wall-adhesion:

Implications in drug delivery vehicle design across nano-to-micro scale. The Royal

Society of Chemistry; 2018; Available from:

http://pubs.rsc.org/en/Content/ArticleLanding/2018/NR/C8NR04042G

7. van Zyl N, Reade MC, Fraser JF. Experimental Animal Models of Traumatic

Coagulopathy. 2015 Jul;44(1):16–24. Available from:

325 http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=00

024382-201507000-00003

326 Bibliography

Acker, J. P., Marks, D. C., & Sheffield, W. P. (2016). Quality Assessment of Established and Emerging Blood Components for Transfusion. Journal of Blood Transfusion, 2016, 1–28. http://doi.org/10.1155/2016/4860284 Agam, G., & Livne, A. A. (1992). Erythrocytes with covalently bound fibrinogen as a cellular replacement for the treatment of thrombocytopenia. European Journal of Clinical Investigation, 22(2), 105–12. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/1572388 Al-Belasy, F. A., & Amer, M. Z. (2003). Hemostatic effect of n-butyl-2-cyanoacrylate (histoacryl) glue in warfarin-treated patients undergoing oral surgery. Journal of Oral and Maxillofacial Surgery, 61(12), 1405–1409. http://doi.org/10.1016/J.JOMS.2002.12.001 Alam, H. B., Burris, D., DaCorta, J. a, & Rhee, P. (2005). Hemorrhage control in the battlefield: role of new hemostatic agents. Military Medicine, 170(1), 63–69. http://doi.org/10.7205/MILMED.170.1.63 Alam, H. B., Chen, Z., Jaskille, A., Querol, R. I. L. C., Koustova, E., Inocencio, R., … Rhee, P. (2004). Application of a Zeolite Hemostatic Agent Achieves 100% Survival in a Lethal Model of Complex Groin Injury in Swine. The Journal of Trauma: Injury, Infection, and Critical Care, 56(5), 974–983. http://doi.org/10.1097/01.TA.0000127763.90890.31 Albala, D. M. (2003). Fibrin sealants in clinical practice. Cardiovascular Surgery, 11, 5– 11. http://doi.org/10.1016/S0967-2109(03)00065-6 Allen, G. A., Monroe, D. M., Roberts, H. R., & Hoffman, M. (2000). The effect of factor X level on thrombin generation and the procoagulant effect of activated factor VII in a cell-based model of coagulation. Blood Coagulation and Fibrinolysis, 11(4 SUPPL. 1), 3–7. http://doi.org/10.1097/00001721-200004001-00002 AlMomani, T., Udaykumar, H. S., Marshall, J. S., & Chandran, K. B. (2008). Micro-scale Dynamic Simulation of Erythrocyte–Platelet Interaction in Blood Flow. Annals of , 36(6), 905–920. http://doi.org/10.1007/s10439-008-9478-z American College of Surgeons. Committee on Trauma. (2008). ATLS, advanced trauma life support for doctors. American College of Surgeons. Andre, P. (2004). P-selectin in haemostasis. British Journal of Haematology, 126(3), 298–306. http://doi.org/10.1111/j.1365-2141.2004.05032.x Andrews, R. K., & Berndt, M. C. (2004). Platelet physiology and thrombosis. Thrombosis Research, 114(5–6), 447–453. http://doi.org/10.1016/J.THROMRES.2004.07.020 Anema, J. G., Morey, A. F., Harris, R., MacPhee, M., & Cornum, R. L. (2001). Potential Uses of Absorbable Fibrin Adhesive Bandage for Genitourinary Trauma. World Journal of Surgery, 25(12), 1573–1577. http://doi.org/10.1007/s00268-001-0152-y

327 Annabi, N., Tamayol, A., Shin, S. R., Ghaemmaghami, A. M., Peppas, N. A., & Khademhosseini, A. (2014). Surgical materials: Current challenges and nano-enabled solutions. Nano Today, 9(5), 574–589. http://doi.org/10.1016/J.NANTOD.2014.09.006 Annen, K., & Olson, J. E. (2015). Optimizing platelet transfusions. Current Opinion in Hematology, 22(6), 559–564. http://doi.org/10.1097/MOH.0000000000000188 Anselmo, A. C., Modery-Pawlowski, C. L., Menegatti, S., Kumar, S., Vogus, D. R., Tian, L. L., … Mitragotri, S. (2014). Platelet-like Nanoparticles: Mimicking Shape, Flexibility, and Surface Biology of Platelets To Target Vascular Injuries. ACS Nano, 8(11), 11243–11253. http://doi.org/10.1021/nn503732m Antonik, L. M., Khabibulina, A. G., Fadeeva, T. V., Kogan, A. S., & Malyshev, A. V. (2007). Hemostatic complexes based on polyacrylic acid. Pharmaceutical Chemistry Journal, 41(5), 278–280. http://doi.org/10.1007/s11094-007-0062-x Apelseth, T. O., Cap, A. P., Spinella, P. C., Hervig, T., & Strandenes, G. (2017). Cold stored platelets in treatment of bleeding. ISBT Science Series, 12(4), 488–495. http://doi.org/10.1111/voxs.12380 Approach to the adult with unexplained thrombocytopenia - UpToDate. (n.d.). Retrieved August 4, 2018, from https://www.uptodate.com/contents/approach-to-the-adult-with- unexplained- thrombocytopenia?search=thrombocytopenia&source=search_result&selectedTitle=1 ~150&usage_type=default&display_rank=1#H20171687 Arat, Y. O., Dorotheo, E. U., Tang, R. A., Boniuk, M., & Schiffman, J. S. (2006). Compressive Optic Neuropathy after Use of Oxidized Regenerated Cellulose in Orbital Surgery: Review of Complications, Prophylaxis, and Treatment. , 113(2), 333–337. http://doi.org/10.1016/J.OPHTHA.2005.11.003 Ariëns, R., Lai, T., Weisel, J., & Greenberg, C. (2002). Role of factor XIII in fibrin clot formation and effects of genetic polymorphisms. Blood. Retrieved from http://www.bloodjournal.org/content/100/3/743.short Arnaud, F., Teranishi, K., Okada, T., Parreño-Sacdalan, D., Hupalo, D., McNamee, G., … McCarron, R. (2011). Comparison of Combat Gauze and TraumaStat in Two Severe Groin Injury Models. Journal of Surgical Research, 169(1), 92–98. http://doi.org/10.1016/J.JSS.2009.09.004 Arnaud, F., Teranishi, K., Tomori, T., Carr, W., & McCarron, R. (2009). Comparison of 10 hemostatic dressings in a groin puncture model in swine. Journal of Vascular Surgery, 50(3), 632–639.e1. http://doi.org/10.1016/J.JVS.2009.06.010 Arnaud, F., Tomori, T., Carr, W., McKeague, A., Teranishi, K., Prusaczyk, K., & McCarron, R. (2008). Exothermic Reaction in Zeolite Hemostatic Dressings: QuikClot ACS and ACS+®. Annals of Biomedical Engineering, 36(10), 1708–1713. http://doi.org/10.1007/s10439-008-9543-7 Arnaud, F., Tomori, T., Saito, R., McKeague, A., Prusaczyk, W. K., & McCarron, R. M. (2007). Comparative Efficacy of Granular and Bagged Formulations of the

328 Hemostatic Agent QuikClot. The Journal of Trauma: Injury, Infection, and Critical Care, 63(4), 775–782. http://doi.org/10.1097/TA.0b013e31805f7023 Arnout, J., Hoylaerts, M. F., & Lijnen, H. R. (2006). Haemostasis. In The Vascular Endothelium II (pp. 1–41). Springer Berlin Heidelberg. http://doi.org/10.1007/3-540- 36028-X_1 Arya, R. C., Wander, G., & Gupta, P. (2011). Blood component therapy: Which, when and how much. Journal of Anaesthesiology, Clinical Pharmacology, 27(2), 278–84. http://doi.org/10.4103/0970-9185.81849 Avanzi, M. P., & Mitchell, W. B. (2014). Ex Vivo production of platelets from stem cells. British Journal of Haematology, 165(2), 237–247. http://doi.org/10.1111/bjh.12764 Avery, R. K., Albadawi, H., Akbari, M., Zhang, Y. S., Duggan, M. J., Sahani, D. V, … Oklu, R. (2016). An injectable shear-thinning biomaterial for endovascular embolization. Science Translational Medicine, 8(365), 365ra156. http://doi.org/10.1126/scitranslmed.aah5533 Aydemir Sezer, U., Kocer, Z., Aru, B., Demirel, G. Y., Gulmez, M., Aktekin, A., … Sezer, S. (2016). Combination of gelatin and tranexamic acid offers improved haemostasis and safe use on internal hemorrhage control. RSC Advances, 6(97), 95189–95198. http://doi.org/10.1039/C6RA16790J Bachet, J., Goudot, B., Dreyfus, G., Banfi, C., Ayle, N. A., Aota, M., … Guilmet, D. (1997). The proper use of glue: a 20-year experience with the GRF glue in acute aortic dissection. Journal of Cardiac Surgery, 12(2 Suppl), 243-53; discussion 253–5. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9271753 Bajerová, M., Krejčová, K., Rabišková, M., Gajdziok, J., & Masteiková, R. (2009). Oxycellulose: Significant characteristics in relation to its pharmaceutical and medical applications. Advances in Polymer Technology, 28(3), 199–208. http://doi.org/10.1002/adv.20161 BAK, J. B., SINGH, A., & SHEKARRIZ, B. (2004). Use of Gelatin Matrix Thrombin Tissue Sealant as an Effective Hemostatic Agent During Laparoscopic Partial Nephrectomy. The Journal of , 171(2), 780–782. http://doi.org/10.1097/01.JU.0000104800.97009.C6 Bakar, B., Oruckaptan, H. H., Hazer, B. D., Saatci, I., Atilla, P., Kılıc, K., & Muftuoglu, S. F. (2010). Evaluation of the toxicity of onyx compared with n-butyl 2- cyanoacrylate in the subarachnoid space of a rabbit model: an experimental research. Neuroradiology, 52(2), 125–134. http://doi.org/10.1007/s00234-009-0594-8 Balakrishnan, B., Mohanty, M., Umashankar, P. R., & Jayakrishnan, A. (2005). Evaluation of an in situ forming hydrogel wound dressing based on oxidized alginate and gelatin. Biomaterials, 26(32), 6335–6342. http://doi.org/10.1016/J.BIOMATERIALS.2005.04.012 Baylis, J. R., Chan, K. Y. T., & Kastrup, C. J. (2016). Halting hemorrhage with self- propelling particles and local drug delivery. Thrombosis Research, 141 Suppl 2(Suppl 2), S36-9. http://doi.org/10.1016/S0049-3848(16)30362-0

329 Baylis, J. R., Yeon, J. H., Thomson, M. H., Kazerooni, A., Wang, X., St. John, A. E., … Kastrup, C. J. (2015). Self-propelled particles that transport cargo through flowing blood and halt hemorrhage. Science Advances, 1(9), e1500379–e1500379. http://doi.org/10.1126/sciadv.1500379 Beer, J. H., Springer, K. T., & Coller, B. S. (1992). Immobilized Arg-Gly-Asp (RGD) peptides of varying lengths as structural probes of the platelet glycoprotein IIb/IIIa receptor. Blood, 79(1), 117–28. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/1728303 Benesch, J., & Tengvall, P. (2002). Blood protein adsorption onto chitosan. Biomaterials, 23(12), 2561–2568. http://doi.org/10.1016/S0142-9612(01)00391-X Bennett-Marsden, M. (2010). How to select a wound dressing. Retrieved August 26, 2018, from https://www.pharmaceutical-journal.com/career/career-feature/how-to- select-a-wound-dressing/11039846.article?firstPass=false Bennett, B. L., & Littlejohn, L. (2014). Review of New Topical Hemostatic Dressings for Combat Casualty Care. Military Medicine, 179(5), 497–514. http://doi.org/10.7205/MILMED-D-13-00199 Bennett, J. S. (1996). Structural biology of glycoprotein IIb-IIIa. Trends in Cardiovascular Medicine, 6(1), 31–36. http://doi.org/10.1016/1050-1738(95)00126-3 Bennett, J. S. (2005). Structure and function of the platelet integrin alphaIIbbeta3. The Journal of Clinical Investigation, 115(12), 3363–9. http://doi.org/10.1172/JCI26989 Bennett, J. S., Shattil, S. J., Power, J. W., & Gartner, T. K. (1988). Interaction of fibrinogen with its platelet receptor. Differential effects of alpha and gamma chain fibrinogen peptides on the glycoprotein IIb-IIIa complex. The Journal of Biological Chemistry, 263(26), 12948–53. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/3417645 Berrettini, M., Mariani, G., Schiavoni, M., Rocino, A., Di Paolantonio, T., Longo, G., & Morfini, M. (2001). Pharmacokinetic evaluation of recombinant, activated factor VII in patients with inherited factor VII deficiency. Haematologica, 86(6), 640–5. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/11418374 Bertoni, F., Barbani, N., Giusti, P., & Ciardelli, G. (2006). Transglutaminase Reactivity with Gelatine: Perspective Applications in Tissue Engineering. Biotechnology Letters, 28(10), 697–702. http://doi.org/10.1007/s10529-006-9046-2 Bertram, J. P., Williams, C. a, Robinson, R., Segal, S. S., Flynn, N. T., & Lavik, E. B. (2009). Intravenous hemostat: nanotechnology to halt bleeding. Science Translational Medicine, 1(11), 11ra22. http://doi.org/10.1126/scitranslmed.3000397 Bhattacharya, S. S., Banerjee, S., Chowdhury, P., Ghosh, A., Hegde, R. R., & Mondal, R. (2013). Tranexamic acid loaded gellan gum-based polymeric microbeads for controlled release: In vitro and in vivo assessment. Colloids and Surfaces B: Biointerfaces, 112, 483–491. http://doi.org/10.1016/j.colsurfb.2013.07.054

330 Bigi, A., Cojazzi, G., Panzavolta, S., Roveri, N., & Rubini, K. (2002). Stabilization of gelatin films by crosslinking with genipin. Biomaterials, 23(24), 4827–4832. http://doi.org/10.1016/S0142-9612(02)00235-1 Biró, E., Akkerman, J. W. N., Hoek, F. J., Gorter, G., Pronk, L. M., Sturk, A., & Nieuwland, R. (2005). The phospholipid composition and cholesterol content of platelet-derived microparticles: a comparison with platelet membrane fractions. Journal of Thrombosis and Haemostasis : JTH, 3(12), 2754–63. http://doi.org/10.1111/j.1538-7836.2005.01646.x Bishop, P., & Lawson, J. (2004). Recombinant biologics for treatment of bleeding disorders. Nature Reviews Drug Discovery, 3(8), 684–694. http://doi.org/10.1038/nrd1443 Björses, K., & Holst, J. (2007). Various Local Hemostatic Agents with Different Modes of Action; an in vivo Comparative Randomized Vascular Surgical Experimental Study. European Journal of Vascular and Endovascular Surgery, 33(3), 363–370. http://doi.org/10.1016/J.EJVS.2006.10.011 Björses K, & Holst J. (2009). Topical haemostatics in renal trauma-an evaluation of four different substances in an experimental setting. J Trauma (Vol. 66). Retrieved from http://www.scielo.br/pdf/ibju/v35n3/v35n3a22 Blackbourne, L. H., Baer, D. G., Eastridge, B. J., Kheirabadi, B., Kragh, J. F., Cap, A. P., … Holcomb, J. B. (2012). Military medical revolution. Journal of Trauma and Acute Care Surgery, 73, S372–S377. http://doi.org/10.1097/TA.0b013e3182755662 Blajchman, M. A. (1998). Bacterial contamination and proliferation during the storage of cellular blood products. Vox Sanguinis, 74 Suppl 2, 155–9. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9704439 Blajchman, M. A. (1999). Substitutes for success. Nature Medicine, 5(1), 17–18. http://doi.org/10.1038/4694 Blajchman, M. A. (2001). Novel platelet products, substitutes and alternatives. Transfusion Clinique et Biologique, 8(3), 267–271. http://doi.org/10.1016/S1246- 7820(01)00127-6 Blajchman, M. A. (2003). Substitutes and alternatives to platelet transfusions in thrombocytopenic patients. Journal of Thrombosis and Haemostasis, 1(7), 1637– 1641. http://doi.org/10.1046/j.1538-7836.2003.00332.x Blann, A., Nadar, S. K., & Lip, G. Y. H. (2003). The adhesion molecule P-selectin and cardiovascular disease. European Heart Journal, 24(24), 2166–2179. http://doi.org/10.1016/j.ehj.2003.08.021 Bochicchio, G., Dunne, J., Bochicchio, K., & Scalea, T. (2004). The Combination of Platelet-Enriched Autologous Plasma with Bovine Collagen and Thrombin Decreases the Need for Multiple Blood Transfusions in Trauma Patients with . The Journal of Trauma: Injury, Infection, and Critical Care, 56(1), 76–79. http://doi.org/10.1097/01.TA.0000103985.26884.D2

331 Bode, A. P., & Fischer, T. H. (2007). Lyophilized Platelets: Fifty Years in the Making. Artificial Cells, Blood Substitutes, and Biotechnology, 35(1), 125–133. http://doi.org/10.1080/10731190600974962 Bode, A. P., & Read, M. S. (2000). Lyophilized platelets: continued development. Transfusion Science, 22(1–2), 99–105. http://doi.org/10.1016/S0955-3886(00)00028- X Boffard, K. D., Riou, B., Warren, B., Choong, P. I. T., Rizoli, S., Rossaint, R., … Kluger, Y. (2005). Recombinant Factor VIIa as Adjunctive Therapy for Bleeding Control in Severely Injured Trauma Patients: Two Parallel Randomized, Placebo-Controlled, Double-Blind Clinical Trials. The Journal of Trauma: Injury, Infection, and Critical Care, 59(1), 8–18. http://doi.org/10.1097/01.TA.0000171453.37949.B7 Bordin, J. O., Heddle, N. M., & Blajchman, M. A. (1994). Biologic effects of leukocytes present in transfused cellular blood products. Blood, 84(6), 1703–21. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/8080981 Borgman, M. A., Spinella, P. C., Perkins, J. G., Grathwohl, K. W., Repine, T., Beekley, A. C., … Holcomb, J. B. (2007). The Ratio of Blood Products Transfused Affects Mortality in Patients Receiving Massive Transfusions at a Combat Support Hospital. The Journal of Trauma: Injury, Infection, and Critical Care, 63(4), 805–813. http://doi.org/10.1097/TA.0b013e3181271ba3 Borst, H. G., Haverich, A., Walterbusch, G., & Maatz, W. (1982). Fibrin adhesive: an important hemostatic adjunct in cardiovascular operations. The Journal of Thoracic and Cardiovascular Surgery, 84(4), 548–53. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/6981733 Boscarino, C., Tien, H., Acker, J., Callum, J., Hansen, A. L., Engels, P., … Beckett, A. (2014). Feasibility and transport of packed red blood cells into Special Forces operational conditions. Journal of Trauma and Acute Care Surgery, 76(4), 1013– 1019. http://doi.org/10.1097/TA.0000000000000173 Boulton, F. (2006). A hundred years of cascading - started by Paul Morawitz (1879- 1936), a pioneer of haemostasis and of transfusion. Transfusion Medicine, 16(1), 1– 10. http://doi.org/10.1111/j.1365-3148.2006.00643.x Bozzuto, G., & Molinari, A. (2015). Liposomes as nanomedical devices. International Journal of Nanomedicine, 10, 975–99. http://doi.org/10.2147/IJN.S68861 Briggs, A., & Askari, R. (2016). Damage control resuscitation. International Journal of Surgery, 33, 218–221. http://doi.org/10.1016/j.ijsu.2016.03.064 Browder, I. W., & Litwin, M. S. (1986). Use of absorbable collagen for hemostasis in general surgical patients. The American Surgeon, 52(9), 492–4. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/3752727 Brown, A. C., Stabenfeldt, S. E., Ahn, B., Hannan, R. T., Dhada, K. S., Herman, E. S., … Barker, T. H. (2014). Ultrasoft microgels displaying emergent platelet-like behaviours. Nature Materials, 13(September), 1–7. http://doi.org/10.1038/nmat4066

332 Brown, J. B., Neal, M. D., Guyette, F. X., Peitzman, A. B., Billiar, T. R., Zuckerbraun, B. S., & Sperry, J. L. (2015). Design of the Study of Tranexamic Acid during Air Medical Prehospital Transport (STAAMP) Trial: Addressing the Knowledge Gaps. Prehospital Emergency Care, 19(1), 79–86. http://doi.org/10.3109/10903127.2014.936635 Brown, M. A., Daya, M. R., & Worley, J. A. (2009). Experience with Chitosan Dressings in a Civilian EMS System. The Journal of , 37(1), 1–7. http://doi.org/10.1016/J.JEMERMED.2007.05.043 Broze, G. J., & Higuchi, D. (1996). Coagulation-dependent inhibition of fibrinolysis: role of carboxypeptidase-U and the premature lysis of clots from hemophilic plasma. Blood, 88(10). Retrieved from http://www.bloodjournal.org/content/88/10/3815.short?sso-checked=true Buchta, C., Dettke, M., Funovics, P. T., Hocker, P., Knobl, P., Macher, M., … Worel, N. (2004). Fibrin sealant produced by the CryoSealR FS System: product chemistry, material properties and possible preparation in the autologous preoperative setting. Vox Sanguinis, 86(4), 257–262. http://doi.org/10.1111/j.0042-9007.2004.00516.x Buckley, M. F., James, J. W., Brown, D. E., Whyte, G. S., Dean, M. G., Chesterman, C. N., & Donald, J. A. (2000). A novel approach to the assessment of variations in the human platelet count. Thrombosis and Haemostasis, 83(3), 480–4. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10744157 Burgert, J. M., Gegel, B. T., Austin, R., Davila, A., Deeds, J., Hodges, L., … Johnson, D. (2010). Effects of arterial blood pressure on rebleeding using Celox and TraumaDEX in a porcine model of lethal femoral injury. AANA Journal, 78(3), 230–6. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/20572410 Burke, J. E., & Dennis, E. A. (2009). Phospholipase A2 structure/function, mechanism, and signaling. Journal of Lipid Research, 50 Suppl, S237-42. http://doi.org/10.1194/jlr.R800033-JLR200 Buskens, E., Meijboom, M. J., Kooijman, H., & Van Hout, B. A. (2006). The use of a surgical sealant (CoSeal ® ) in cardiac and vascular reconstructive surgery: an economic analysis. THE JOURNAL OF CARDIOVASCULAR SURGERY J CARDIOVASC SURG (Vol. 47). Retrieved from https://pdfs.semanticscholar.org/2d9a/6cbbc75294564117f2013964658f7d239ebf.pdf Busuttil, R. W. (2003). A comparison of antifibrinolytic agents used in hemostatic fibrin sealants. Journal of the American College of Surgeons, 197(6), 1021–1028. http://doi.org/10.1016/j.jamcollsurg.2003.07.002 Butenas, S., Brummel, K. E., Branda, R. F., Paradis, S. G., & Mann, K. G. (2002). Mechanism of factor VIIa-dependent coagulation in hemophilia blood. Blood, 99(3), 923–930. http://doi.org/10.1182/blood.V99.3.923 Butler, F. K., & Blackbourne, L. H. (2012). Battlefield trauma care then and now. Journal of Trauma and Acute Care Surgery, 73, S395–S402. http://doi.org/10.1097/TA.0b013e3182754850 Cain, J. (2008). No Title. J. Truma: War Trauma, 3(16).

333 Calvini, P., Conio, G., Lorenzoni, M., & Pedemonte, E. (2004). Viscometric determination of dialdehyde content in periodate oxycellulose. Part I. Methodology. Cellulose, 11(1), 99–107. http://doi.org/10.1023/B:CELL.0000014766.31671.8d Calvini, P., Conio, G., Princi, E., Vicini, S., & Pedemonte, E. (2006). Viscometric determination of dialdehyde content in periodate oxycellulose Part II. Topochemistry of oxidation. Cellulose, 13(5), 571–579. http://doi.org/10.1007/s10570-005-9035-y Cannon, J. W. (2018). Hemorrhagic Shock. New England Journal of Medicine, 378(4), 370–379. http://doi.org/10.1056/NEJMra1705649 Cap, A. P., & Perkins, J. G. (2011). Lyophilized Platelets: Challenges and Opportunities. The Journal of Trauma: Injury, Infection, and Critical Care, 70, S59–S60. http://doi.org/10.1097/TA.0b013e31821a606d Cap, A. P., & Spinella, P. C. (2017). Just chill-it’s worth it! Transfusion, 57(12), 2817– 2820. http://doi.org/10.1111/trf.14399 Capdevila, X., Calvet, Y., Biboulet, P., Biron, C., Rubenovitch, J., & d’Athis, F. (1998). Aprotinin decreases blood loss and homologous transfusions in patients undergoing major orthopedic surgery. , 88(1), 50–7. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9447855 Cardenas, J. C., Zhang, X., Fox, E. E., Cotton, B. A., Hess, J. R., Schreiber, M. A., … PROPPR Study Group. (2018). Platelet transfusions improve hemostasis and survival in a substudy of the prospective, randomized PROPPR trial. Blood Advances, 2(14), 1696–1704. http://doi.org/10.1182/bloodadvances.2018017699 Carmen, R. (1993). The selection of plastic materials for blood bags. Transfusion Medicine Reviews, 7(1), 1–10. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/8431654 Carr, M., Kenawy, E., Layman, J., Wnek, G., & Ward, K. (2002). Development Of The Biohemostat-A Treatment Modality For High Pressure Bleeding Based On Super Absorbent Polymers. Shock, 17, 54. Retrieved from https://journals.lww.com/shockjournal/Citation/2002/06001/DEVELOPMENT_OF_T HE_BIOHEMOSTAT___A_TREATMENT.163.aspx Carraway, J. W., Kent, D., Young, K., Cole, A., Friedman, R., & Ward, K. R. (2008). Comparison of a new mineral based hemostatic agent to a commercially available granular zeolite agent for hemostasis in a swine model of lethal extremity arterial hemorrhage. Resuscitation, 78(2), 230–235. http://doi.org/10.1016/J.RESUSCITATION.2008.02.019 Casati, V., Guzzon, D., Oppizzi, M., Cossolini, M., Torri, G., Calori, G., & Alfieri, O. (1999). Hemostatic effects of aprotinin, tranexamic acid and epsilon-aminocaproic acid in primary cardiac surgery. The Annals of Thoracic Surgery, 68(6), 2252-6; discussion 2256-7. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10617012 Casey, B. J., Behrens, A. M., Hess, J. R., Wu, Z. J., Griffith, B. P., & Kofinas, P. (2010). FVII Dependent Coagulation Activation in Citrated Plasma by Polymer Hydrogels. Biomacromolecules, 11(12), 3248–3255. http://doi.org/10.1021/bm101147w

334 Cejas, M. A., Chen, C., Kinney, W. A., & Maryanoff, B. E. (2007). Nanoparticles That Display Short Collagen-Related Peptides. Potent Stimulation of Human Platelet Aggregation by Triple Helical Motifs. Bioconjugate Chemistry, 18(4), 1025–1027. http://doi.org/10.1021/bc070105s Champion, H. R., Bellamy, R. F., Roberts, C. P., & Leppaniemi, A. (2003). A profile of combat injury. The Journal of Trauma, 54(5 Suppl), S13–S19. http://doi.org/10.1097/01.TA.0000057151.02906.27 Chan, L. W., Kim, C. H., Wang, X., Pun, S. H., White, N. J., & Kim, T. H. (2016). PolySTAT-modified chitosan gauzes for improved hemostasis in external hemorrhage. Acta Biomaterialia, 31, 178–185. http://doi.org/10.1016/j.actbio.2015.11.017 Chan, L. W., Wang, X., Wei, H., Pozzo, L. D., White, N. J., & Pun, S. H. (2015). A synthetic fibrin cross-linking polymer for modulating clot properties and inducing hemostasis. Science Translational Medicine, 7(277), 1–11 (online). http://doi.org/10.1126/scitranslmed.3010383 Chanan-Khan, a. (2003). Complement activation following first exposure to pegylated liposomal doxorubicin (Doxil(R)): possible role in hypersensitivity reactions. Annals of Oncology, 14(9), 1430–1437. http://doi.org/10.1093/annonc/mdg374 Chandler, M. H., Roberts, M., Sawyer, M., & Myers, G. (2012). The US military experience with fresh whole blood during the conflicts in Iraq and Afghanistan. Seminars in Cardiothoracic and Vascular Anesthesia, 16(3), 153–9. http://doi.org/10.1177/1089253212452344 Chang, R., Cardenas, J. C., Wade, C. E., & Holcomb, J. B. (2016). Advances in the understanding of trauma-induced coagulopathy. Blood, 128(8), 1043–9. http://doi.org/10.1182/blood-2016-01-636423 Chao, F. C., Kim, B. K., Houranieh, A. M., Liang, F. H., Konrad, M. W., Swisher, S. N., & Tullis, J. L. (1996). Infusible platelet membrane microvesicles: a potential transfusion substitute for platelets. Transfusion, 36(6), 536–42. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/8669086 Chapin, J., & Hajjar, K. (2015). Fibrinolysis and the control of blood coagulation. Blood Reviews. Retrieved from https://www.sciencedirect.com/science/article/pii/S0268960X14000721 Chapman, W. C., Sherman, R., Boyce, S., Malawer, M., Hill, A., Buncke, G., … Goldsein, R. (2001). A novel collagen-based composite offers effective hemostasis for multiple surgical indications: Results of a randomized controlled trial. Surgery, 129(4), 445–450. http://doi.org/10.1067/MSY.2001.112365 Chaudry, I. H. (1983). Cellular mechanisms in shock and ischemia and their correction. The American Journal of Physiology, 245(2), R117-34. http://doi.org/10.1152/ajpregu.1983.245.2.R117 CHEN, J., & LÓPEZ, J. A. (2005). Interactions of Platelets with Subendothelium and Endothelium. Microcirculation, 12(3), 235–246. http://doi.org/10.1080/10739680590925484

335 Cheng, C. M., Meyer-Massetti, C., & Kayser, S. R. (2009). A review of three stand-alone topical thrombins for surgical hemostasis. Clinical Therapeutics, 31(1), 32–41. http://doi.org/10.1016/J.CLINTHERA.2009.01.005 Cheng, S., Craig, W. S., Mullen, D., Tschopp, J. F., Dixon, D., & Pierschbacher, M. D. (1994). Design and synthesis of novel cyclic RGD-containing peptides as highly potent and selective integrin alpha IIb beta 3 antagonists. Journal of Medicinal Chemistry, 37(1), 1–8. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/7507165 Chou, T.-C., Fu, E., Wu, C.-J., & Yeh, J.-H. (2003). Chitosan enhances platelet adhesion and aggregation. Biochemical and Biophysical Research Communications, 302(3), 480–483. http://doi.org/10.1016/S0006-291X(03)00173-6 Christenson, L. J., Otley, C. C., & Roenigk, R. K. (2004). Oxidized regenerated cellulose gauze for hemostasis of a two-stage interpolation flap pedicle. Dermatologic Surgery, 30(12 Pt 2), 1593–4. http://doi.org/10.1111/j.1524-4725.2004.30574.x Cierniewski, C. S., Byzova, T., Papierak, M., Haas, T. A., Niewiarowska, J., Zhang, L., … Plow, E. F. (1999). Peptide ligands can bind to distinct sites in integrin alphaIIbbeta3 and elicit different functional responses. The Journal of Biological Chemistry, 274(24), 16923–32. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10358039 Clark, H. (1929). The history of hemostasis. Retrieved from https://scholar.google.com/scholar_lookup?hl=en&volume=37&publication_year=19 29&journal=Semin+Hematol.&author=H.+S.+Clark&title=The+History+of+Hemost asis Coats, T. J., Brazil, E., & Heron, M. (2006). The effects of commonly used resuscitation fluids on whole blood coagulation. Emergency Medicine Journal, 23(7), 546–549. http://doi.org/10.1136/emj.2005.032334 Cohen, M. J. (2014). Acute traumatic coagulopathy: Clinical characterization and mechanistic investigation. Thrombosis Research, 133, S25–S27. http://doi.org/10.1016/J.THROMRES.2014.03.013 Cohen, M. J., Kutcher, M., Redick, B., Nelson, M., Call, M., Knudson, M. M., … PROMMTT Study Group, on behalf of the P. study. (2013). Clinical and mechanistic drivers of acute traumatic coagulopathy. The Journal of Trauma and Acute Care Surgery, 75(1 Suppl 1), S40-7. http://doi.org/10.1097/TA.0b013e31828fa43d Coller, B. (1980). Interaction of normal, thrombasthenic, and Bernard-Soulier platelets with immobilized fibrinogen: defective platelet-fibrinogen interaction in thrombasthenia. Blood, 55(2). Retrieved from http://www.bloodjournal.org/content/55/2/169.short Coller, B. S., Springer, K. T., Beer, J. H., Mohandas, N., Scudder, L. E., Norton, K. J., & West, S. M. (1992). Thromboerythrocytes. In vitro studies of a potential autologous, semi-artificial alternative to platelet transfusions. The Journal of Clinical Investigation, 89(2), 546–55. http://doi.org/10.1172/JCI115619

336 Connolly, R. J. (2004). Application of the Poly-N-Acetyl Glucosamine-Derived Rapid Deployment Hemostat Trauma Dressing in Severe/Lethal Swine Hemorrhage Trauma Models. The Journal of Trauma: Injury, Infection, and Critical Care, 57(Supplement), S26–S28. http://doi.org/10.1097/01.TA.0000136746.97192.B2 Cooley, M., Sarode, A., Hoore, M., Fedosov, D. A., Mitragotri, S., & Sen Gupta, A. (2018). Influence of particle size and shape on their margination and wall-adhesion: Implications in drug delivery vehicle design across nano-to-micro scale. Nanoscale. http://doi.org/10.1039/C8NR04042G Corash, L. (1999). Inactivation of viruses, bacteria, protozoa, and leukocytes in platelet concentrates: Current research perspectives. Transfusion Medicine Reviews, 13(1), 18–30. http://doi.org/10.1016/S0887-7963(99)80085-6 Cormier, A. R., Pang, X., Zimmerman, M. I., Zhou, H.-X., & Paravastu, A. K. (2013). Molecular Structure of RADA16-I Designer Self-Assembling Peptide Nanofibers. ACS Nano, 7(9), 7562–7572. http://doi.org/10.1021/nn401562f CORNUM, R., BELL, J., GRESHAM, V., BRINKLEY, W., BEALL, D., & MACPHEE, M. (1999). INTRAOPERATIVE USE OF THE ABSORBABLE FIBRIN ADHESIVE BANDAGE: LONG TERM EFFECTS. The Journal of Urology, 162(5), 1817–1820. http://doi.org/10.1016/S0022-5347(05)68244-4 Cotton, B. A., Gunter, O. L., Isbell, J., Au, B. K., Robertson, A. M., Morris, J. A., … Young, P. P. (2008). Damage Control Hematology: The Impact of a Trauma Exsanguination Protocol on Survival and Blood Product Utilization. The Journal of Trauma: Injury, Infection, and Critical Care, 64(5), 1177–1183. http://doi.org/10.1097/TA.0b013e31816c5c80 Cowley, R. (1975). A total emergency medical system for the State of Maryland. Md State Med J, 24(7), 37–45. Retrieved from https://www.ncbi.nlm.nih.gov/pubmed/1142842 CRASH-2 trial collaborators, H., Shakur, H., Roberts, I., Bautista, R., Caballero, J., Coats, T., … Yutthakasemsunt, S. (2010). Effects of tranexamic acid on death, vascular occlusive events, and blood transfusion in trauma patients with significant haemorrhage (CRASH-2): a randomised, placebo-controlled trial. Lancet (London, England), 376(9734), 23–32. http://doi.org/10.1016/S0140-6736(10)60835-5 Cresilon. (n.d.). Retrieved June 13, 2017, from https://cresilon.com/ CRONKITE, E. P., LOZNER, E. L., & DEAVER, J. M. (1944). USE OF THROMBIN AND FIBRINOGEN IN SKIN GRAFTING. Journal of the American Medical Association, 124(14), 976. http://doi.org/10.1001/jama.1944.02850140022006 Curry, N., Rourke, C., Davenport, R., Stanworth, S., & Brohi, K. (2014). Fibrinogen replacement in trauma haemorrhage. Scandinavian Journal of Trauma, Resuscitation and Emergency Medicine, 22(Suppl 1), A5. http://doi.org/10.1186/1757-7241-22-S1- A5 Dahlbäck, B., & Villoutreix, B. O. (2005). The anticoagulant protein C pathway. FEBS Letters, 579(15), 3310–3316. http://doi.org/10.1016/j.febslet.2005.03.001

337 Dan Dimitrijevich, S., Tatarko, M., Gracy, R. W., Linsky, C. B., & Olsen, C. (1990). Biodegradation of oxidized regenerated cellulose. Carbohydrate Research, 195(2), 247–256. http://doi.org/10.1016/0008-6215(90)84169-U Dan Dimitrijevich, S., Tatarko, M., Gracy, R. W., Wise, G. E., Oakford, L. X., Linsky, C. B., & Kamp, L. (1990). In vivo degradation of oxidized, regenerated cellulose. Carbohydrate Research, 198(2), 331–341. http://doi.org/10.1016/0008- 6215(90)84303-C Davenport, R. (2014). Haemorrhage control of the pre-hospital trauma patient. Scandinavian Journal of Trauma, Resuscitation and Emergency Medicine, 22(Suppl 1), A4. http://doi.org/10.1186/1757-7241-22-S1-A4 Davenport, R. A., Guerreiro, M., Frith, D., Rourke, C., Platton, S., Cohen, M., … Brohi, K. (2017). Activated Protein C Drives the Hyperfibrinolysis of Acute Traumatic Coagulopathy. Anesthesiology, 126(1), 115–127. http://doi.org/10.1097/ALN.0000000000001428 Davidson, B. R., Burnett, S., Javed, M. S., Seifalian, A., Moore, D., & Doctor, N. (2000). Experimental study of a novel fibrin sealant for achieving haemostasis following partial hepatectomy. British Journal of Surgery, 87(6), 790–795. http://doi.org/10.1046/j.1365-2168.2000.01427.x Davie, E. W., & Ratnoff, O. D. (1964). Waterfall Sequence for Intrinsic Blood Clotting. Science, 145(3638), 1310–1312. http://doi.org/10.1126/science.145.3638.1310 Davis, J. S., Satahoo, S. S., Butler, F. K., Dermer, H., Naranjo, D., Julien, K., … Schulman, C. I. (2014). An analysis of prehospital deaths: Who can we save? Accreditation Statement AMA PRA Category 1 Creditsi. http://doi.org/10.1097/TA.0000000000000292

De Ceunynck, K., De Meyer, S. F., & Vanhoorelbeke, K. (2013). Unwinding the von Willebrand factor strings puzzle. Blood, 121(2), 270–277. http://doi.org/10.1182/blood-2012-07-442285 del Carpio Munoz, C., Campbell, W., Constantinescu, I., & Gyongyossy-Issa, M. I. C. (2008). Rational design of antithrombotic peptides to target the von Willebrand Factor (vWf) - GPIb integrin interaction. Journal of Molecular Modeling, 14(12), 1191–1202. http://doi.org/10.1007/s00894-008-0375-z Delgado, A. V., Kheirabadi, B. S., Fruchterman, T. M., Scherer, M., Cortez, D., Wade, C. E., … Holcomb, J. B. (2008). A Novel Biologic Hemostatic Dressing (Fibrin Patch) Reduces Blood Loss and Resuscitation Volume and Improves Survival in Hypothermic, Coagulopathic Swine With Grade V Liver Injury. The Journal of Trauma: Injury, Infection, and Critical Care, 64(1), 75–80. http://doi.org/10.1097/TA.0b013e31815b843c Demetri, G. D. (2000). Pharmacologic treatment options in patients with thrombocytopenia. Seminars in Hematology, 37, 11–18. http://doi.org/10.1016/S0037-1963(00)90048-9

338 Devine, D. V., & Serrano, K. (2010). The Platelet Storage Lesion. Clinics in Laboratory Medicine, 30(2), 475–487. http://doi.org/10.1016/j.cll.2010.02.002 Dézsi, L., Fülöp, T., Mészáros, T., … G. S.-J. of controlled, & 2014, undefined. (n.d.). Features of complement activation-related pseudoallergy to liposomes with different surface charge and PEGylation: comparison of the porcine and rat. Elsevier. Retrieved from https://www.sciencedirect.com/science/article/pii/S0168365914005914 di Lena, F. (2014). Hemostatic polymers: the concept, state of the art and perspectives. J. Mater. Chem. B, 2(23), 3567–3577. http://doi.org/10.1039/C3TB21739F Dorlac, W. C., DeBakey, M. E., Holcomb, J. B., Fagan, S. P., Kwong, K. L., Dorlac, G. R., … Mattox, K. L. (2005). Mortality from isolated civilian penetrating extremity injury. The Journal of Trauma, 59(1), 217–22. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/16096567 Doshi, N., Orje, J. N., Molins, B., Smith, J. W., Mitragotri, S., & Ruggeri, Z. M. (2012). Platelet Mimetic Particles for Targeting Thrombi in Flowing Blood. Advanced Materials, 24(28), 3864–3869. http://doi.org/10.1002/adma.201200607 Dowling, M. B., Smith, W., Balogh, P., Duggan, M. J., MacIntire, I. C., Harris, E., … King, D. R. (2015). Hydrophobically-modified chitosan foam: description and hemostatic efficacy. Journal of Surgical Research, 193(1), 316–323. http://doi.org/10.1016/J.JSS.2014.06.019 Drake, D. B., & Wong, L. G. (2003). Hemostatic Effect of Vivostat Patient-Derived Fibrin Sealant on Split-Thickness Skin Graft Donor Sites. Annals of , 50(4), 367–372. http://doi.org/10.1097/01.SAP.0000041484.22953.6D Du, X. P., Plow, E. F., Frelinger, A. L., O’Toole, T. E., Loftus, J. C., & Ginsberg, M. H. (1991). Ligands "activate" integrin alpha IIb beta 3 (platelet GPIIb-IIIa). Cell, 65(3), 409–16. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/2018974

Duggan, M., Rago, A., Sharma, U., Zugates, G., Freyman, T., Busold, R., … King, L. D. R. (2013). Self-expanding polyurethane polymer improves survival in a model of noncompressible massive abdominal hemorrhage. Journal of Trauma and Acute Care Surgery, 74(6), 1462–1467. http://doi.org/10.1097/TA.0b013e31828da937 Dutton, R. P., Parr, M., Tortella, B. J., Champion, H. R., Bernard, G. R., Boffard, K., … Hauser, C. J. (2011). Recombinant Activated Factor VII Safety in Trauma Patients: Results From the CONTROL Trial. The Journal of Trauma: Injury, Infection, and Critical Care, 71(1), 12–19. http://doi.org/10.1097/TA.0b013e31821a42cf Dyer, M., Haldeman, S., Gutierrez, A., Kohut, L., Sen Gupta, A., & Neal, M. D. (2017). Uncontrolled Hemorrhagic Shock Modeled via Liver Laceration in Mice with Real Time Hemodynamic Monitoring. Journal of Visualized Experiments : JoVE, (123). http://doi.org/10.3791/55554 Dyer, M. R., Hickman, D., Luc, N., Haldeman, S., Loughran, P., Pawlwoski, C., … Neal, M. D. (2018). Intravenous administration of synthetic platelets (SynthoPlateTM) in a

339 mouse liver injury model of uncontrolled hemorrhage improves hemostasis. Journal of Trauma and Acute Care Surgery, 84(6), 1. http://doi.org/10.1097/TA.0000000000001893 Eastoe, J. E., & Leach, A. A. (Eds.). (1977). Chemical Constitution of Gelatin. London, UK: Academic Press. Edwards, J., Graves, E., Bopp, A., Prevost, N., Santiago, M., & Condon, B. (2014). Electrokinetic and Hemostatic Profiles of Nonwoven Cellulosic/Synthetic Fiber Blends with Unbleached Cotton. Journal of Functional Biomaterials, 5(4), 273–287. http://doi.org/10.3390/jfb5040273 Edwards, J. V., & Prevost, N. (2011). Thrombin production and human neutrophil elastase sequestration by modified cellulosic dressings and their electrokinetic analysis. Journal of Functional Biomaterials, 2(4), 391–413. http://doi.org/10.3390/jfb2040391 Elliott, J. T., & Prestwich, G. D. (2000). Maleimide-Functionalized Lipids that Anchor Polypeptides to Lipid Bilayers and Membranes. Bioconjugate Chemistry, 11(6), 832– 841. http://doi.org/10.1021/bc000022a Eltzschig, H. K., & Eckle, T. (2011). Ischemia and reperfusion--from mechanism to translation. Nature Medicine, 17(11), 1391–401. http://doi.org/10.1038/nm.2507 Englehart, M. S., Cho, S. D., Tieu, B. H., Morris, M. S., Underwood, S. J., Karahan, A., … Schreiber, M. A. (2008). A Novel Highly Porous Silica and Chitosan-Based Hemostatic Dressing Is Superior to HemCon and Gauze Sponges. The Journal of Trauma: Injury, Infection, and Critical Care, 65(4), 884–892. http://doi.org/10.1097/TA.0b013e318187800b Epstein, J. S., & Vostal, J. G. (2003). FDA approach to evaluation of pathogen reduction technology. Transfusion, 43(10), 1347–1350. http://doi.org/10.1046/j.1537- 2995.2003.00584.x Erdogan, D., & van Gulik, T. M. (2008). Evolution of fibrinogen-coated collagen patch for use as a topical hemostatic agent. Journal of Biomedical Materials Research Part B: Applied Biomaterials, 85B(1), 272–278. http://doi.org/10.1002/jbm.b.30916 Ersoy, G., Kaynak, M. F., Yilmaz, O., Rodoplu, U., Maltepe, F., & Gokmen, N. (2007). Hemostatic effects of Microporous Polysaccharide Hemosphere® in a rat model with severe femoral artery bleeding. Advances in Therapy, 24(3), 485–492. http://doi.org/10.1007/BF02848770 Etchill, E. W., Myers, S. P., Raval, J. S., Hassoune, A., SenGupta, A., & Neal, M. D. (2017). Platelet Transfusion in Critical Care and Surgery. SHOCK, 47(5), 537–549. http://doi.org/10.1097/SHK.0000000000000794 EVATT, B. L., FARRUGIA, A., SHAPIRO, A. D., & WILDE, J. T. (2002). Haemophilia 2002: emerging risks of treatment. Haemophilia, 8(3), 221–229. http://doi.org/10.1046/j.1365-2516.2002.00612.x

340 Eyre, L., & Gamlin, F. (2010). Haemostasis, blood platelets and coagulation. Anaesthesia & , 11(6), 244–246. http://doi.org/10.1016/J.MPAIC.2010.02.015 FAIRBAIRN, J. W., & WHITTET, T. D. (1948). Absorbable haemostatics, their uses and identification. The Pharmaceutical Journal, 106(4400), 149. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/18911857 Fan, Y., Sun, H., Pei, G., & Ruan, C. (2008). Haemostatic efficacy of an ethyl-2- cyanoacrylate-based aerosol in combination with tourniquet application in a large wound model with an arterial injury. Injury, 39(1), 61–66. http://doi.org/10.1016/J.INJURY.2007.10.004 Farndale, R. W., Sixma, J. J., Barnes, M. J., & De Groot, P. G. (2004). The role of collagen in thrombosis and hemostasis. Journal of Thrombosis and Haemostasis, 2(4), 561–573. http://doi.org/10.1111/j.1538-7836.2004.00665.x Fedel, M., Endogan, T., Hasirci, N., Maniglio, D., Morelli, A., Chiellini, F., & Motta, A. (2012). Blood compatibility of polymers derived from natural materials. Journal of Bioactive and Compatible Polymers, 27(4), 295–312. http://doi.org/10.1177/0883911512446060 FINLEY, D. S., LEE, D. I., EICHEL, L., URIBE, C. A., McDOUGALL, E. M., & CLAYMAN, R. V. (2005). FIBRIN GLUE-OXIDIZED CELLULOSE SANDWICH FOR LAPAROSCOPIC WEDGE RESECTION OF SMALL RENAL LESIONS. The Journal of Urology, 173(5), 1477–1481. http://doi.org/10.1097/01.JU.0000154165.12738.7F Fischer, D., Li, Y., Ahlemeyer, B., Krieglstein, J., & Kissel, T. (2003). In vitro cytotoxicity testing of polycations: influence of polymer structure on cell viability and hemolysis. Biomaterials, 24(7), 1121–1131. http://doi.org/10.1016/S0142- 9612(02)00445-3 Fischer, T. H., Bode, A. P., Demcheva, M., & Vournakis, J. N. (2007). Hemostatic properties of glucosamine-based materials. Journal of Biomedical Materials Research Part A, 80A(1), 167–174. http://doi.org/10.1002/jbm.a.30877 Fitzpatrick, G. M., Cliff, R., & Tandon, N. (2013). Thrombosomes: a platelet-derived hemostatic agent for control of noncompressible hemorrhage. Transfusion, 53 Suppl 1, 100S–106S. http://doi.org/10.1111/trf.12043 Franchini, M. (2007). The use of desmopressin as a hemostatic agent: A concise review. American Journal of Hematology, 82(8), 731–735. http://doi.org/10.1002/ajh.20940 Frantz, V. K. (1943). ABSORBABLE COTTON, PAPER AND GAUZE : (OXIDIZED CELLULOSE). Annals of Surgery, 118(1), 116–26. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/17858245 FRANTZ, V. K., & LATTES, R. (1945). OXIDIZED CELLULOSE—ABSORBABLE GAUZE. Journal of the American Medical Association, 129(12), 798. http://doi.org/10.1001/jama.1945.02860460022006

341 Fries, D., & Martini, W. Z. (2010). Role of fibrinogen in trauma-induced coagulopathy. British Journal of Anaesthesia, 105(2), 116–121. http://doi.org/10.1093/bja/aeq161 Friess, W. (1998). Collagen – biomaterial for drug delivery. European Journal of Pharmaceutics and Biopharmaceutics, 45(2), 113–136. http://doi.org/10.1016/S0939- 6411(98)00017-4 Fukunaga, N., Uchida, T., Kaetsu, H., Kawakami, K., Ishihara, Y., & Funatsu, A. (1998). Effective application of fibrin sealant using a spray device with a double-tube structure. International Journal of Adhesion and Adhesives, 18(5), 345–349. http://doi.org/10.1016/S0143-7496(98)00016-5 Fülöp, A., Turóczi, Z., Garbaisz, D., Harsányi, L., & Szijártó, A. (2013). Experimental models of hemorrhagic shock: A review. European Surgical Research, 50(2), 57–70. http://doi.org/10.1159/000348808 Furie, M. B., Atkins, M., & Mayer, R. (2003). Clinical Hematology and Oncology: Presentation, Diagnosis and Treatment, 1e. Retrieved from https://mrz3jdqtk01.storage.googleapis.com/MDQ0MzA2NTU2WA==01.pdf Fuss, C., Palmaz, J. C., & Sprague, E. A. (2001). Fibrinogen: Structure, Function, and Surface Interactions. Journal of Vascular and Interventional , 12(6), 677– 682. http://doi.org/10.1016/S1051-0443(07)61437-7 Gabay, M., & Boucher, B. A. (2013). An Essential Primer for Understanding the Role of Topical Hemostats, Surgical Sealants, and Adhesives for Maintaining Hemostasis. Pharmacotherapy: The Journal of Human Pharmacology and Drug Therapy, 33(9), 935–955. http://doi.org/10.1002/phar.1291 Gaharwar, A. K., Avery, R. K., Assmann, A., Paul, A., McKinley, G. H., Khademhosseini, A., & Olsen, B. D. (2014). Shear-Thinning Nanocomposite Hydrogels for the Treatment of Hemorrhage. ACS Nano, 8(10), 9833–9842. http://doi.org/10.1021/nn503719n Galán, A. M., Bozzo, J., Hernández, M. R., Pino, M., Reverter, J. C., Mazzara, R., … Ordinas, A. (2000). Infusible platelet membranes improve hemostasis in thrombocytopenic blood: experimental studies under flow conditions. Transfusion, 40(9), 1074–80. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10988310 Galan, A. M., Lozano, M., Molina, P., Navalon, F., Marschner, S., Goodrich, R., & Escolar, G. (2011). Impact of pathogen reduction technology and storage in platelet additive solutions on platelet function. Transfusion, 51(4), 808–815. http://doi.org/10.1111/j.1537-2995.2010.02914.x Gale, A. J. (2011). Continuing Education Course #2: Current Understanding of Hemostasis. Toxicologic Pathology, 39(1), 273–280. http://doi.org/10.1177/0192623310389474 Garraud, O., Cognasse, F., Tissot, J.-D., Chavarin, P., Laperche, S., Morel, P., … Osselaer, J.-C. (2016). Improving platelet transfusion safety: biomedical and technical considerations. Blood Transfusion, 14(2), 109–22. http://doi.org/10.2450/2015.0042-15

342 Gegel, B., Burgert, J., Gasko, J., Campbell, C., Martens, M., Keck, J., … Johnson, D. (2012). The Effects of QuikClot Combat Gauze and Movement on Hemorrhage Control in a Porcine Model. Military Medicine, 177(12), 1543–1547. http://doi.org/10.7205/MILMED-D-12-00165 Gegel, B. T., Austin, P. N., & Johnson, A. D. (2013). An evidence-based review of the use of a combat gauze (QuikClot) for hemorrhage control. AANA Journal, 81(6), 453–8. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/24597007 Gegel, B. T., Burgert, J. M., Lockhart, C., Austin, R., Davila, A., Deeds, J., … Johnson, D. (2010). Effects of Celox and TraumaDEX on hemorrhage control in a porcine model. AANA Journal, 78(2), 115–20. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/20583456 Gelse, K., Pöschl, E., & Aigner, T. (2003). Collagens—structure, function, and biosynthesis. Advanced Drug Delivery Reviews, 55(12), 1531–1546. http://doi.org/10.1016/J.ADDR.2003.08.002 Gerotziafas, G. T., Zervas, C., Gavrielidis, G., Tokmaktsis, A., Hatjiharissi, E., Papaioannou, M., … Christakis, J. (2002). Effective hemostasis with rFVIIa treatment in two patients with severe thrombocytopenia and life-threatening hemorrhage. American Journal of Hematology, 69(3), 219–222. http://doi.org/10.1002/ajh.10056 Gillespie, C. J., Sutherland, A. D., Mossop, P. J., Woods, R. J., Keck, J. O., & Heriot, A. G. (2010). Mesenteric Embolization for Lower Gastrointestinal Bleeding. Diseases of the Colon & Rectum, 53(9), 1258–1264. http://doi.org/10.1007/DCR.0b013e3181e10e90 Gils, A., & Declerck, P. J. (2004). Plasminogen activator inhibitor-1. Current Medicinal Chemistry, 11(17), 2323–34. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/15379715 Ginsburg, I., de Vries, A., & Katchalski, E. (1952). The Action of Some Water-Soluble Poly-a-Amino Acids on Fibrinolysis. Science, 116(3001), 15–6. http://doi.org/10.1126/science.116.3001.15 Golling, M., Schaudt, A., Mehrabi, A., Mood, Z. A., & Bechstein, W. O. (2008). Clinical application of soft polyglycolic acid felt for hemostasis and repair of a lacerated liver: Report of two cases. Surgery Today, 38(2), 188–192. http://doi.org/10.1007/s00595- 007-3587-4 Goodnough, L. T., Brecher, M. E., Kanter, M. H., & AuBuchon, J. P. (1999). Transfusion Medicine — Blood Transfusion. New England Journal of Medicine, 340(6), 438–447. http://doi.org/10.1056/NEJM199902113400606 Graham, S. S., Gonchoroff, N. J., & Miller, J. L. (2001). Infusible Platelet Membranes Retain Partial Functionality of the Platelet GPIb/IX/V Receptor Complex. American Journal of , 115(1), 144–147. http://doi.org/10.1309/CCDV-3BEP- XXKP-BKDM Greenberg, E. M. (2017). Thrombocytopenia: A Destruction of Platelets. Journal of Infusion , 40(1), 41–50. http://doi.org/10.1097/NAN.0000000000000204

343 Greinacher, A., Selleng, K., & Warkentin, T. E. (2017). Autoimmune heparin-induced thrombocytopenia. Journal of Thrombosis and Haemostasis, 15(11), 2099–2114. http://doi.org/10.1111/jth.13813 Grey, E. (1915). Fibrin as a hemostatic in cerebral surgery. Surg Gynecol Obstet, (21), 452. GRIFFITH, B. C., MOREY, A. F., ROZANSKI, T. A., HARRIS, R., DALTON, S. R., TORGERSON, S. J., & PARTYKA, S. R. (2004). Central Renal Stab Wounds:: Treatment With Augmented Fibrin Sealant in a Porcine Model. The Journal of Urology, 171(1), 445–447. http://doi.org/10.1097/01.JU.0000101000.08371.79 Gryshchuk, V., & Galagan, N. (2016). Silica Nanoparticles Effects on Blood Coagulation Proteins and Platelets. Biochemistry Research International, 2016, 2959414. http://doi.org/10.1155/2016/2959414 Gurney, J. M., & Spinella, P. C. (2018). Blood transfusion management in the severely bleeding military patient. Current Opinion in Anaesthesiology, 31(2), 1. http://doi.org/10.1097/ACO.0000000000000574 Gustafson, S. B., Fulkerson, P., Bildfell, R., Aguilera, L., & Hazzard, T. M. (2007). Chitosan Dressing Provides Hemostasis in Swine Femoral Arterial Injury Model. Prehospital Emergency Care, 11(2), 172–178. http://doi.org/10.1080/10903120701205893 Gyongyossy‐Issa, M. I. C., Kizhakkedathu, J., Constantinescu, I., Campbell, W., & del Carpio Munoz, C. A. (2007, September 7). Synthetic platelets. US. Retrieved from https://patents.google.com/patent/CA2599849C/en?oq=Gyongyossy‐ Issa%2C+Kizhakkedathu%2C+Constantinescu%2C+Campbell%2C+del+Carpio+Mu noz Hagen, F. S., Gray, C. L., O’Hara, P., Grant, F. J., Saari, G. C., Woodbury, R. G., … Kurachi, K. (1986). Characterization of a cDNA coding for human factor VII. Proceedings of the National Academy of Sciences of the United States of America, 83(8), 2412–6. http://doi.org/10.1073/PNAS.83.8.2412 Hait, M. R. (1970). Microcrystalline collagen. A new hemostatic agent. American Journal of Surgery, 120(3), 330. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/5456913 Hait, M. R., Robb, C. A., Baxter, C. R., Borgmann, A. R., & Tippett, L. O. (1973). Comparative evaluation of avitene microcrystalline collagen hemostat in experimental animal wounds. The American Journal of Surgery, 125(3), 284–287. http://doi.org/10.1016/0002-9610(73)90042-1 Ham, S. W., Lew, W. K., & Weaver, F. A. (2010). Thrombin use in surgery: an evidence- based review of its clinical use. Journal of Blood Medicine, 1, 135–42. http://doi.org/10.2147/JBM.S6622 Harker, L. A., & Finch, C. A. (1969). Thrombokinetics in man. The Journal of Clinical Investigation, 48(6), 963–74. http://doi.org/10.1172/JCI106077

344 Harmon, P., Cabral-Lilly, D., Reed, R. A., Maurio, F. P., Franklin, J. C., & Janoff, A. (1997). The Release and Detection of Endotoxin from Liposomes. Analytical Biochemistry, 250(2), 139–146. http://doi.org/10.1006/abio.1997.2216 Harvey, S. (1916). The use of fibrin paper and forms in surgery. Boston Med Surg J, (174), 658. Hattori, H., Amano, Y., Nogami, Y., Takase, B., & Ishihara, M. (2010). Hemostasis for Severe Hemorrhage with Photocrosslinkable Chitosan Hydrogel and Calcium Alginate. Annals of Biomedical Engineering, 38(12), 3724–3732. http://doi.org/10.1007/s10439-010-0121-4 Heal, J. M., & Blumberg, N. (2004). Optimizing platelet transfusion therapy. Blood Reviews, 18(3), 149–165. http://doi.org/10.1016/S0268-960X(03)00057-2 Heddle, N. M., Klama, L. N., Griffith, L., Roberts, R., Shukla, G., & Kelton, J. G. (1993). A prospective study to identify the risk factors associated with acute reactions to platelet and red cell transfusions. Transfusion, 33(10), 794–7. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/8236418 Hedner, U. (2004). Dosing with recombinant factor viia based on current evidence. Seminars in Hematology, 41(1 Suppl 1), 35–39. http://doi.org/10.1053/j.seminhematol.2003.11.008 Hedner, U., & Lee, C. A. (2011). First 20 years with recombinant FVIIa (NovoSeven). Haemophilia, 17(1), 172–182. http://doi.org/10.1111/j.1365-2516.2010.02352.x Heemskerk, J., Bevers, E., & Lindhout, T. (2002). Platelet Activation and Blood Coagulation. Thrombosis and Haemostasis, 88(08), 186–193. http://doi.org/10.1055/s-0037-1613209 HEEMSKERK, J. W. M., MATTHEIJ, N. J. A., & COSEMANS, J. M. E. M. (2013). Platelet-based coagulation: different populations, different functions. Journal of Thrombosis and Haemostasis, 11(1), 2–16. http://doi.org/10.1111/jth.12045 Henderson, P. W., Kadouch, D. J. M., Singh, S. P., Zawaneh, P. N., Weiser, J., Yazdi, S., … Spector, J. A. (2009). A rapidly resorbable hemostatic biomaterial based on dihydroxyacetone. Journal of Biomedical Materials Research Part A, 9999A, NA- NA. http://doi.org/10.1002/jbm.a.32586 Herr, A. B., & Farndale, R. W. (2009). Structural insights into the interactions between platelet receptors and fibrillar collagen. The Journal of Biological Chemistry, 284(30), 19781–5. http://doi.org/10.1074/jbc.R109.013219 Hess, J. R. (2013). Resuscitation of trauma-induced coagulopathy. Hematology Am Soc Hematol Educ Program., 664–7. http://doi.org/10.1182/asheducation-2013.1.664 Hess, J. R., Brohi, K., Dutton, R. P., Hauser, C. J., Holcomb, J. B., Kluger, Y., … Bouillon, B. (2008). The Coagulopathy of Trauma: A Review of Mechanisms. The Journal of Trauma: Injury, Infection, and Critical Care, 65(4), 748–754. http://doi.org/10.1097/TA.0b013e3181877a9c

345 Hickman, D. A., Pawlowski, C. L., Sekhon, U. D. S., Marks, J., & Gupta, A. Sen. (2017). Biomaterials and Advanced Technologies for Hemostatic Management of Bleeding. Advanced Materials, 1700859, 1–40. http://doi.org/10.1002/adma.201700859 Hickman, D. A., Pawlowski, C. L., Shevitz, A., Luc, N. F., Kim, A., Girish, A., … Neal, M. D. (2018). Intravenous synthetic platelet ( SynthoPlate ) nanoconstructs reduce bleeding and improve ‘ golden hour ’ survival in a porcine model of traumatic arterial hemorrhage. Scientific Reports, (February), 1–14. http://doi.org/10.1038/s41598-018- 21384-z Hill, A., Estridge, T. D., Maroney, M., Monnet, E., Egbert, B., Cruise, G., & Coker, G. T. (2001). Treatment of suture line bleeding with a novel synthetic surgical sealant in a canine iliac PTFE graft model. Journal of Biomedical Materials Research, 58(3), 308–312. http://doi.org/10.1002/1097-4636(2001)58:3<308::AID- JBM1022>3.0.CO;2-P Hoekstra, M. J., Hermans, M. H. E., Richters, C. D., & Dutrieux, R. P. (2002). A histological comparison of acute inflammatory responses with a hydrofibre or tulle gauze dressing. Journal of Wound Care, 11(3), 113–117. http://doi.org/10.12968/jowc.2002.11.3.26384 Hoffman, M. (2003). Remodeling the Blood Coagulation Cascade. Journal of Thrombosis and Thrombolysis, 16(1/2), 17–20. http://doi.org/10.1023/B:THRO.0000014588.95061.28 Hoffman, M., & Monroe, D. M. 3rd. (2001). A cell-based model of hemostasis. Thrombosis and Haemostasis, 85(6), 958–965. http://doi.org/01060958 [pii] Hoffmeister, K. M., Josefsson, E. C., Isaac, N. A., Clausen, H., Hartwig, J. H., & Stossel, T. P. (2003). Glycosylation restores survival of chilled blood platelets. Science (New York, N.Y.), 301(5639), 1531–4. http://doi.org/10.1126/science.1085322 Holcomb, J. B. (2010). Optimal use of blood products in severely injured trauma patients. Hematology. American Society of Hematology. Education Program, 2010(1), 465–9. http://doi.org/10.1182/asheducation-2010.1.465 Holcomb, J. B., del Junco, D. J., Fox, E. E., Wade, C. E., Cohen, M. J., Schreiber, M. A., … PROMMTT Study Group. (2013). The prospective, observational, multicenter, major trauma transfusion (PROMMTT) study: comparative effectiveness of a time- varying treatment with competing risks. JAMA Surgery, 148(2), 127–36. http://doi.org/10.1001/2013.jamasurg.387 Holcomb, J. B., Jenkins, D., Rhee, P., Johannigman, J., Mahoney, P., Mehta, S., … Hess, J. R. (2007). Damage Control Resuscitation: Directly Addressing the Early Coagulopathy of Trauma. The Journal of Trauma: Injury, Infection, and Critical Care, 62(2), 307–310. http://doi.org/10.1097/TA.0b013e3180324124 Holcomb, J. B., McClain, J. M., Pusateri, A. E., Beall, D., Macaitis, J. M., Harris, R. A., … Hess, J. R. (2000). Fibrin sealant foam sprayed directly on liver injuries decreases blood loss in resuscitated rats. Journal of Trauma - Injury, Infection and Critical Care, 49(2), 246–250. http://doi.org/10.1097/00005373-200008000-00010

346 Holcomb, J. B., McMullin, N. R., Pearse, L., Caruso, J., Wade, C. E., Oetjen-Gerdes, L., … Butler, F. K. (2007). Causes of Death in U.S. Special Operations Forces in the Global War on Terrorism. Annals of Surgery, 245(6), 986–991. http://doi.org/10.1097/01.sla.0000259433.03754.98 Holcomb, J. B., Pusateri, A. E., Harris, R. A., Charles, N. C., Gomez, R. R., Cole, J. P., … Hess, J. R. (1999). Effect of Dry Fibrin Sealant Dressings versus Gauze Packing on Blood Loss in Grade V Liver Injuries in Resuscitated Swine. Journal of Trauma and Acute Care Surgery, 46(1). Retrieved from https://journals.lww.com/jtrauma/Fulltext/1999/01000/Effect_of_Dry_Fibrin_Sealant _Dressings_versus.9.aspx Holcomb, J. B., Tilley, B. C., Baraniuk, S., Fox, E. E., Wade, C. E., Podbielski, J. M., … van Belle, G. (2015). Transfusion of Plasma, Platelets, and Red Blood Cells in a 1:1:1 vs a 1:1:2 Ratio and Mortality in Patients With Severe Trauma. Jama, 313, 471. http://doi.org/10.1001/jama.2015.12 Holcomb, J., MacPhee, M., Hetz, S., Harris, R., Pusateri, A., & Hess, J. (1998). Efficacy of a Dry Fibrin Sealant Dressing for Hemorrhage Control After Ballistic Injury. Archives of Surgery, 133(1), 32. http://doi.org/10.1001/archsurg.133.1.32 Horák, D., Švec, F., Isakov, Y., Polyaev, Y., Adamyan, A., Konstantinov, K., … Trostenyuk, N. (1992). Use of poly(2-hydroxyethyl methacrylate) for endovascular occlusion in pediatric surgery. Clinical Materials, 9(1), 43–48. http://doi.org/10.1016/0267-6605(92)90009-I Houdebine, L.-M. (2009). Production of pharmaceutical proteins by transgenic animals. Comparative , Microbiology and Infectious Diseases, 32(2), 107–121. http://doi.org/10.1016/J.CIMID.2007.11.005 Hoylaerts, M., Rijken, D. C., Lijnen, H. R., & Collen, D. (1982). Kinetics of the activation of plasminogen by human tissue plasminogen activator. Role of fibrin. The Journal of Biological Chemistry, 257(6), 2912–9. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/7199524 Hsu, B. B., Conway, W., Tschabrunn, C. M., Mehta, M., Perez-Cuevas, M. B., Zhang, S., & Hammond, P. T. (2015). Clotting Mimicry from Robust Hemostatic Bandages Based on Self-Assembling Peptides. ACS Nano, 9(9), 9394–9406. http://doi.org/10.1021/acsnano.5b02374 Hu, C.-M. J., Fang, R. H., Wang, K.-C., Luk, B. T., Thamphiwatana, S., Dehaini, D., … Zhang, L. (2015). Nanoparticle biointerfacing by platelet membrane cloaking. Nature, 526(7571), 118–121. http://doi.org/10.1038/nature15373 Hu, Q., Sun, W., Qian, C., Wang, C., Bomba, H. N., & Gu, Z. (2015). Anticancer Platelet-Mimicking Nanovehicles. Advanced Materials, 27(44), 7043–7050. http://doi.org/10.1002/adma.201503323 Huang, G., Zhou, Z., Srinivasan, R., Penn, M. S., Kottke-Marchant, K., Marchant, R. E., & Gupta, A. S. (2008). Affinity manipulation of surface-conjugated RGD peptide to modulate binding of liposomes to activated platelets. Biomaterials, 29(11), 1676–85. http://doi.org/10.1016/j.biomaterials.2007.12.015

347 Humphries, J. D., Byron, A., & Humphries, M. J. (2006). Integrin ligands at a glance. Journal of Cell Science, 119(Pt 19), 3901–3. http://doi.org/10.1242/jcs.03098 Hunt, B. J. (2015). The current place of tranexamic acid in the management of bleeding. Anaesthesia, 70, 50-e18. http://doi.org/10.1111/anae.12910 Hunt, H., Stanworth, S., Curry, N., Woolley, T., Cooper, C., Ukoumunne, O., … Hyde, C. (2015). Thromboelastography (TEG) and rotational thromboelastometry (ROTEM) for trauma-induced coagulopathy in adult trauma patients with bleeding. Cochrane Database of Systematic Reviews, (2). http://doi.org/10.1002/14651858.CD010438.pub2 Immordino, M. L., Dosio, F., & Cattel, L. (2006). Stealth liposomes: review of the basic science, rationale, and clinical applications, existing and potential. International Journal of Nanomedicine, 1(3), 297–315. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/17717971 Inaba, K., Rhee, P., Teixeira, P. G., Barmparas, G., Putty, B., Branco, B. C., … Demetriades, D. (2011). Intracorporeal use of advanced local hemostatics in a damage control swine model of grade IV liver injury. The Journal of Trauma, 71(5), 1312–8. http://doi.org/10.1097/TA.0b013e31821cb7cd Initial management of trauma in adults - UpToDate. (n.d.). Retrieved August 4, 2018, from https://www.uptodate.com/contents/initial-management-of-trauma-in- adults?search=trauma&source=search_result&selectedTitle=1~150&usage_type=def ault&display_rank=1 Ishikura, H., & Kitamura, T. (2017). Trauma-induced coagulopathy and critical bleeding: the role of plasma and platelet transfusion. Journal of Intensive Care, 5(1), 2. http://doi.org/10.1186/s40560-016-0203-y Ivanova, E. P., Bazaka, K., & Crawford, R. J. (n.d.). New functional biomaterials for medicine and healthcare. Izumi, Y., Gika, M., Shinya, N., Miyabashira, S., Imamura, T., Nozaki, C., … Kobayashi, K. (2007). Hemostatic Efficacy of a Recombinant Thrombin-Coated Polyglycolic Acid Sheet Coupled With Liquid Fibrinogen, Evaluated in a Canine Model of Pulmonary Arterial Hemorrhage. The Journal of Trauma: Injury, Infection, and Critical Care, 63(4), 783–787. http://doi.org/10.1097/TA.0b013e318151ffdc Jackson, M. R. (2001). Fibrin sealants in surgical practice: An overview. The American Journal of Surgery, 182(2), S1–S7. http://doi.org/10.1016/S0002-9610(01)00770-X Jackson, M. R., Friedman, S. A., Carter, A. J., Bayer, V., Burge, J. R., MacPhee, M. J., … Alving, B. M. (1997). Hemostatic efficacy of a fibrin sealant–based topical agent in a femoral artery injury model: A randomized, blinded, placebo-controlled study. Journal of Vascular Surgery, 26(2), 274–280. http://doi.org/10.1016/S0741- 5214(97)70189-7 Jackson, M. R., Taher, M. M., Burge, J. R., Krishnamurti, C., Reid, T. J., & Alving, B. M. (1998). Hemostatic Efficacy of a Fibrin Sealant Dressing in an Animal Model of Kidney Injury. Journal of Trauma and Acute Care Surgery, 45(4). Retrieved from

348 https://journals.lww.com/jtrauma/Fulltext/1998/10000/Hemostatic_Efficacy_of_a_Fi brin_Sealant_Dressing.3.aspx Janetzko, K., Hinz, K., Marschner, S., Goodrich, R., & Klüter, H. (2009). Pathogen reduction technology (Mirasol®) treated single-donor platelets resuspended in a mixture of autologous plasma and PAS. Vox Sanguinis, 97(3), 234–239. http://doi.org/10.1111/j.1423-0410.2009.01193.x Jen, C. J., & McIntire, L. V. (1982). The structural properties and contractile force of a clot. Cell Motility, 2(5), 445–455. http://doi.org/10.1002/cm.970020504 Jenkins, H., Senz, E., Owen, H., & Jampolis, R. (1946). Present status of gelatin sponge for the control of hemorrhage: With experimental data on its use for wounds of the great vessels and the heart. J. Am. Med. Assoc., 132, 614. Retrieved from https://jamanetwork.com/data/journals/JAMA/7264/jama_132_11_002.pdf Jewelewicz, D. D., Cohn, S. M., Crookes, B. A., & Proctor, K. G. (2003). Modified Rapid Deployment Hemostat Bandage Reduces Blood Loss and Mortality in Coagulopathic Pigs with Severe Liver Injury. The Journal of Trauma: Injury, Infection, and Critical Care, 55(2), 275–281. http://doi.org/10.1097/01.TA.0000079375.69610.89 Jhang, J. S., & Spitalnik, S. L. (2005). Glycosylation and cold platelet storage. Current Hematology Reports, 4(6), 483–7. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/16232387 Ji, Y., Ji, T., Liang, K., & Zhu, L. (2016). Mussel-inspired soft-tissue adhesive based on poly(diol citrate) with catechol functionality. Journal of Materials Science: Materials in Medicine, 27(2), 30. http://doi.org/10.1007/s10856-015-5649-2 Joch, C. (2003). The safety of fibrin sealants. Cardiovascular Surgery, 11, 23–28. http://doi.org/10.1016/S0967-2109(03)00068-1 Johansson, P. I., Stensballe, J., Oliveri, R., Wade, C. E., Ostrowski, S. R., & Holcomb, J. B. (2014). How I treat patients with massive hemorrhage. Blood, 124(20), 3052– 3058. http://doi.org/10.1182/blood-2014-05-575340 Johnson, D., Bates, S., Nukalo, S., Staub, A., Hines, A., Leishman, T., … Burgert, J. (2014). The effects of QuikClot Combat Gauze on hemorrhage control in the presence of hemodilution and hypothermia. Annals of Medicine and Surgery, 3(2), 21–25. http://doi.org/10.1016/j.amsu.2014.03.001 Kabiri, M., Emami, S. H., Rafinia, M., & Tahriri, M. (2011). Preparation and characterization of absorbable hemostat crosslinked gelatin sponges for surgical applications. Current Applied Physics, 11(3), 457–461. http://doi.org/10.1016/J.CAP.2010.08.031 Kattula, S., Byrnes, J. R., & Wolberg, A. S. (2017). Fibrinogen and Fibrin in Hemostasis and Thrombosis. Arteriosclerosis, Thrombosis, and Vascular Biology, 37(3), e13– e21. http://doi.org/10.1161/ATVBAHA.117.308564 Kaufman, R. J., Wasley, L. C., Furie, B. C., Furie, B., & Shoemaker, C. B. (1986). Expression, purification, and characterization of recombinant gamma-carboxylated

349 factor IX synthesized in Chinese hamster ovary cells. The Journal of Biological Chemistry, 261(21), 9622–8. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/3733688 Kaufman, R. M., Djulbegovic, B., Gernsheimer, T., Kleinman, S., Tinmouth, A. T., Capocelli, K. E., … Tobian, A. A. R. (2015). Platelet Transfusion: A Clinical Practice Guideline From the AABB. Annals of , 162(3), 205. http://doi.org/10.7326/M14-1589 Kauvar, D. S., Holcomb, J. B., Norris, G. C., & Hess, J. R. (2006). Fresh Whole Blood Transfusion: A Controversial Military Practice. The Journal of Trauma: Injury, Infection, and Critical Care, 61(1), 181–184. http://doi.org/10.1097/01.ta.0000222671.84335.64 Kauvar, D. S., Lefering, R., & Wade, C. E. (2006). Impact of Hemorrhage on Trauma Outcome: An Overview of Epidemiology, Clinical Presentations, and Therapeutic Considerations. The Journal of Trauma: Injury, Infection, and Critical Care, 60(Supplement), S3–S11. http://doi.org/10.1097/01.ta.0000199961.02677.19 Ker, K., Edwards, P., Perel, P., Shakur, H., & Roberts, I. (2012). Effect of tranexamic acid on surgical bleeding: systematic review and cumulative meta-analysis. BMJ (Clinical Research Ed.), 344, e3054. http://doi.org/10.1136/BMJ.E3054 Kheirabadi, B. (2011). Evaluation of topical hemostatic agents for combat wound treatment. U.S. Army Medical Department Journal, 25–37. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/21607904 Kheirabadi, B. S., Acheson, E. M., Deguzman, R., Crissey, J. M., Delgado, A. V., Estep, S. J., & Holcomb, J. B. (2007). The Potential Utility of Fibrin Sealant Dressing in Repair of Vascular Injury in Swine. The Journal of Trauma: Injury, Infection, and Critical Care, 62(1), 94–103. http://doi.org/10.1097/01.ta.0000251595.45451.d0 Kheirabadi, B. S., Edens, J. W., Terrazas, I. B., Estep, J. S., Klemcke, H. G., Dubick, M. A., & Holcomb, J. B. (2009). Comparison of New Hemostatic Granules/Powders With Currently Deployed Hemostatic Products in a Lethal Model of Extremity Arterial Hemorrhage in Swine. The Journal of Trauma: Injury, Infection, and Critical Care, 66(2), 316–328. http://doi.org/10.1097/TA.0b013e31819634a1 Kheirabadi, B. S., & Klemcke, H. G. (2004). Symp. Paper RTO‐MP‐HFM‐109. In NATO Science and Technology Organization (p. 20). Kheirabadi, B. S., Mace, J. E., Terrazas, I. B., Fedyk, C. G., Estep, J. S., Dubick, M. A., & Blackbourne, L. H. (2010). Safety Evaluation of New Hemostatic Agents, Smectite Granules, and Kaolin-Coated Gauze in a Vascular Injury Wound Model in Swine. The Journal of Trauma: Injury, Infection, and Critical Care, 68(2), 269–278. http://doi.org/10.1097/TA.0b013e3181c97ef1 Kheirabadi, B. S., Pearson, R., Rudnicka, K., Somwaru, L., MacPhee, M., Drohan, W., & Tuthill, D. (2001). Development of an Animal Model for Assessment of the Hemostatic Efficacy of Fibrin Sealant in Vascular Surgery. Journal of Surgical Research, 100(1), 84–92. http://doi.org/10.1006/JSRE.2001.6226

350 Kheirabadi, B. S., Pearson, R., Tuthill, D., Rudnicka, K., Holcomb, J. B., Drohan, W., & MacPhee, M. J. (2002). Comparative Study of the Hemostatic Efficacy of a New Human Fibrin Sealant: Is an Antifibrinolytic Agent Necessary? Journal of Trauma and Acute Care Surgery, 52(6). Retrieved from https://journals.lww.com/jtrauma/Fulltext/2002/06000/Comparative_Study_of_the_H emostatic_Efficacy_of_a.14.aspx Kheirabadi, B. S., Pusateri, A. E., Sondeen, J. L., Delgado, A. V., Modrow, H. E., Hess, J. R., & Holcomb, J. B. (2004). Development of Hemostatic Dressings for Use in Military Operations. Combat Casualty Care in Ground Based Tactical Situations: Trauma Technology and Emergency Medical Procedures, (August), P34-1-P34-12. Kheirabadi, B. S., Scherer, M. R., Estep, J. S., Dubick, M. A., & Holcomb, J. B. (2009). Determination of Efficacy of New Hemostatic Dressings in a Model of Extremity Arterial Hemorrhage in Swine. The Journal of Trauma: Injury, Infection, and Critical Care, 67(3), 450–460. http://doi.org/10.1097/TA.0b013e3181ac0c99 Kheirabadi, B. S., Sieber, J., Bukhari, T., Rudnicka, K., Murcin, L. A., & Tuthill, D. (2008). High-Pressure Fibrin Sealant Foam: An Effective Hemostatic Agent for Treating Severe Parenchymal Hemorrhage. Journal of Surgical Research, 144(1), 145–150. http://doi.org/10.1016/J.JSS.2007.02.012 King, D. R., Cohn, S. M., & Proctor, K. G. (2004). Modified Rapid Deployment Hemostat Bandage Terminates Bleeding in Coagulopathic Patients with Severe Visceral Injuries. The Journal of Trauma: Injury, Infection, and Critical Care, 57(4), 756–759. http://doi.org/10.1097/01.TA.0000147501.64610.AFS Kjaergard, H. K., & Trumbull, H. R. (2000). Bleeding from the sternal marrow can be stopped using vivostat patient-derived fibrin sealant. The Annals of Thoracic Surgery, 69(4), 1173–1175. http://doi.org/10.1016/S0003-4975(99)01560-X Kjaergard, H. K., & Weis-Fogh, U. S. (1994). Important factors influencing the strength of autologous fibrin glue; the fibrin concentration and reaction time--comparison of strength with commercial fibrin glue. European Surgical Research, 26(5), 273–6. http://doi.org/10.1159/000129346 Klein, H. G. (1995). Allogeneic transfusion risks in the surgical patient. American Journal of Surgery, 170(6A Suppl), 21S–26S. http://doi.org/10.1016/S0002- 9610(99)80054-3 Knill, C. ., Kennedy, J. ., Mistry, J., Miraftab, M., Smart, G., Groocock, M. ., & Williams, H. . (2004). Alginate fibres modified with unhydrolysed and hydrolysed chitosans for wound dressings. Carbohydrate Polymers, 55(1), 65–76. http://doi.org/10.1016/J.CARBPOL.2003.08.004 Kodama, K., Doi, O., Higashiyama, M., & Yokouchi, H. (1997). Pneumostatic effect of gelatin-resorcinol formaldehyde-glutaraldehyde glue on thermal injury of the lung: an experimental study on rats. European Journal of Cardio-thoracic Surgery (Vol. 11). Retrieved from https://academic.oup.com/ejcts/article-abstract/11/2/333/374926 Kolev, K., & Longstaff, C. (2016). Bleeding related to disturbed fibrinolysis. British Journal of Haematology, 175(1), 12–23. http://doi.org/10.1111/bjh.14255

351 Kolodziej, A. F., Nair, S. A., Graham, P., McMurry, T. J., Ladner, R. C., Wescott, C., … Caravan, P. (2012). Fibrin Specific Peptides Derived by Phage Display: Characterization of Peptides and Conjugates for Imaging. Bioconjugate Chemistry, 23(3), 548–556. http://doi.org/10.1021/bc200613e Kragh, J. F., Aden, J. K., Steinbaugh, J., Bullard, M., & Dubick, M. A. (2015). Gauze vs XSTAT in wound packing for hemorrhage control. The American Journal of Emergency Medicine, 33(7), 974–976. http://doi.org/10.1016/J.AJEM.2015.03.048 Kragh, J. F., Swan, K. G., Smith, D. C., Mabry, R. L., & Blackbourne, L. H. (2012). Historical review of emergency tourniquet use to stop bleeding. The American Journal of Surgery, 203(2), 242–252. http://doi.org/10.1016/J.AMJSURG.2011.01.028 Kragh, J. F., Walters, T. J., Baer, D. G., Fox, C. J., Wade, C. E., Salinas, J., & Holcomb, J. B. (2009). Survival With Emergency Tourniquet Use to Stop Bleeding in Major Limb Trauma. Annals of Surgery, 249(1), 1–7. http://doi.org/10.1097/SLA.0b013e31818842ba Krishnan, L. K., Mohanty, M., Umashankar, P. R., & Vijayan Lal, A. (2004). Comparative evaluation of absorbable hemostats: advantages of fibrin-based sheets. Biomaterials, 25(24), 5557–5563. http://doi.org/10.1016/J.BIOMATERIALS.2004.01.013 Krug, E. G., Sharma, G. K., & Lozano, R. (2000). The global burden of injuries. American Journal of , 90(4), 523–6. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10754963 Kudela, D., Smith, S. A., May-Masnou, A., Braun, G. B., Pallaoro, A., Nguyen, C. K., … Stucky, G. D. (2015). Clotting Activity of Polyphosphate-Functionalized Silica Nanoparticles. Angewandte Chemie International Edition, 54(13), 4018–4022. http://doi.org/10.1002/anie.201409639 Kumar Dutta, P., Dutta, J., & Tripathi, V. S. (2004). Chitin and chitosan: Chemistry, properties and applications. Journal of Scientific & Industrial Research (Vol. 63). Retrieved from http://nopr.niscair.res.in/bitstream/123456789/5397/1/JSIR 63%281%29 20-31.pdf Kunihara, T., Shiiya, N., Matsuzaki, K., Murashita, T., & Matsui, Y. (2008). Recommendation for Appropriate Use of GRF Glue in the Operation for Acute Aortic Dissection. Ann Thorac Cardiovasc Surg (Vol. 14). Retrieved from https://pdfs.semanticscholar.org/8312/a0ec657dc25906092a289592d9acee8cbe42.pdf Kurachi, K., & Davie, E. W. (1982). Isolation and characterization of a cDNA coding for human factor IX. Proceedings of the National Academy of Sciences of the United States of America, 79(21), 6461–4. http://doi.org/10.1073/PNAS.79.21.6461 Kutcher, M. E., Kornblith, L. Z., Narayan, R., Curd, V., Daley, A. T., Redick, B. J., … Cohen, M. J. (2013). A Paradigm Shift in Trauma Resuscitation. JAMA Surgery, 148(9), 834. http://doi.org/10.1001/jamasurg.2013.2911 L’Esperance, J. O., Sung, J. C., Marguet, C. G., Maloney, M. E., Springhart, W. P., Preminger, G. M., & Albala, D. M. (2005). Controlled Survival Study of the Effects

352 of Tisseel or a Combination of FloSeal and Tisseel on Major Vascular Injury and Major Collecting-System Injury during Partial Nephrectomy in a Porcine Model. Journal of Endourology, 19(9), 1114–1121. http://doi.org/10.1089/end.2005.19.1114 Laidmäe, I., McCormick, M. E., Herod, J. L., Pastore, J. J., Salum, T., Sawyer, E. S., … Uibo, R. (2006). Stability, sterility, coagulation, and immunologic studies of salmon coagulation proteins with potential use for mammalian wound healing and cell engineering. Biomaterials, 27(34), 5771–5779. http://doi.org/10.1016/J.BIOMATERIALS.2006.07.035 Lakstein, D., Blumenfeld, A., Sokolov, T., Lin, G., Bssorai, R., Lynn, M., & Ben- Abraham, R. (2003). Tourniquets for hemorrhage control on the battlefield: a 4-year accumulated experience. The Journal of Trauma, 54(5 Suppl), S221–S225. http://doi.org/10.1097/01.TA.0000047227.33395.49 Lam, W. A., Chaudhuri, O., Crow, A., Webster, K. D., Li, T.-D., Kita, A., … Fletcher, D. A. (2011). Mechanics and contraction dynamics of single platelets and implications for clot stiffening. Nature Materials, 10(1), 61–66. http://doi.org/10.1038/nmat2903 Lambert, M. P., Sullivan, S. K., Fuentes, R., French, D. L., & Poncz, M. (2013). Challenges and promises for the development of donor-independent platelet transfusions. Blood, 121(17), 3319–24. http://doi.org/10.1182/blood-2012-09-455428 Lambert, M. P., Sullivan, S. K., Fuentes, R., French, D. L., Poncz, M., & Dc, W. (2013). Challenges and promises for the development of donor-independent platelet transfusions Review Article Challenges and promises for the development of donor- independent platelet transfusions, 121(17), 3319–3324. http://doi.org/10.1182/blood- 2012-09-455428 Larson, M. J., Bowersox, J. C., Lim, R. C., & Hess, J. R. (1995). Efficacy of a Fibrin Hemostatic Bandage in Controlling Hemorrhage From Experimental Arterial Injuries. Archives of Surgery, 130(4), 420. http://doi.org/10.1001/archsurg.1995.01430040082018 Lashof-Sullivan, M. M., Shoffstall, E., Atkins, K. T., Keane, N., Bir, C., VandeVord, P., & Lavik, E. B. (2014). Intravenously administered nanoparticles increase survival following blast trauma. Proceedings of the National Academy of Sciences, 111(28), 10293–10298. http://doi.org/10.1073/pnas.1406979111 Lashof-Sullivan, M., Shoffstall, A., & Lavik, E. (2013). Intravenous hemostats: challenges in translation to patients. Nanoscale, 5(22), 10719–28. http://doi.org/10.1039/c3nr03595f Laupacis, A., & Fergusson, D. (1997). Drugs to minimize perioperative blood loss in cardiac surgery: meta-analyses using perioperative blood transfusion as the outcome. The International Study of Peri-operative Transfusion (ISPOT) Investigators. Anesthesia and Analgesia, 85(6), 1258–67. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9390590 Laurence, S., Bareille, R., Baquey, C., & Fricain, J. C. (2005). Development of a resorbable macroporous cellulosic material used as hemostatic in an osseous

353 environment. Journal of Biomedical Materials Research Part A, 73A(4), 422–429. http://doi.org/10.1002/jbm.a.30280 Lee, D. H., & Blajchman, M. A. (2001). Novel treatment modalities: New platelet preparations and subsititutes. British Journal of Haematology, 114(3), 496–505. http://doi.org/10.1046/j.1365-2141.2001.03004.x Lee, D. H., Blajchman, M. A., Ho, T. W. C., & Yen, R. C. K. (1995). In vitro characteristics of Thrombospheres, a novel hemostatic agent. Blood, 86(suppl 1), 902a. Lee, K. Y., & Mooney, D. J. (2012). Alginate: Properties and biomedical applications. Progress in Polymer Science, 37(1), 106–126. http://doi.org/10.1016/J.PROGPOLYMSCI.2011.06.003 Leixnering, M., Reichetseder, J., Schultz, A., Figl, M., Wassermann, E., Thurnher, M., & Redl, H. (2008). Gelatin Thrombin Granules for Hemostasis in a Severe Traumatic Liver and Spleen Rupture Model in Swine. The Journal of Trauma Injury Infection and Critical Care, 64(2), 456–461. http://doi.org/10.1097/TA.0b013e3180340de1 Lentz, B. (2003). Exposure of platelet membrane phosphatidylserine regulates blood coagulation. Progress in Lipid Research. Retrieved from https://www.sciencedirect.com/science/article/pii/S0163782703000250 Levi, M., Cromheecke, M. E., de Jonge, E., Prins, M. H., de Mol, B. J., Briët, E., & Büller, H. R. (1999). Pharmacological strategies to decrease excessive blood loss in cardiac surgery: a meta-analysis of clinically relevant endpoints. The Lancet, 354(9194), 1940–1947. http://doi.org/10.1016/S0140-6736(99)01264-7 Levi, M., Friederich, P. W., Middleton, S., de Groot, P. G., Wu, Y. P., Harris, R., … ten Cate, J. W. (1999). Fibrinogen-coated albumin microcapsules reduce bleeding in severely thrombocytopenic rabbits. Nature Medicine, 5(1), 107–11. http://doi.org/10.1038/4795 Levi, M., Levy, J. H., Andersen, H. F., & Truloff, D. (2010). Safety of Recombinant Activated Factor VII in in Randomized Clinical Trials, 1791–1800. Levy, J. H. (2004). Hemostatic agents. Transfusion, 44(S2), 58S–62S. http://doi.org/10.1111/j.0041-1132.2004.04173.x Levy, J. H., & Goodnough, L. T. (2015). How I use fibrinogen replacement therapy in acquired bleeding. Blood, 125(9), 1387–93. http://doi.org/10.1182/blood-2014-08- 552000 Lewis, K. M., Atlee, H. D., Mannone, A. J., Dwyer, J., Lin, L., Goppelt, A., & Redl, H. (2013). Comparison of Two Gelatin and Thrombin Combination Hemostats in a Porcine Liver Abrasion Model. Journal of Investigative Surgery, 26(3), 141–148. http://doi.org/10.3109/08941939.2012.724519 Lewis, K. M., Spazierer, D., Urban, M. D., Lin, L., Redl, H., & Goppelt, A. (2013). Comparison of regenerated and non-regenerated oxidized cellulose hemostatic agents. European Surgery, 45(4), 213–220. http://doi.org/10.1007/s10353-013-0222-z

354 Li, D., Li, P., Zang, J., & Liu, J. (2012). Enhanced hemostatic performance of tranexamic acid-loaded chitosan/alginate composite microparticles. Journal of and Biotechnology, 2012. http://doi.org/10.1155/2012/981321 Li, J., Ai, Y., Wang, L., Bu, P., Sharkey, C. C., Wu, Q., … King, M. R. (2016). to circulating tumor cells via platelet membrane-functionalized particles. Biomaterials, 76, 52–65. http://doi.org/10.1016/J.BIOMATERIALS.2015.10.046 Li, J., Cao, W., Lv, X., Jiang, L., Li, Y., Li, W., … Li, X. (2013). Zeolite-based hemostat QuikClot releases calcium into blood and promotes blood coagulation in vitro. Acta Pharmacologica Sinica, 34(3), 367–372. http://doi.org/10.1038/aps.2012.159 Li, Y., & Boraschi, D. (2016). Endotoxin contamination: a key element in the interpretation of nanosafety studies. Nanomedicine, 11(3), 269–287. http://doi.org/10.2217/nnm.15.196 Li, Y., Li, H., Xiao, L., Zhou, L., Shentu, J., Zhang, X., … Fan, J. (2012). Hemostatic Efficiency and Wound Healing Properties of Natural Zeolite Granules in a Lethal Rabbit Model of Complex Groin Injury. Materials, 5(12), 2586–2596. http://doi.org/10.3390/ma5122586 Lin, W.-C., Tseng, C.-H., & Yang, M.-C. (2005). In-Vitro Hemocompatibility Evaluation of a Thermoplastic Polyurethane Membrane with Surface-Immobilized Water- Soluble Chitosan and Heparin. Macromolecular Bioscience, 5(10), 1013–1021. http://doi.org/10.1002/mabi.200500077 Liumbruno, G., Bennardello, F., Lattanzio, A., Piccoli, P., Rossetti, G., & Italian Society of Transfusion Medicine and Immunohaematology (SIMTI) Work Group, as I. S. of T. M. and I. (SIMTI) W. (2009). Recommendations for the transfusion of plasma and platelets. Blood Transfusion = Trasfusione Del Sangue, 7(2), 132–50. http://doi.org/10.2450/2009.0005-09 Lomas-Niera, J. L., Perl, M., Chung, C.-S., & Ayala, A. (2005). SHOCK AND HEMORRHAGE: AN OVERVIEW OF ANIMAL MODELS. Shock, 24(Supplement 1), 33–39. http://doi.org/10.1097/01.shk.0000191411.48719.ab Lozano, M. (2008). Platelets come back in from the cold. Blood, 111(6), 2951–2951. http://doi.org/10.1182/blood-2008-01-131920 Lubarsky, M., Ray, C., & Funaki, B. (2010). Embolization agents-which one should be used when? Part 2: small-vessel embolization. Seminars in , 27(1), 99–104. http://doi.org/10.1055/s-0030-1247891 Lundblad, R. L., Bradshaw, R. A., Gabriel, D., Ortel, T., Lawson, J., & Mann, K. G. (2004). A review of the therapeutic uses of thrombin. Thrombosis and Haemostasis, 91(05), 851–860. http://doi.org/10.1160/TH03-12-0792 Lusher, J. M., Arkin, S., Abildgaard, C. F., & Schwartz, R. S. (1993). Recombinant Factor VIII for the Treatment of Previously Untreated Patients with Hemophilia A -- Safety, Efficacy, and Development of Inhibitors. New England Journal of Medicine, 328(7), 453–459. http://doi.org/10.1056/NEJM199302183280701

355 MACFARLANE, R. G. (1964). An Enzyme Cascade in the Blood Clotting Mechanism and its Function as a Biochemical Amplifier. Nature, 202(4931), 498–499. http://doi.org/10.1038/202498a0 Mackenzie, E. J., Rivara, F. P., Jurkovich, G. J., Nathens, A. B., Frey, K. P., Egleston, B. L., … Scharfstein, D. O. (2007). The National Study on Costs and Outcomes of Trauma. J Trauma, 63, 54–67. http://doi.org/10.1097/TA.0b013e31815acb09 Mair, H., Kaczmarek, I., Oberhoffer, M., Groetzner, J., Daebritz, S., & Reichart, B. (2005). Surgicel Nu-Knit hemostat for bleeding control of fragile sternum. The Journal of Thoracic and Cardiovascular Surgery, 130(2), 605–606. http://doi.org/10.1016/j.jtcvs.2005.04.013 Malette, W. G., Quigley, H. J., Gaines, R. D., Johnson, N. D., & Rainer, W. G. (1983). Chitosan: A New Hemostatic. The Annals of Thoracic Surgery, 36(1), 55–58. http://doi.org/10.1016/S0003-4975(10)60649-2 Mandell, S. P., & Gibran, N. S. (2014). Fibrin sealants: surgical hemostat, sealant and adhesive. Expert Opinion on Biological Therapy, 14(6), 821–830. http://doi.org/10.1517/14712598.2014.897323 MARGULIS, V., MATSUMOTO, E. D., SVATEK, R., KABBANI, W., CADEDDU, J. A., & LOTAN, Y. (2005). APPLICATION OF NOVEL HEMOSTATIC AGENT DURING LAPAROSCOPIC PARTIAL NEPHRECTOMY. The Journal of Urology, 174(2), 761–764. http://doi.org/10.1097/01.JU.0000164727.53999.FE Mathias, J.-D., Tessier-Doyen, N., & Michaud, P. (2011). Development of a Chitosan- Based Biofoam: Application to the Processing of a Porous Ceramic Material. International Journal of Molecular Sciences, 12(2), 1175–1186. http://doi.org/10.3390/ijms12021175 Matthew, I. R., Browne, R. M., Frame, J. W., & Millar, B. G. (1993). Tissue response to a haemostatic alginate wound dressing in tooth extraction sockets. The British Journal of Oral & Maxillofacial Surgery, 31(3), 165–9. http://doi.org/10.1016/0266- 4356(93)90117-F Matthew, I. R., Browne, R. M., Frame, J. W., & Millar, B. G. (1994). Alginate fiber dressing for oral mucosal wounds. Oral Surgery, , Oral Pathology, 77(5), 456–460. http://doi.org/10.1016/0030-4220(94)90223-2 Mayo, A., Misgav, M., Kluger, Y., Geenberg, R., Pauzner, D., Klausner, J., & Ben-Tal, O. (2004). Recombinant activated factor VII (NovoSevenTM): Addition to replacement therapy in acute, uncontrolled and life-threatening bleeding. Vox Sanguinis, 87(1), 34–40. http://doi.org/10.1111/j.1423-0410.2004.00533.x McCormack, P. L. (2012). Tranexamic Acid. Drugs, 72(5), 585–617. http://doi.org/10.2165/11209070-000000000-00000 McManus, J., Hurtado, T., Pusateri, A., & Knoop, K. J. (2007). A case series describing thermal injury resulting from zeolite use for hemorrhage control in combat operations. Prehospital Emergency Care : Official Journal of the National Association of EMS Physicians and the National Association of State EMS Directors, 11(1), 67– 71. http://doi.org/10.1080/10903120601021176

356 Mehdizadeh, M., Weng, H., Gyawali, D., Tang, L., & Yang, J. (2012). Injectable citrate- based mussel-inspired tissue bioadhesives with high wet strength for sutureless wound closure. Biomaterials, 33(32), 7972–7983. http://doi.org/10.1016/J.BIOMATERIALS.2012.07.055 Mehdizadeh, M., & Yang, J. (2013). Design Strategies and Applications of Tissue Bioadhesives. Macromolecular Bioscience, 13(3), 271–288. http://doi.org/10.1002/mabi.201200332 Michaud, S. E., Wang, L. Z., Korde, N., Bucki, R., Randhawa, P. K., Pastore, J. J., … Janmey, P. A. (2002). Purification of salmon thrombin and its potential as an alternative to mammalian thrombins in fibrin sealants. Thrombosis Research, 107(5), 245–254. http://doi.org/10.1016/S0049-3848(02)00333-X Midathada, M. V, Mehta, P., Waner, M., & Fink, L. M. (2004). Recombinant Factor VIIa in the Treatment of Bleeding. American Journal of Clinical Pathology, 121(1), 124– 137. http://doi.org/10.1309/D0G0-C96V-05CJ-5B4J Milford, E. M., & Reade, M. C. (2016). Comprehensive review of platelet storage methods for use in the treatment of active hemorrhage. Transfusion, 56 Suppl 2, S140-8. http://doi.org/10.1111/trf.13504 MILLER, J. M., JACKSON, D. A., & COLLIER, C. S. (1961). An investigation of the chemical reactions of oxidized regenerated cellulose. Experimental Medicine and Surgery, 19, 196–201. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/14474048 Millner, R. W. J., Lockhart, A. S., Bird, H., & Alexiou, C. (2009). A New Hemostatic Agent: Initial Life-Saving Experience With Celox (Chitosan) in Cardiothoracic Surgery. The Annals of Thoracic Surgery, 87(2), e13–e14. http://doi.org/10.1016/J.ATHORACSUR.2008.09.046 Mitra, B., Cameron, P. A., Parr, M. J., & Phillips, L. (2012). Recombinant factor VIIa in trauma patients with the “triad of death.” Injury, 43(9), 1409–1414. http://doi.org/10.1016/j.injury.2011.01.033 Mo, X., An, Y., Yun, C.-S., & Yu, S. M. (2006). Nanoparticle-Assisted Visualization of Binding Interactions between Collagen Mimetic Peptide and Collagen Fibers. Angewandte Chemie, 118(14), 2325–2328. http://doi.org/10.1002/ange.200504529 Modery-Pawlowski, C. L., Kuo, H.-H., Baldwin, W. M., & Gupta, A. Sen. (2013). A platelet-inspired paradigm for nanomedicine targeted to multiple diseases. Nanomedicine, 8(10), 1709–1727. http://doi.org/10.2217/nnm.13.113 Modery-Pawlowski, C. L., Tian, L. L., Pan, V., McCrae, K. R., Mitragotri, S., & Sen Gupta, A. (2013). Approaches to synthetic platelet analogs. Biomaterials, 34(2), 526– 541. http://doi.org/10.1016/j.biomaterials.2012.09.074 Modery-Pawlowski, C. L., Tian, L. L., Ravikumar, M., Wong, T. L., & Gupta, A. Sen. (2013). In vitro and in vivo hemostatic capabilities of a functionally integrated platelet-mimetic liposomal nanoconstruct. Biomaterials, 34(12), 3031–3041. http://doi.org/10.1016/j.biomaterials.2012.12.045

357 Mogoşanu, G. D., & Grumezescu, A. M. (2014). Natural and synthetic polymers for wounds and burns dressing. International Journal of Pharmaceutics, 463(2), 127–136. http://doi.org/10.1016/J.IJPHARM.2013.12.015 Mohanty, D. (2009). Current concepts in platelet transfusion. Asian Journal of Transfusion Science, 3(1), 18–21. http://doi.org/10.4103/0973-6247.45257 Montanaro, L., Arciola, C. ., Cenni, E., Ciapetti, G., Savioli, F., Filippini, F., & Barsanti, L. . (2000). Cytotoxicity, blood compatibility and antimicrobial activity of two cyanoacrylate glues for surgical use. Biomaterials, 22(1), 59–66. http://doi.org/10.1016/S0142-9612(00)00163-0 Moody, A. (2006). Use of a hydrogel dressing for management of a painful leg ulcer. British Journal of Community Nursing, 11(Sup3), S12–S17. http://doi.org/10.12968/bjcn.2006.11.Sup3.21212 Moore, F. A. (2009). Tourniquets. Annals of Surgery, 249(1), 8–9. http://doi.org/10.1097/SLA.0b013e3181932329 Moore, H. B., Moore, E. E., Huebner, B. R., Stettler, G. R., Nunns, G. R., Einersen, P. M., … Sauaia, A. (2017). Tranexamic acid is associated with increased mortality in patients with physiological fibrinolysis. Journal of Surgical Research, 220, 438–443. http://doi.org/10.1016/j.jss.2017.04.028 Morawitz, P. (1905). Die Chemie der Blutgerinnung. Ergebnisse Der Physiologie, 4(1), 307–422. http://doi.org/10.1007/BF02321003 MOREY, A. F., ANEMA, J. G., HARRIS, R., GRESHAM, V., DANIELS, R., KNIGHT, R. W., … CORNUM, R. L. (2001). TREATMENT OF GRADE 4 RENAL STAB WOUNDS WITH ABSORBABLE FIBRIN ADHESIVE BANDAGE IN A PORCINE MODEL. The Journal of Urology, 165(3), 955–958. http://doi.org/10.1016/S0022-5347(05)66583-4 Morgan, C. E., Prakash, V. S., Vercammen, J. M., Pritts, T., & Kibbe, M. R. (2015). Development and Validation of 4 Different Rat Models of Uncontrolled Hemorrhage. JAMA Surg, 150(4), 316–324. http://doi.org/10.1001/jamasurg.2014.1685 Morgan, D. (1999). Wound management products in the drug tariff. Pharmaceutical Journal. Retrieved from https://scholar.google.com/scholar_lookup?hl=en&volume=263&publication_year=1 999&pages=820&journal=Pharm.+J.&author=D.+A.+Morgan&title=Pharm.+J. Moroff, G., & Holme, S. (1991). Concepts about current conditions for the preparation and storage of platelets. Transfusion Medicine Reviews, 5(1), 48–59. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/1802276 Morrison, J. J., Dubose, J. J., Rasmussen, T. E., & Midwinter, M. J. (2012). Military Application of Tranexamic Acid in Trauma Emergency Resuscitation (MATTERs) Study. Archives of Surgery, 147(2), 113. http://doi.org/10.1001/archsurg.2011.287 Morrissey, J., Choi, S., & Smith, S. (2012). Polyphosphate: an ancient molecule that links platelets, coagulation and inflammation. Blood. Retrieved from http://www.bloodjournal.org/content/early/2012/04/18/blood-2012-03-306605.short

358 Morrissey, J. H. (2012). Polyphosphate: a link between platelets, coagulation and inflammation. International Journal of Hematology, 95(4), 346–352. http://doi.org/10.1007/s12185-012-1054-5 MOSESSON, M. W. (2005). Fibrinogen and fibrin structure and functions. Journal of Thrombosis and Haemostasis, 3(8), 1894–1904. http://doi.org/10.1111/j.1538- 7836.2005.01365.x Mosnier, L., Lisman, T., Berg, M. van den, Nieuwenhuis, K., Meijers, J., & Bouma, B. (2001). The Defective Down Regulation of Fibrinolysis in Can Be Restored by Increasing the TAFI Plasma Concentration. Thrombosis and Haemostasis, 86(10), 1035–1039. http://doi.org/10.1055/s-0037-1616530 Mueller, G. R., Pineda, T. J., Xie, H. X., Teach, J. S., Barofsky, A. D., Schmid, J. R., & Gregory, K. W. (2012). A novel sponge-based wound stasis dressing to treat lethal noncompressible hemorrhage. Journal of Trauma and Acute Care Surgery, 73, S134– S139. http://doi.org/10.1097/TA.0b013e3182617c3c Muench, T. R., Kong, W., & Harmon, A. M. (2010). The performance of a hemostatic agent based on oxidized regenerated cellulose – polyglactin 910 composite in a liver defect model in immunocompetent and athymic rats. Biomaterials, 31(13), 3649– 3656. http://doi.org/10.1016/J.BIOMATERIALS.2010.01.070 Mullin, J. L., Gorkun, O. V, Binnie, C. G., & Lord, S. T. (2000). Recombinant fibrinogen studies reveal that thrombin specificity dictates order of fibrinopeptide release. The Journal of Biological Chemistry, 275(33), 25239–46. http://doi.org/10.1074/jbc.M004142200 Muszbek, L., Bereczky, Z., Bagoly, Z., Komáromi, I., & Katona, É. (2011). Factor XIII: A Coagulation Factor With Multiple Plasmatic and Cellular Functions. Physiological Reviews, 91(3), 931–972. http://doi.org/10.1152/physrev.00016.2010 Naimer, S. A. (2014). A Review of Methods to Control Bleeding from Life-Threatening Traumatic Wounds. Health, 6, 479–490. http://doi.org/10.4236/health.2014.66067 Naimer, S. A., Anat, N., & Katif, G. (2004). Evaluation of techniques for treating the bleeding wound. Injury, 35(10), 974–979. http://doi.org/10.1016/S0020- 1383(03)00316-4 Naimer, S. A., & Chemla, F. (2000). Elastic adhesive dressing treatment of bleeding wounds in trauma victims. The American Journal of Emergency Medicine, 18(7), 816–819. http://doi.org/10.1053/AJEM.2000.18126 Nair, A. B., & Jacob, S. (2016). A simple practice guide for dose conversion between animals and human. Journal of Basic and Clinical , 7(2), 27–31. http://doi.org/10.4103/0976-0105.177703 Nair, P. M., Pidcoke, H. F., Cap, A. P., & Ramasubramanian, A. K. (2014). Effect of cold storage on shear-induced platelet aggregation and clot strength. Journal of Trauma and Acute Care Surgery, 77, S88–S93. http://doi.org/10.1097/TA.0000000000000327 Nakamura, S., Takayama, N., Hirata, S., Seo, H., Endo, H., Ochi, K., … Eto, K. (2014). Expandable Megakaryocyte Cell Lines Enable Clinically Applicable Generation of

359 Platelets from Human Induced Pluripotent Stem Cells. Cell Stem Cell, 14(4), 535– 548. http://doi.org/10.1016/J.STEM.2014.01.011 Nasiri, S. (2013). Infusible platelet membrane as a platelet substitute for transfusion: an overview. Blood Transfusion, 11(3), 337–42. http://doi.org/10.2450/2013.0209-12 Nasiri, S., Heidari, M., & Rivandi, S. (2012). INFUSIBLE PLATELET MEMBRANES IMPROVE HEMOSTASIS IN THROMBOCYTOPENIC RABBITS: STUDIES WITH TWO DIFFERENT INJECTION DOSES. IJPSR, 3(12). Retrieved from www.ijpsr.com Neal, M. D., Hoffman, M. K., Cuschieri, J., Minei, J. P., Maier, R. V, Harbrecht, B. G., … Investigators, T. I. and the H. R. to I. (2012). Crystalloid to packed red blood cell transfusion ratio in the massively transfused patient: when a little goes a long way. The Journal of Trauma and Acute Care Surgery, 72(4), 892–8. http://doi.org/10.1097/TA.0b013e31823d84a7 Nelson, P. A., Powers, J. N., Estridge, T. D., Elder, E. A., Alea, A. D., Sidhu, P. K., … DeLustro, F. A. (2001). Serological analysis of patients treated with a new surgical hemostat containing bovine proteins and autologous plasma. Journal of Biomedical Materials Research, 58(6), 710–719. http://doi.org/10.1002/jbm.10025 Nie, W., Yuan, X., Zhao, J., Zhou, Y., & Bao, H. (2013). Rapidly in situ forming chitosan/ε-polylysine hydrogels for adhesive sealants and hemostatic materials. Carbohydrate Polymers, 96(1), 342–348. http://doi.org/10.1016/J.CARBPOL.2013.04.008 Nikolajsen, C. L., Dyrlund, T. F., Poulsen, E. T., Enghild, J. J., & Scavenius, C. (2014). Coagulation factor XIIIa substrates in human plasma: identification and incorporation into the clot. The Journal of Biological Chemistry, 289(10), 6526–34. http://doi.org/10.1074/jbc.M113.517904 Nishiya, T., Kainoh, M., Murata, M., Handa, M., & Ikeda, Y. (2001). PLATELET INTERACTIONS WITH LIPOSOMES CARRYING RECOMBINANT PLATELET MEMBRANE GLYCOPROTEINS OR FIBRINOGEN: APPROACH TO PLATELET SUBSTITUTES. Artificial Cells, Blood Substitutes, and Biotechnology, 29(6), 453–464. http://doi.org/10.1081/BIO-100108550 Nishiya, T., Kainoh, M., Murata, M., Handa, M., & Ikeda, Y. (2002). Reconstitution of adhesive properties of human platelets in liposomes carrying both recombinant glycoproteins Ia/IIa and Ib?? under flow conditions: Specific synergy of receptor- ligand interactions. Blood, 100(1), 136–142. http://doi.org/10.1182/blood.V100.1.136 Nogami, K., Shima, M., Giddings, J. C., Takeyama, M., Tanaka, I., & Yoshioka, A. (2007). Relationship between the binding sites for von Willebrand factor, phospholipid, and human factor VIII C2 inhibitor alloantibodies within the factor VIII C2 domain. International Journal of Hematology, 85(4), 317–22. http://doi.org/10.1532/IJH97.06192 Noorman, F., van Dongen, T. T. C. F., Plat, M.-C. J., Badloe, J. F., Hess, J. R., & Hoencamp, R. (2016). Transfusion: -80°C Frozen Blood Products Are Safe and

360 Effective in Military Casualty Care. PLOS ONE, 11(12), e0168401. http://doi.org/10.1371/journal.pone.0168401 O’Hara, P. J., Grant, F. J., Haldeman, B. A., Gray, C. L., Insley, M. Y., Hagen, F. S., & Murray, M. J. (1987). Nucleotide sequence of the gene coding for human factor VII, a vitamin K-dependent protein participating in blood coagulation. Proceedings of the National Academy of Sciences of the United States of America, 84(15), 5158–62. http://doi.org/10.1073/PNAS.84.15.5158 Ofosu, F. A., Freedman, J., & Semple, J. W. (2008). Plasma-derived biological medicines used to promote haemostasis. Thrombosis and Haemostasis, 99(11), 851–862. http://doi.org/10.1160/TH07-10-0592 Okamoto, S., Hijikata-Okunomiya, A., Wanaka, K., Okada, Y., & Okamoto, U. (1997). Enzyme-Controlling Medicines: Introduction. Seminars in Thrombosis and Hemostasis, 23(06), 493–501. http://doi.org/10.1055/s-2007-996127 OKAMOTO, S., & OKAMOTO, U. (1962). AMINO-METHYL-CYCLOHEXANE- CARBOXYLIC ACID: AMCHA. The Keio Journal of Medicine, 11(3), 105–115. http://doi.org/10.2302/kjm.11.105 Okamoto, Y., Yano, R., Miyatake, K., Tomohiro, I., Shigemasa, Y., & Minami, S. (2003). Effects of chitin and chitosan on blood coagulation. Carbohydrate Polymers, 53(3), 337–342. http://doi.org/10.1016/S0144-8617(03)00076-6 Okamura, Y., Fujie, T., Maruyama, H., Handa, M., Ikeda, Y., & Takeoka, S. (2007). Prolonged hemostatic ability of polyethylene glycol-modified polymerized albumin particles carrying fibrinogen γ-chain dodecapeptide. Transfusion, 47(7), 1254–1262. http://doi.org/10.1111/j.1537-2995.2007.01265.x Okamura, Y., Fujie, T., Nogawa, M., Maruyama, H., Handa, M., Ikeda, Y., & Takeoka, S. (2008). Haemostatic effects of polymerized albumin particles carrying fibrinogen γ-chain dodecapeptide as platelet substitutes in severely thrombocytopenic rabbits. Transfusion Medicine, 18(3), 158–166. http://doi.org/10.1111/j.1365- 3148.2008.00860.x Okamura, Y., Handa, M., Suzuki, H., Ikeda, Y., & Takeoka, S. (2006). New strategy of platelet substitutes for enhancing platelet aggregation at high shear rates: cooperative effects of a mixed system of fibrinogen γ-chain dodecapeptide- or glycoprotein Ibα- conjugated latex beads under flow conditions. Journal of Artificial Organs, 9(4), 251– 258. http://doi.org/10.1007/s10047-006-0345-0 Okamura, Y., Maekawa, I., Teramura, Y., Maruyama, H., Handa, M., Ikeda, Y., & Takeoka, S. (2005). Hemostatic Effects of Phospholipid Vesicles Carrying Fibrinogen γ Chain Dodecapeptide in Vitro and in Vivo. Bioconjugate Chemistry, 16(6), 1589– 1596. http://doi.org/10.1021/bc050178g OKAMURA, Y., TAKEOKA, S., ETO, K., MAEKAWA, I., FUJIE, T., MARUYAMA, H., … HANDA, M. (2009). Development of fibrinogen γ-chain peptide-coated, adenosine diphosphate-encapsulated liposomes as a synthetic platelet substitute. Journal of Thrombosis and Haemostasis, 7(3), 470–477. http://doi.org/10.1111/j.1538-7836.2008.03269.x

361 Okamura, Y., Takeoka, S., Teramura, Y., Maruyama, H., Tsuchida, E., Handa, M., & Ikeda, Y. (2005). Hemostatic effects of fibrinogen gamma-chain dodecapeptide- conjugated polymerized albumin particles in vitro and in vivo. Transfusion, 45(7), 1221–1228. http://doi.org/10.1111/j.1537-2995.2005.00173.x Olsen, D., Yang, C., Bodo, M., Chang, R., Leigh, S., Baez, J., … Polarek, J. (2003). Recombinant collagen and gelatin for drug delivery. Advanced Drug Delivery Reviews, 55(12), 1547–1567. http://doi.org/10.1016/J.ADDR.2003.08.008 Onwukwe, C., Maisha, N., Holland, M., Varley, M., Groynom, R., Hickman, D., … Lavik, E. (2018). Engineering Intravenously Administered Nanoparticles to Reduce Infusion Reaction and Stop Bleeding in a Large Animal Model of Trauma. Bioconjugate Chemistry, acs.bioconjchem.8b00335. http://doi.org/10.1021/acs.bioconjchem.8b00335 Ortel, T. L., Mercer, M. C., Thames, E. H., Moore, K. D., & Lawson, J. H. (2001). Immunologic impact and clinical outcomes after surgical exposure to bovine thrombin. Annals of Surgery, 233(1), 88–96. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/11141230 Ostomel, T. A., Shi, Q., Stoimenov, P. K., & Stucky, G. D. (2007). Metal Oxide Surface Charge Mediated Hemostasis. Langmuir, 23(22), 11233–11238. http://doi.org/10.1021/LA701281T Ostomel, T. A., Stoimenov, P. K., Holden, P. A., Alam, H. B., & Stucky, G. D. (2006). Host-guest composites for induced hemostasis and therapeutic healing in traumatic injuries. Journal of Thrombosis and Thrombolysis, 22(1), 55–67. http://doi.org/10.1007/s11239-006-7658-y Oz, M. C., Rondinone, J. F., & Shargill, N. S. (2003). Floseal Matrix:. New Generation Topical Hemostatic Sealant. Journal of Cardiac Surgery, 18(6), 486–493. http://doi.org/10.1046/j.0886-0440.2003.00302.x Pabinger, I., Fries, D., Schöchl, H., Streif, W., & Toller, W. (2017). Tranexamic acid for treatment and prophylaxis of bleeding and hyperfibrinolysis. Wiener Klinische Wochenschrift, 129(9–10), 303–316. http://doi.org/10.1007/s00508-017-1194-y Palarasah, Y., Nielsen, C., Sprogøe, U., Christensen, M. L., Lillevang, S., Madsen, H. O., … Skjoedt, M.-O. (2011). Novel assays to assess the functional capacity of the classical, the alternative and the lectin pathways of the complement system. Clinical and Experimental Immunology, 164(3), 388–95. http://doi.org/10.1111/j.1365- 2249.2011.04322.x Panagiotopoulos, V., Gizewski, E., Asgari, S., Regel, J., Forsting, M., & Wanke, I. (2009). Embolization of intracranial arteriovenous malformations with ethylene-vinyl alcohol copolymer (Onyx). AJNR. American Journal of Neuroradiology, 30(1), 99– 106. http://doi.org/10.3174/ajnr.A1314 Pasic, M., Unbehaun, A., Drews, T., & Hetzer, R. (2011). Late wound healing problems after use of BioGlue for apical hemostasis during transapical aortic valve implantation. Interactive Cardiovascular and Thoracic Surgery, 13(5), 532–4. http://doi.org/10.1510/icvts.2011.276360

362 Paul, W., & Sharma, C. (2004). Chitosan and alginate wound dressings: a short review. Trends Biomater Artif Organs, 18(1), 18–23. Retrieved from http://medind.nic.in/taa/t04/i1/taat04i1p18.pdf Pawlowski, Christa L; Li, Wei; Sun, Michael; Ravichandran, Kavya; Hickman, DaShawn; Kos, Clarissa; Kaur, Gurbani; Sen Gupta, A. (2017). Platelet microparticle-inspired clot-responsive nanomedicine for targeted fibrinolysis. Biomaterials, 128, 94–108. http://doi.org/10.1016/j.biomaterials.2017.03.012 Pedersen, A. H., Lund-Hansen, T., Bisgaard-Frantzen, H., Olsen, F., & Petersen, L. C. (1989). Autoactivation of human recombinant coagulation factor VII. Biochemistry, 28(24), 9331–9336. http://doi.org/10.1021/bi00450a013 Peev, M. P., Rago, A., Hwabejire, J. O., Duggan, M. J., Beagle, J., Marini, J., … King, D. R. (2014). Self-expanding foam for prehospital treatment of severe intra-abdominal hemorrhage. Journal of Trauma and Acute Care Surgery, 76(3), 619–624. http://doi.org/10.1097/TA.0000000000000126 Peng, T. (2010). Biomaterials for hemorrhage control. Trends in Biomaterials and Artificial Organs, 24(1), 27–68. Permpongkosol, S., Nicol, T. L., Bagga, H. S., Kohanim, S., Kavoussi, L., & Solomon, S. B. (2006). Prophylactic Gelatin Sponge Tract Injection to Prevent Bleeding after Percutaneous Renal Cryoablation in a Swine Model. Journal of Vascular and Interventional Radiology, 17(9), 1505–1509. http://doi.org/10.1097/01.RVI.0000235773.23394.51 Peyvandi, F., Garagiola, I., & Baronciani, L. (2011). Role of von Willebrand factor in the haemostasis. Blood Transfusion = Trasfusione Del Sangue, 9 Suppl 2(Suppl 2), s3-8. http://doi.org/10.2450/2011.002S Pharmacia & Upjohn Company. (2000). CYKLOKAPRON ® tranexamic acid injection Antifibrinolytic agent. Retrieved from https://www.accessdata.fda.gov/drugsatfda_docs/label/2011/019281s030lbl.pdf Phillips, M., Dickneite, G., & Metzner, H. (2003). Fibrin sealants in supporting surgical techniques: strength in factor XIII. Cardiovascular Surgery, 11, 13–16. http://doi.org/10.1016/S0967-2109(03)00066-8 Picker, S. M. (2012). Pathogen Reduction Technologies: The Best Solution for Safer Blood? J Blood Disorders Transf, 3, 133. http://doi.org/10.4172/2155-9864.1000133 Pidcoke, H. F., McFaul, S. J., Ramasubramanian, A. K., Parida, B. K., Mora, A. G., Fedyk, C. G., … Cap, A. P. (2013). Primary hemostatic capacity of whole blood: a comprehensive analysis of pathogen reduction and refrigeration effects over time. Transfusion, 53, 137S–149S. http://doi.org/10.1111/trf.12048 Pierce, A. M., Wiebkin, O. W., & Wilson, D. F. (1984). SurgicelR: its fate following implantation. Journal of Oral Pathology and Medicine, 13(6), 661–670. http://doi.org/10.1111/j.1600-0714.1984.tb01468.x Pilcher, L. (1883). The Treatment of Wounds: Its Principles and Practice, General and Special. Retrieved from

363 https://books.google.com/books?hl=en&lr=&id=FzK6AAAAIAAJ&oi=fnd&pg=PA3 7&ots=4ZYop1aJ1m&sig=eGq6CrJ6y4dyNLTUL6cJbbgqwsY Pipe, S. W. (2008). Recombinant clotting factors. Thromb Haemost, 99, 840–850. http://doi.org/10.1160/TH07-10-0593 PIPE, S. W. (2005). The promise and challenges of bioengineered recombinant clotting factors. Journal of Thrombosis and Haemostasis, 3(8), 1692–1701. http://doi.org/10.1111/j.1538-7836.2005.01367.x Plow, E., & Collen, D. (1981). The presence and release of alpha 2-antiplasmin from human platelets. Blood, 58(6). Retrieved from http://www.bloodjournal.org/content/58/6/1069.short Pohlman, T. H., Walsh, M., Aversa, J., Hutchison, E. M., Olsen, K. P., & Lawrence Reed, R. (2015). Damage control resuscitation. Blood Reviews, 29(4), 251–262. http://doi.org/10.1016/J.BLRE.2014.12.006 Pozza, M., & Millner, R. W. J. (2011). Celox (chitosan) for haemostasis in massive traumatic bleeding. European Journal of Emergency Medicine, 18(1), 31–33. http://doi.org/10.1097/MEJ.0b013e32833a5ee4 Presolski, S. I., Hong, V. P., & Finn, M. G. (2011). Copper-Catalyzed Azide-Alkyne Click Chemistry for Bioconjugation. In Current Protocols in Chemical Biology. Hoboken, NJ, USA: John Wiley & Sons, Inc. http://doi.org/10.1002/9780470559277.ch110148 Prevention, C. for D. C. and. (2014). Deaths and Mortality. Prior, J. J., Wallace, D. G., Harner, A., & Powers, N. (1999). A sprayable hemostat containing fibrillar collagen, bovine thrombin, and autologous plasma. The Annals of Thoracic Surgery, 68(2), 479–485. http://doi.org/10.1016/S0003-4975(99)00552-4 Pusateri, A. E., Delgado, A. V., Dick, E. J., Martinez, R. S., Holcomb, J. B., & Ryan, K. L. (2004). Application of a Granular Mineral-Based Hemostatic Agent (QuikClot) to Reduce Blood Loss After Grade V Liver Injury in Swine. The Journal of Trauma: Injury, Infection, and Critical Care, 57(3), 555–562. http://doi.org/10.1097/01.TA.0000136155.97758.CD Pusateri, A. E., Holcomb, J. B., Harris, R. A., MacPhee, M. J., Charles, N. C., Beall, L. D., & Hess, J. R. (2001). Effect of Fibrin Bandage Fibrinogen Concentration on Blood Loss after Grade V Liver Injury in Swine. Military Medicine, 166(3), 217–222. http://doi.org/10.1093/milmed/166.3.217 Pusateri, A. E., Holcomb, J. B., Kheirabadi, B. S., Alam, H. B., Wade, C. E., & Ryan, K. L. (2006). Making Sense of the Preclinical Literature on Advanced Hemostatic Products. The Journal of Trauma: Injury, Infection, and Critical Care, 60(3), 674– 682. http://doi.org/10.1097/01.ta.0000196672.47783.fd Pusateri, A. E., Kheirabadi, B. S., Delgado, A. V, Doyle, J. W., Kanellos, J., Uscilowicz, J. M., … Modrow, H. E. (2004). Structural design of the dry fibrin sealant dressing and its impact on the hemostatic efficacy of the product. Journal of Biomedical

364 Materials Research. Part B, Applied Biomaterials, 70(1), 114–21. http://doi.org/10.1002/jbm.b.30031 Pusateri, A. E., McCarthy, S. J., Gregory, K. W., Harris, R. A., Cardenas, L., McManus, A. T., & Goodwin, C. W. (2003). Effect of a chitosan-based hemostatic dressing on blood loss and survival in a model of severe venous hemorrhage and hepatic injury in swine. Journal of Trauma, 54(1), 177–182. http://doi.org/10.1097/00005373- 200301000-00023 Qin, Y. (2008). Alginate fibers: an overwiew of the production processes and applications in wound management. Polymer International, 57(April), 171–180. http://doi.org/10.1002/pi Quayle, J. M., & Thomas, G. O. R. (2011). A pre-hospital technique for controlling haemorrhage from traumatic perineal and high amputation injuries. Journal of the Royal Army Medical Corps, 157(4), 419–20. http://doi.org/10.1136/JRAMC-157-04- 17 R, H. M., Battista, O. A., Stark, R. B., & McCord, C. W. (1970). Mirocrystalline collagen as a biologic dressing, vascular prosthesis, and hemostatic agent. Plastic and Reconstructive Surgery, 45(5), 523. RAHE-MEYER, N., & SØRENSEN, B. (2011). Fibrinogen concentrate for management of bleeding. Journal of Thrombosis and Haemostasis, 9(1), 1–5. http://doi.org/10.1111/j.1538-7836.2010.04099.x Rall, J. M., Cox, J. M., Songer, A. G., Cestero, R. F., & Ross, J. D. (2013). Comparison of novel hemostatic dressings with QuikClot combat gauze in a standardized swine model of uncontrolled hemorrhage. Journal of Trauma and Acute Care Surgery, 75, S150–S156. http://doi.org/10.1097/TA.0b013e318299d909 Rathinamoorthy, R., & Sasikala, L. (2011). Polysaccharide Fibers in Wound Management. International Journal of Pharmacy and Pharmaceutical Sciences, 3(3), 38–44. Retrieved from https://innovareacademics.in/journal/ijpps/Vol3Issue3/2282.pdf Ravikumar, M., Modery, C. L., Wong, T. L., Dzuricky, M., & Sen Gupta, A. (2012). Mimicking Adhesive Functionalities of Blood Platelets using Ligand-Decorated Liposomes. Bioconjugate Chemistry, 23(6), 1266–1275. http://doi.org/10.1021/bc300086d Ravikumar, M., Modery, C. L., Wong, T. L., & Gupta, a S. (2012). Peptide-decorated liposomes promote arrest and aggregation of activated platelets under flow on vascular injury relevant protein surfaces in vitro. Biomacromolecules, 13(5), 1495– 1502. http://doi.org/10.1021/bm300192t [doi] Ravikumar, M., Modery, C. L., Wong, T. L., & Sen Gupta, A. (2012). Peptide-Decorated Liposomes Promote Arrest and Aggregation of Activated Platelets under Flow on Vascular Injury Relevant Protein Surfaces in Vitro. Biomacromolecules, 13(5), 1495– 1502. http://doi.org/10.1021/bm300192t

365 RC Becker. (2005). Cell-based models of coagulation: a paradigm in evolution. Journal of Thrombosis and Thrombolysis. Retrieved from https://link.springer.com/content/pdf/10.1007/s11239-005-3118-3.pdf Reddoch, K. M., Pidcoke, H. F., Montgomery, R. K., Fedyk, C. G., Aden, J. K., Ramasubramanian, A. K., & Cap, A. P. (2014). Hemostatic Function of Apheresis Platelets Stored at 4°C and 22°C. Shock, 41, 54–61. http://doi.org/10.1097/SHK.0000000000000082 Reddy, H. L., Dayan, A. D., Cavagnaro, J., Gad, S., Li, J., & Goodrich, R. P. (2008). Toxicity testing of a novel riboflavin-based technology for pathogen reduction and white blood cell inactivation. Transfusion Medicine Reviews, 22(2), 133–53. http://doi.org/10.1016/j.tmrv.2007.12.003 Rhee, P., Brown, C., Martin, M., Salim, A., Plurad, D., Green, D., … Alam, H. (2008). QuikClot Use in Trauma for Hemorrhage Control: Case Series of 103 Documented Uses. The Journal of Trauma: Injury, Infection, and Critical Care, 64(4), 1093–1099. http://doi.org/10.1097/TA.0b013e31812f6dbc Ricklin, D., Hajishengallis, G., Yang, K., & Lambris, J. D. (2010). Complement: a key system for immune surveillance and homeostasis. Nature Immunology, 11(9), 785– 797. http://doi.org/10.1038/ni.1923 Riehemann, K., Schneider, S. W., Luger, T. A., Godin, B., Ferrari, M., & Fuchs, H. (2009). Nanomedicine-Challenge and Perspectives. Angewandte Chemie International Edition, 48(5), 872–897. http://doi.org/10.1002/anie.200802585 Roberts, H., Hoffman, M., & Monroe, D. (2006). A Cell-Based Model of Thrombin Generation. Seminars in Thrombosis and Hemostasis, 32(S 1), 032–038. http://doi.org/10.1055/s-2006-939552 Roberts, H. R. (2001). Recombinant factor VIIa (NovoSeven®) and the safety of treatment. Seminars in Hematology, 38, 48–50. http://doi.org/10.1016/S0037- 1963(01)90148-9 Roberts, I. (2015). Tranexamic acid in trauma: how should we use it? Journal of Thrombosis and Haemostasis, 13, S195–S199. http://doi.org/10.1111/jth.12878 Roberts, I., & Prieto-Merino, D. (2014). Applying results from clinical trials: tranexamic acid in trauma patients. Journal of Intensive Care, 2(1), 56. http://doi.org/10.1186/s40560-014-0056-1 Roberts, I., Prieto-Merino, D., & Manno, D. (2014). Mechanism of action of tranexamic acid in bleeding trauma patients: an exploratory analysis of data from the CRASH-2 trial. Critical Care, 18(6), 1–5. http://doi.org/10.1186/s13054-014-0685-8 Roberts, I., Shakur, H., Coats, T., Hunt, B., Balogun, E., Barnetson, L., … Guerriero, C. (2013). The CRASH-2 trial: a randomised controlled trial and economic evaluation of the effects of tranexamic acid on death, vascular occlusive events and transfusion requirement in bleeding trauma patients. Assessment (Winchester, England), 17(10), 1–79. http://doi.org/10.3310/hta17100

366 Rong, J., Liang, M., Xuan, F., Sun, J., Zhao, L., Zhen, H., … Yu, W. (2015). Alginate- calcium microsphere loaded with thrombin: A new composite biomaterial for hemostatic embolization. International Journal of Biological Macromolecules, 75, 479–488. http://doi.org/10.1016/J.IJBIOMAC.2014.12.043 Roos, A., Bouwman, L. H., van Gijlswijk-Janssen, D. J., Faber-Krol, M. C., Stahl, G. L., & Daha, M. R. (2001). Human IgA Activates the Complement System Via the Mannan-Binding Lectin Pathway. The Journal of Immunology, 167(5), 2861–2868. http://doi.org/10.4049/jimmunol.167.5.2861 Rothwell, S. W., Fudge, J. M., Reid, T. J., & Krishnamurti, C. (2002). ε-Amino caproic acid additive decreases fibrin bandage performance in a swine arterial bleeding model. Thrombosis Research, 108(5–6), 341–345. http://doi.org/10.1016/S0049- 3848(03)00111-7 Rothwell, S. W., Reid, T. J., Dorsey, J., Flournoy, W. S., Bodo, M., Janmey, P. A., & Sawyer, E. (2005). A Salmon Thrombin-Fibrin Bandage Controls Arterial Bleeding in a Swine Aortotomy Model. The Journal of Trauma: Injury, Infection, and Critical Care, 59(1), 143–149. http://doi.org/10.1097/01.TA.0000171528.43746.53 Rothwell, S. W., Settle, T., Wallace, S., Dorsey, J., Simpson, D., Bowman, J. R., … Sawyer, E. (2010). The long term immunological response of swine after two exposures to a salmon thrombin and fibrinogen hemostatic bandage. Biologicals, 38(6), 619–628. http://doi.org/10.1016/J.BIOLOGICALS.2010.07.001 Ruggeri, Z. M., & Mendolicchio, G. L. (2007). Adhesion Mechanisms in Platelet Function. Circulation Research, 100(12), 1673–1685. http://doi.org/10.1161/01.RES.0000267878.97021.ab Ruiz, F., Lea, C., Oldfield, E., & Docampo, R. (2004). Human platelet dense granules contain polyphosphate and are similar to acidocalcisomes of bacteria and unicellular eukaryotes. Journal of Biological Chemistry. Retrieved from http://www.jbc.org/content/279/43/44250.short Ruoslahti, E. (1996). RGD and other recognition sequences for integrins. Annual Review of Cell and Developmental Biology, 12, 697–715. http://doi.org/10.1146/annurev.cellbio.12.1.697 Russo, R. M., Williams, T. K., Grayson, J. K., Lamb, C. M., Cannon, J. W., Clement, N. F., … Neff, L. P. (2016). Extending the golden hour. Journal of Trauma and Acute Care Surgery, 80(3), 372–380. http://doi.org/10.1097/TA.0000000000000940 Ryan, K. L., Cortez, D. S., Dick, E. J., & Pusateri, A. E. (2006). Efficacy of FDA- approved hemostatic drugs to improve survival and reduce bleeding in rat models of uncontrolled hemorrhage. Resuscitation, 70(1), 133–144. http://doi.org/10.1016/J.RESUSCITATION.2005.11.008 Rybak, M. E. M., & Renzulli, L. A. (1993). A Liposome Based Platelet Substitute, the Plateletsome, with Hemostatic Efficacy. Biomaterials, Artificial Cells and Immobilization Biotechnology, 21(2), 101–118. http://doi.org/10.3109/10731199309117350

367 Saltz, R., Sierra, D., Feldman, D., Saltz, M. B., Dimick, A., & Vasconez, L. O. (1991). Experimental and clinical applications of fibrin glue. Plastic and Reconstructive Surgery, 88(6), 1005-15; discussion 1016–7. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/1946751 Sambasivan, C. N., Cho, S. D., Zink, K. A., Differding, J. A., & Schreiber, M. A. (2009). A highly porous silica and chitosan-based hemostatic dressing is superior in controlling hemorrhage in a severe groin injury model in swine. The American Journal of Surgery, 197(5), 576–580. http://doi.org/10.1016/J.AMJSURG.2008.12.011 Sassan, K., Macdougall, M., Lulic, D., Kasasbeh, A., & Levy, M. (2013). Clinical Experience with the Surgicel Family of Absorbable Hemostats (Oxidized Regenerated Cellulose) in Neurosurgical Applications: A Review. Wounds, 25(6), 160–167. Satterly, S., Nelson, D., Zwintscher, N., Oguntoye, M., Causey, W., Theis, B., … Rush, R. M. (2013). Hemostasis in a Noncompressible Hemorrhage Model: An End-User Evaluation of Hemostatic Agents in a Proximal Arterial Injury. Journal of Surgical Education, 70(2), 206–211. http://doi.org/10.1016/J.JSURG.2012.11.001 Sauaia, A., Moore, F., Moore, E., & Moser, K. (1995). Epidemiology of trauma deaths: a reassessment. Journal of Trauma. Retrieved from https://journals.lww.com/jtrauma/Abstract/1995/02000/Epidemiology_of_Trauma_D eaths__A_Reassessment.6.aspx Savage, B., Almus-Jacobs, F., & Ruggeri, Z. M. (1998). Specific Synergy of Multiple Substrate–Receptor Interactions in Platelet Thrombus Formation under Flow. Cell, 94(5), 657–666. http://doi.org/10.1016/S0092-8674(00)81607-4 Sawamura, Y., Asaoka, K., Terasaka, S., Tada, M., & Uchida, T. (1999). Evaluation of Application Techniques of Fibrin Sealant to Prevent Cerebrospinal Fluid Leakage: A New Device for the Application of Aerosolized Fibrin Glue. Neurosurgery, 44(2), 332–337. http://doi.org/10.1097/00006123-199902000-00048 Scheyer, M., & Zimmermann, G. (1996). Tachocomb used in endoscopic surgery. Surgical Endoscopy, 10(5), 501–503. http://doi.org/10.1007/BF00188394 Schiele, U., Kuntz, G., & Riegler, A. (1992). Haemostyptic preparations on the basis of collagen alone and as fixed combination with fibrin glue. Clinical Materials, 9(3–4), 169–177. http://doi.org/10.1016/0267-6605(92)90097-D Schöchl, H., Maegele, M., Solomon, C., Görlinger, K., & Voelckel, W. (2012). Early and individualized goal-directed therapy for trauma-induced coagulopathy. Scandinavian Journal of Trauma, Resuscitation and Emergency Medicine, 20, 15. http://doi.org/10.1186/1757-7241-20-15 Schreiber, G. B., Busch, M. P., Kleinman, S. H., & Korelitz, J. J. (1996). The risk of transfusion-transmitted viral infections. The Retrovirus Epidemiology Donor Study. The New England Journal of Medicine, 334(26), 1685–90. http://doi.org/10.1056/NEJM199606273342601

368 Segal, H. C., Hunt, B. J., & Gilding, K. (1998). The Effects of Alginate and Non-Alginate Wound Dressings on Blood Coagulation and Platelet Activation. Journal of Biomaterials Applications, 12(3), 249–257. http://doi.org/10.1177/088532829801200305 Seghatchian, J., & de Sousa, G. (2006a). Pathogen-reduction systems for blood components: the current position and future trends. Transfusion and Apheresis Science : Official Journal of the World Apheresis Association : Official Journal of the European Society for Haemapheresis, 35(3), 189–96. http://doi.org/10.1016/j.transci.2006.10.002 Seghatchian, J., & de Sousa, G. (2006b). Pathogen-reduction systems for blood components: The current position and future trends. Transfusion and Apheresis Science, 35(3), 189–196. http://doi.org/10.1016/J.TRANSCI.2006.10.002 Seifman, B. D., Rubin, M. A., Williams, A. L., & Wolf, J. S. (2002). Use Of Absorbable Cyanoacrylate Glue To Repair An Open Cystotomy. The Journal of Urology, 167(4), 1872–1875. http://doi.org/10.1016/S0022-5347(05)65252-4 Sen Gupta, A. (2016). Role of particle size, shape, and stiffness in design of intravascular drug delivery systems: insights from computations, experiments, and nature. Wiley Interdisciplinary Reviews: Nanomedicine and , 8(2), 255–270. http://doi.org/10.1002/wnan.1362 Sen Gupta, A. (2017). Bio-inspired Nanomedicine Strategies for Synthetic Blood Substitutes. WIREs Nanomedicine & Nanobiotechnology, 54(4), 1–9. http://doi.org/10.1111/j.1399-0012.2012.01641.x Sen Gupta, A., Ravikumar, M., & Modery‐Pawlowski, C. (2012, April 13). Synthetic platelets. US. Retrieved from https://patents.google.com/patent/US9107845?oq=9107845%2C+2015 Sercombe, L., Veerati, T., Moheimani, F., Wu, S. Y., Sood, A. K., & Hua, S. (2015). Advances and Challenges of Liposome Assisted Drug Delivery. Frontiers in Pharmacology, 6, 286. http://doi.org/10.3389/fphar.2015.00286 Shipman, N., & Lessard, C. S. (2009). Pressure Applied by the Emergency/Israeli Bandage. Military Medicine, 174(1), 086–092. http://doi.org/10.7205/MILMED-D- 00-9908 Shukla, M., Sekhon, U., Betapudi, V., Li, W., Hickman, D., Modery-Pawlowski, C., … Sen Gupta, A. (2016). In vitro characterization and in vivo evaluation of SynthoPlateTM (synthetic platelet) technology in severely thrombocytopenic mice. Journal of Thrombosis and Haemostasis. Shukla, M., Sekhon, U. D. S., Betapudi, V., Li, W., Hickman, D. A., Pawlowski, C. L., … Sen Gupta, A. (2017). In vitro characterization of SynthoPlateTM (synthetic platelet) technology and its in vivo evaluation in severely thrombocytopenic mice. Journal of Thrombosis and Haemostasis, 15(2), 375–387. http://doi.org/10.1111/jth.13579

369 Sierra, D. H. (1993). Fibrin Sealant Adhesive Systems: A Review of Their Chemistry, Material Properties and Clinical Applications. Journal of Biomaterials Applications, 7(4), 309–352. http://doi.org/10.1177/088532829300700402 Silverstein, M. E., & Chvapil, M. (1981). Experimental and clinical experiences with collagen fleece as a hemostatic agent. The Journal of Trauma, 21(5), 388–93. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/7230285 Silverstein, M. E., Keown, K., Owen, J. A., & Chvapil, M. (1980). Collagen fibers as a fleece hemostatic agent. The Journal of Trauma, 20(8), 688–94. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/7401211 Sim, X., Poncz, M., Gadue, P., & French, D. (2016). Understanding platelet generation from megakaryocytes: implications for in vitro-derived platelets. Blood, 127(10), 1227–1233. Retrieved from http://www.bloodjournal.org/content/early/2016/01/19/blood-2015-08- 607929.abstract Simmons, J. W., Pittet, J.-F., & Pierce, B. (2014). Trauma-Induced Coagulopathy. Current Anesthesiology Reports, 4(3), 189–199. http://doi.org/10.1007/s40140-014- 0063-8 Singer, A. J., McClain, S. A., & Katz, A. (2004). A Porcine Epistaxis Model: Hemostatic Effects of Octylcyanoacrylate. Otolaryngology-Head and Neck Surgery, 130(5), 553– 557. http://doi.org/10.1016/j.otohns.2003.09.035 Smith, E. R., Shapiro, G., & Sarani, B. (2016). The profile of wounding in civilian public mass shooting fatalities. Journal of Trauma and Acute Care Surgery, 81(1), 86–92. http://doi.org/10.1097/TA.0000000000001031 Smith, S., Mutch, N., & Baskar, D.-. (2006). Polyphosphate modulates blood coagulation and fibrinolysis. Proceedings of the National Acad Sciences. Retrieved from http://www.pnas.org/content/103/4/903.short Snyder, E. L., & Rinder, H. M. (2003). Platelet Storage — Time to Come in from the Cold? New England Journal of Medicine, 348(20), 2032–2033. http://doi.org/10.1056/NEJMcibr035099 Sohn, V. Y., Eckert, M. J., Martin, M. J., Arthurs, Z. M., Perry, J. R., Beekley, A., … Rush, R. M. (2009). Efficacy of Three Topical Hemostatic Agents Applied by Medics in a Lethal Groin Injury Model. Journal of Surgical Research, 154(2), 258–261. http://doi.org/10.1016/J.JSS.2008.07.040 Solheim, B. G. (2008). Pathogen reduction of blood components. Transfusion and Apheresis Science, 39(1), 75–82. http://doi.org/10.1016/J.TRANSCI.2008.05.003 Sperry, J. L., Guyette, F. X., Brown, J. B., Yazer, M. H., Triulzi, D. J., Early-Young, B. J., … Zenati, M. S. (2018). Prehospital Plasma during Air Medical Transport in Trauma Patients at Risk for Hemorrhagic Shock. New England Journal of Medicine, 379(4), 315–326. http://doi.org/10.1056/NEJMoa1802345

370 Spinella, P. C., & Cap, A. P. (2017). Prehospital hemostatic resuscitation to achieve zero preventable deaths after traumatic injury. Current Opinion in Hematology, 24(6), 529–535. http://doi.org/10.1097/MOH.0000000000000386 Spinella, P. C., Dunne, J., Beilman, G. J., O’Connell, R. J., Borgman, M. A., Cap, A. P., & Rentas, F. (2012). Constant challenges and evolution of US military transfusion medicine and blood operations in combat. Transfusion, 52(5), 1146–1153. http://doi.org/10.1111/j.1537-2995.2012.03594.x Spinella, P. C., Perkins, J. G., Grathwohl, K. W., Repine, T., Beekley, A. C., Sebesta, J., … Holcomb, J. B. (2007). Risks associated with fresh whole blood and red blood cell transfusions in a combat support hospital. Critical Care Medicine, 35(11), 2576–2581. http://doi.org/10.1097/01.CCM.0000285996.65226.A9 Spitalnik, S. L., Triulzi, D., Devine, D. V., Dzik, W. H., Eder, A. F., Gernsheimer, T., … Glynn, S. (2015). 2015 proceedings of the National Heart, Lung, and Blood Institute’s State of the Science in Transfusion Medicine symposium. Transfusion, 55(9), 2282–2290. http://doi.org/10.1111/trf.13250 Spotnitz, W. D. (2010). Fibrin Sealant: Past, Present, and Future: A Brief Review. World Journal of Surgery, 34(4), 632–634. http://doi.org/10.1007/s00268-009-0252-7 Spotnitz, W. D. (2014). Fibrin Sealant: The Only Approved Hemostat, Sealant, and Adhesive-a Laboratory and Clinical Perspective. ISRN Surgery, 2014, 203943. http://doi.org/10.1155/2014/203943 Sum, R., Hager, S., Pietramaggiori, G., Orgill, D. P., Dee, J., Rudolph, A., … Ho, D. (2007). Wound-healing properties of trehalose-stabilized freeze-dried outdated platelets. Transfusion, 47(4), 672–9. http://doi.org/10.1111/j.1537-2995.2007.01170.x Szebeni, J. (2005). Complement activation-related pseudoallergy: A new class of drug- induced acute immune toxicity. Toxicology, 216(2–3), 106–121. http://doi.org/10.1016/j.tox.2005.07.023 Szebeni, J. (2012). Hemocompatibility testing for nanomedicines and biologicals: predictive assays for complement mediated infusion reactions. European Journal of Nanomedicine, 4(1), 33–53. http://doi.org/10.1515/ejnm-2012-0002 Szebeni, J., Baranyi, L., Savay, S., Bodo, M., Morse, D. S., Basta, M., … Alving, C. R. (2000). Liposome-induced pulmonary hypertension: properties and mechanism of a complement-mediated pseudoallergic reaction. American Journal of Physiology- Heart and Circulatory Physiology, 279(3), H1319–H1328. http://doi.org/10.1152/ajpheart.2000.279.3.H1319 Szebeni, J., Bedőcs, P., Csukás, D., Rosivall, L., Bünger, R., & Urbanics, R. (2012). A porcine model of complement-mediated infusion reactions to drug carrier nanosystems and other medicines. Advanced Drug Delivery Reviews, 64(15), 1706– 1716. http://doi.org/10.1016/j.addr.2012.07.005 Szebeni, J., Bedocs, P., Rozsnyay, Z., Weiszhár, Z., Urbanics, R., Rosivall, L., … Barenholz, Y. (2012). Liposome-induced complement activation and related cardiopulmonary distress in pigs: Factors promoting reactogenicity of Doxil and

371 AmBisome. Nanomedicine: Nanotechnology, Biology, and Medicine, 8(2), 176–184. http://doi.org/10.1016/j.nano.2011.06.003 Szebeni, J., Bedőcs, P., Urbanics, R., … R. B.-J. of controlled, & 2012, undefined. (n.d.). Prevention of infusion reactions to PEGylated liposomal doxorubicin via tachyphylaxis induction by placebo vesicles: a porcine model. Elsevier. Retrieved from https://www.sciencedirect.com/science/article/pii/S0168365912001563 Szebeni, J., Fontana, J., Wassef, N., Circulation, P. M.-, & 1999, undefined. (n.d.). Hemodynamic changes induced by liposomes and liposome-encapsulated hemoglobin in pigs: a model for pseudoallergic cardiopulmonary reactions to. Am Heart Assoc. Retrieved from https://scholar.google.com/scholar?hl=en&as_sdt=0%2C36&q=Hemodynamic+chan ges+induced+by+liposomes+and+liposome- encapsulated+hemoglobin+in+pigs&btnG= Szebeni, J., Muggia, F., Gabizon, A., & Barenholz, Y. (2011). Activation of complement by therapeutic liposomes and other lipid excipient-based therapeutic products: Prediction and prevention. Advanced Drug Delivery Reviews, 63(12), 1020–1030. http://doi.org/10.1016/j.addr.2011.06.017 Takeoka, S., Okamura, Y., Teramura, Y., Watanabe, N., Suzuki, H., Tsuchida, E., … Ikeda, Y. (2003). Function of fibrinogen γ-chain dodecapeptide-conjugated latex beads under flow. Biochemical and Biophysical Research Communications, 312(3), 773–779. http://doi.org/10.1016/j.bbrc.2003.10.184 Takeoka, S., Teramura, Y., Ohkawa, H., Ikeda, Y., & Tsuchida, E. (2000). Conjugation of von Willebrand factor-binding domain of platelet glycoprotein Ib alpha to size- controlled albumin microspheres. Biomacromolecules, 1(2), 290–5. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/11710113 Takeoka, S., Teramura, Y., Okamura, Y., Handa, M., Ikeda, Y., & Tsuchida, E. (2001). Fibrinogen-conjugated albumin polymers and their interaction with platelets under flow conditions. Biomacromolecules, 2(4), 1192–7. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/11777392 Takeoka, S., Teramura, Y., Okamura, Y., Tsuchida, E., Handa, M., & Ikeda, Y. (2002). Rolling properties of rGPIbα-conjugated phospholipid vesicles with different membrane flexibilities on vWf surface under flow conditions. Biochemical and Biophysical Research Communications, 296(3), 765–770. http://doi.org/10.1016/S0006-291X(02)00934-8 Teixeira, P. G. R., Inaba, K., Hadjizacharia, P., Brown, C., Salim, A., Rhee, P., … Demetriades, D. (2007). Preventable or Potentially Preventable Mortality at a Mature Trauma Center. http://doi.org/10.1097/TA.0b013e31815078ae Testouri, A., Honorez, C., Barillec, A., Langevin, D., & Drenckhan, W. (2010). Highly Structured Foams from Chitosan Gels. Macromolecules, 43(14), 6166–6173. http://doi.org/10.1021/ma100819j Thiex, R., Williams, A., Smith, E., Scott, R. M., & Orbach, D. B. (2010). The use of Onyx for embolization of central nervous system arteriovenous lesions in pediatric

372 patients. AJNR. American Journal of Neuroradiology, 31(1), 112–20. http://doi.org/10.3174/ajnr.A1786 Thomas, S. (2000). Alginate dressings in surgery and wound management — part 1. Journal of Wound Care, 9(2), 56–60. http://doi.org/10.12968/jowc.2000.9.2.26338 Thon, J. N., Mazutis, L., Wu, S., Sylman, J. L., Ehrlicher, A., Machlus, K. R., … Italiano, J. E. (2014). Platelet bioreactor-on-a-chip. Blood, blood-2014-05-574913. http://doi.org/10.1182/BLOOD-2014-05-574913 Thu, H.-E., Zulfakar, M. H., & Ng, S.-F. (2012). Alginate based bilayer hydrocolloid films as potential slow-release modern wound dressing. International Journal of Pharmaceutics, 434(1–2), 375–383. http://doi.org/10.1016/J.IJPHARM.2012.05.044 Tidrick, R. T., & Warner, E. D. (1944). Fibrin fixation of skin transplants. Surgery, 15(1), 90–95. http://doi.org/10.5555/URI:PII:S0039606044900528 Tokarev, A. A., Butylin, A. A., Ermakova, E. A., Shnol, E. E., Panasenko, G. P., & Ataullakhanov, F. I. (2011). Finite Platelet Size Could Be Responsible for Platelet Margination Effect. Biophysical Journal, 101(8), 1835–1843. http://doi.org/10.1016/J.BPJ.2011.08.031 Toole, J. J., Knopf, J. L., Wozney, J. M., Sultzman, L. A., Buecker, J. L., Pittman, D. D., … Hewick, R. M. (1984). Molecular cloning of a cDNA encoding human antihaemophilic factor. Nature, 312(5992), 342–347. http://doi.org/10.1038/312342a0 Traver, M. A., & Assimos, D. G. (2006). New generation tissue sealants and hemostatic agents: innovative urologic applications. Reviews in Urology, 8(3), 104–11. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/17043707 Travis, J., & Salvesen, G. S. (1983). Human plasma proteinase inhibitors. Annual Review of Biochemistry, 52(1), 655–709. http://doi.org/10.1146/annurev.bi.52.070183.003255 Tyan, Y.-C., Liao*, J.-D., & Lin, S.-P. (2003). Surface properties and in vitro analyses of immobilized chitosan onto polypropylene non-woven fabric surface using antenna- coupling microwave plasma. Journal of Materials Science: Materials in Medicine, 14(9), 775–781. http://doi.org/10.1023/A:1025036421604 Tynngård, N., Lindahl, T. L., & Ramström, S. (2015). Assays of different aspects of haemostasis - what do they measure? Thrombosis Journal, 13, 8. http://doi.org/10.1186/s12959-015-0036-2 Urbanics, R., Bedőcs, P., & Szebeni, J. (2015). Lessons learned from the porcine CARPA model: constant and variable responses to different nanomedicines and administration protocols. European Journal of Nanomedicine, 7(3), 219–231. http://doi.org/10.1515/ejnm-2015-0011 Vaidya, S., Tozer, K. R., & Chen, J. (2008). An overview of embolic agents. Seminars in Interventional Radiology, 25(3), 204–15. http://doi.org/10.1055/s-0028-1085930 van Oostendorp, S. E., Tan, E. C. T. H., & Geeraedts, L. M. G. (2016). Prehospital control of life-threatening truncal and junctional haemorrhage is the ultimate challenge in optimizing trauma care; a review of treatment options and their

373 applicability in the civilian trauma setting. Scandinavian Journal of Trauma, Resuscitation and Emergency Medicine, 24(1), 110. http://doi.org/10.1186/s13049- 016-0301-9 van Raath, M. I., Weijer, R., Nguyen, G. H., Choi, B., de Kroon, A. I., & Heger, M. (2016). Tranexamic Acid-Encapsulating Thermosensitive Liposomes for Site- Specific Pharmaco-Laser Therapy of Port Wine Stains. Journal of Biomedical Nanotechnology, 12(8), 1617–40. http://doi.org/10.1166/JBN.2016.2277 van Zyl, N., Reade, M. C., & Fraser, J. F. (2015). Experimental Animal Models of Traumatic Coagulopathy. Shock, 44(1), 16–24. http://doi.org/10.1097/SHK.0000000000000372 Varga-Szabo, D., Pleines, I., & Nieswandt, B. (2008). Cell Adhesion Mechanisms in Platelets. Arteriosclerosis, Thrombosis, and Vascular Biology, 28(3), 403–412. http://doi.org/10.1161/ATVBAHA.107.150474 Versteeg, H. H., Heemskerk, J. W. M., Levi, M., & Reitsma, P. H. (2013). New Fundamentals in Hemostasis. Physiological Reviews, 93(1), 327–358. http://doi.org/10.1152/physrev.00016.2011 Vournakis, J. N., Demcheva, M., Whitson, a, Guirca, R., & Pariser, E. R. (2004). Isolation, purification, and characterization of poly-N-acetyl glucosamine use as a hemostatic agent. Journal of Trauma-Injury Infection and Critical Care, 57(1, S), S2– S6. http://doi.org/10.1097/01.TA.0000136741.66698.9D Wagner, W. R., Pachence, J. M., Ristich, J., & Johnson, P. C. (1996). Comparativein VitroAnalysis of Topical Hemostatic Agents. Journal of Surgical Research, 66(2), 100–108. http://doi.org/10.1006/JSRE.1996.0379 Wallace, D. G., Cruise, G. M., Rhee, W. M., Schroeder, J. A., Prior, J. J., Ju, J., … Coker, G. C. (2001). A tissue sealant based on reactive multifunctional polyethylene glycol. Journal of Biomedical Materials Research, 58(5), 545–555. http://doi.org/10.1002/jbm.1053 Walsh, M., Shreve, J., Thomas, S., Moore, E., Moore, H., Hake, D., … Castellino, F. (2017). Fibrinolysis in Trauma: “Myth,” “Reality,” or “Something in Between.” Seminars in Thrombosis and Hemostasis, 43(02), 200–212. http://doi.org/10.1055/s- 0036-1597900

Walters, T. J., Wenke, J. C., Kauvar, D. S., McManus, J. G., Holcomb, J. B., & Baer, D. G. (2005). Effectiveness of Self-Applied Tourniquets in Human Volunteers. Prehospital Emergency Care, 9(4), 416–422. http://doi.org/10.1080/10903120500255123 Wang, L. Z., Gorlin, J., Michaud, S. E., Janmey, P. A., Goddeau, R. P., Kuuse, R., … Sawyer, E. S. (2000). Purification of Salmon Clotting Factors and Their Use as Tissue Sealants. Thrombosis Research, 100(6), 537–548. http://doi.org/10.1016/S0049-3848(00)00362-5

374 Wang, M. Y., Armstrong, J. K., Fisher, T. C., Meiselman, H. J., McComb, G. J., & Levy, M. L. (2001). A New, Pluronic-based, Bone Hemostatic Agent That Does Not Impair Osteogenesis. Neurosurgery, 49(4), 962–968. http://doi.org/10.1097/00006123- 200110000-00031 Wang, R., Zhou, B., Liu, W., Feng, X., Li, S., Yu, D., … Xu, H. (2015). Fast in situ generated ɛ-polylysine-poly (ethylene glycol) hydrogels as tissue adhesives and hemostatic materials using an enzyme-catalyzed method. Journal of Biomaterials Applications, 29(8), 1167–1179. http://doi.org/10.1177/0885328214553960 Wang, T., Zhong, X., Wang, S., Lv, F., Zhao, X., Wang, T., … Zhao, X. (2012). Molecular Mechanisms of RADA16-1 Peptide on Fast Stop Bleeding in Rat Models. International Journal of Molecular Sciences, 13(12), 15279–15290. http://doi.org/10.3390/ijms131115279 Ward, K. R., Tiba, M. H., Holbert, W. H., Blocher, C. R., Draucker, G. T., Proffitt, E. K., … Diegelmann, R. F. (2007). Comparison of a New Hemostatic Agent to Current Combat Hemostatic Agents in a Swine Model of Lethal Extremity Arterial Hemorrhage. The Journal of Trauma: Injury, Infection, and Critical Care, 63(2), 276– 284. http://doi.org/10.1097/TA.0b013e3180eea8a5 Weaver, F. A., Hood, D. B., Zatina, M., Messina, L., & Badduke, B. (2002). Gelatin– thrombin-based Hemostatic Sealant for Intraoperative Bleeding in Vascular Surgery. Annals of Vascular Surgery, 16(3), 286–293. http://doi.org/10.1007/s10016-001- 0073-0 Wedmore, I., McManus, J. G., Pusateri, A. E., & Holcomb, J. B. (2006). A Special Report on the Chitosan-based Hemostatic Dressing: Experience in Current Combat Operations. The Journal of Trauma: Injury, Infection, and Critical Care, 60(3), 655– 658. http://doi.org/10.1097/01.ta.0000199392.91772.44 Weisel, J. (2004). The mechanical properties of fibrin for basic scientists and clinicians. Biophysical Chemistry. Retrieved from https://www.sciencedirect.com/science/article/pii/S030146220400198X Wellisz, T., Armstrong, J. K., Cambridge, J., & Fisher, T. C. (2006). Ostene, a new water-soluble bone hemostasis agent. The Journal of Craniofacial Surgery, 17(3), 420–5. http://doi.org/10.1097/00001665-200605000-00006 Wells, P. S. (2002). Safety and efficacy of methods for reducing perioperative allogeneic transfusion: a critical review of the literature. American Journal of Therapeutics, 9(5), 377–88. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/12237729 Whang, H. S., Kirsch, W., Zhu, Y. H., Yang, C. Z., & Hudson, S. M. (2005). Hemostatic Agents Derived from Chitin and Chitosan. Journal of Macromolecular Science, Part C: Polymer Reviews, 45(4), 309–323. http://doi.org/10.1080/15321790500304122 White, G. C., Beebe, A., & Nielsen, B. (1997). Recombinant factor IX. Thrombosis and Haemostasis, 78(1), 261–5. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9198163 White, N. J. (2013). Mechanisms of trauma-induced coagulopathy. Hematology, 2013(1), 660–663. http://doi.org/10.1182/asheducation-2013.1.660

375 WHO | Global Health Estimates. (2018). WHO. Retrieved from http://www.who.int/healthinfo/global_burden_disease/en/ Williamson, D. R., Albert, M., Heels-Ansdell, D., Arnold, D. M., Lauzier, F., Zarychanski, R., … Cook, D. (2013). Thrombocytopenia in Critically Ill Patients Receiving Thromboprophylaxis: Frequency, Risk Factors, and Outcomes. Chest, 144(4), 1207–1215. http://doi.org/10.1378/CHEST.13-0121 Wright, J. K., Kalns, J., Wolf, E. A., Traweek, F., Schwarz, S., Loeffler, C. K., … Eggers, J. (2004). Thermal Injury Resulting from Application of a Granular Mineral Hemostatic Agent. The Journal of Trauma: Injury, Infection, and Critical Care, 57(2), 224–230. http://doi.org/10.1097/01.TA.0000105916.30158.06 Wu, Y., He, J., Cheng, W., Gu, H., Guo, Z., Gao, S., & Huang, Y. (2012). Oxidized regenerated cellulose-based hemostat with microscopically gradient structure. Carbohydrate Polymers, 88(3), 1023–1032. http://doi.org/10.1016/J.CARBPOL.2012.01.058 Xie, X., Tian, J., Lv, F., Wu, R., Tang, W., Luo, Y., … Tang, J. (2013). A novel hemostatic sealant composed of gelatin, transglutaminase and thrombin effectively controls liver trauma-induced bleeding in dogs. Acta Pharmacologica Sinica, 34(7), 983–988. http://doi.org/10.1038/aps.2013.17 Yackel, E. C., & Kenyon, W. O. (1942). The Oxidation of Cellulose by Nitrogen Dioxide. Journal of the American Chemical Society, 64(1), 121–127. http://doi.org/10.1021/ja01253a032 Yamaji-Hasegawa, A., & Tsujimoto, M. (2006). Asymmetric distribution of phospholipids in biomembranes. Biological and Pharmaceutical Bulletin. Retrieved from https://www.jstage.jst.go.jp/article/bpb/29/8/29_8_1547/_article/-char/ja/ Yang, C., Hillas, P. J., Baez, J. A., Nokelainen, M., Balan, J., Tang, J., … Polarek, J. W. (2004). The Application of Recombinant Human Collagen in Tissue Engineering. BioDrugs, 18(2), 103–119. http://doi.org/10.2165/00063030-200418020-00004 Yang, C., Hillas, P., Tang, J., Balan, J., Notbohm, H., & Polarek, J. (2004). Development of a recombinant human collagen-type III based hemostat. Journal of Biomedical Materials Research, 69B(1), 18–24. http://doi.org/10.1002/jbm.b.20030 Yang, L., Stanworth, S., Hopewell, S., Doree, C., & Murphy, M. (2012). Is fresh-frozen plasma clinically effective? An update of a systematic review of randomized controlled trials (CME). Transfusion, 52(8), 1673–1686. http://doi.org/10.1111/j.1537-2995.2011.03515.x Ye, Z., Zhang, H., Luo, H., Wang, S., Zhou, Q., DU, X., … Zhao, X. (2008). Temperature and pH effects on biophysical and morphological properties of self- assembling peptide RADA16-I. Journal of Peptide Science, 14(2), 152–162. http://doi.org/10.1002/psc.988 Yen, R. C. K., Ho, T. W. C., & Blajchman, M. A. (1995). A novel approach to correcting the bleeding associated with thormbocytopenia. Transfusion, 35, 41S.

376 Zahedi, P., Rezaeian, I., Ranaei-Siadat, S.-O., Jafari, S.-H., & Supaphol, P. (2009). A review on wound dressings with an emphasis on electrospun nanofibrous polymeric bandages. Polymers for Advanced Technologies, 21(2), 77–95. http://doi.org/10.1002/pat.1625 Zaragoza, C., Gomez-Guerrero, C., Martin-Ventura, J. L., Blanco-Colio, L., Lavin, B., Mallavia, B., … Egido, J. (2011). Animal Models of Cardiovascular Diseases. Journal of Biomedicine and Biotechnology, 2011, 1–13. http://doi.org/10.1155/2011/497841 Zhang, H., Bré, L. P., Zhao, T., Zheng, Y., Newland, B., & Wang, W. (2014). Mussel- inspired hyperbranched poly(amino ester) polymer as strong wet tissue adhesive. Biomaterials, 35(2), 711–719. http://doi.org/10.1016/J.BIOMATERIALS.2013.10.017 Zhu, G., Mock, J. N., Aljuffali, I., Cummings, B. S., & Arnold, R. D. (2011). Secretory phospholipase A responsive liposomes. Journal of Pharmaceutical Sciences, 100(8), 3146–59. http://doi.org/10.1002/jps.22530 ₂ Zimnitsky, D. S., Yurkshtovich, T. L., & Bychkovsky, P. M. (2004). Synthesis and characterization of oxidized cellulose. Journal of Polymer Science Part A: Polymer Chemistry, 42(19), 4785–4791. http://doi.org/10.1002/pola.20302 Zwaal, R. F. A., Comfurius, P., & Bevers, E. M. (2005). Surface exposure of phosphatidylserine in pathological cells. CMLS Cellular and Molecular Life Sciences, 62(9), 971–988. http://doi.org/10.1007/s00018-005-4527-3 Zwischenberger, J. B., Robert L., R. L. B., Swann, J. R., & Conti, V. R. (1999). Comparison of Two Topical Collagen-Based Hemostatic Sponges During Cardiothoracic Procedures. Journal of Investigative Surgery, 12(2), 101–106. http://doi.org/10.1080/089419399272656

377