novel strategies for improving the pharmacological - Wake ...

568
NOVEL STRATEGIES FOR IMPROVING THE PHARMACOLOGICAL PROPERTIES OF PLATINUM-ACRIDINE ANTICANCER AGENTS BY SONG DING A Dissertation Submitted to the Graduate Faculty of WAKE FOREST UNIVERSITY GRADUATE SCHOOL OF ARTS AND SCIENCES in Partial Fulfillment of the Requirements for the Degree of DOCTOR OF PHILOSOPHY Chemistry December 2015 Winston-Salem, North Carolina Approved By: Ulrich Bierbach, Ph.D, Advisor Martin Guthold, Ph.D, Chair Patricia C. Dos Santos, Ph.D S. Bruce King, Ph.D Mark E. Welker, Ph.D

Transcript of novel strategies for improving the pharmacological - Wake ...

NOVEL STRATEGIES FOR IMPROVING THE PHARMACOLOGICAL

PROPERTIES OF PLATINUM-ACRIDINE ANTICANCER AGENTS

BY

SONG DING

A Dissertation Submitted to the Graduate Faculty of

WAKE FOREST UNIVERSITY GRADUATE SCHOOL OF ARTS AND SCIENCES

in Partial Fulfillment of the Requirements

for the Degree of

DOCTOR OF PHILOSOPHY

Chemistry

December 2015

Winston-Salem, North Carolina

Approved By:

Ulrich Bierbach, Ph.D, Advisor

Martin Guthold, Ph.D, Chair

Patricia C. Dos Santos, Ph.D

S. Bruce King, Ph.D

Mark E. Welker, Ph.D

II

ACKNOWLEDGEMENTS

The work shown in this dissertation would become immeasurably difficult without the

help from many people. First of all, I would like to thank my research advisor, Professor

Ulrich Bierbach, who constantly inspired me and firmly supported my every research

project. His real dedication to science is contagious and encouraged me to be a scientist

in the future.

I also want to express my sincere appreciation to my committee members Dr. Patricia

Dos Santos, Dr. Stephen Bruce King, Dr. Martin Guthold and Dr. Mark Welker for

taking the time to read my dissertation and offering me valuable suggestions during my

Ph.D. study. I would like to thank Dr. Marcus Wright and Dr. Julie Reisz Haines for

support with NMR and mass spectrometry.

I have been fortunate to be able to work with great lab members: Dr. Leigh Ann

Graham, Dr. Jimmy Suryadi, Dr. Amanda Pickard, Dr. Xin Qiao, Ye Zheng, Mu Yang,

and Xiyuan Yao. I also had the opportunity to work with talented undergraduate students:

Ben Fontaine, Zach Hood, Thomas Bartenstein, Layne Raborn and Chris Hackett.

Working with you has been a great experience. To all of my friends in Winston-Salem,

thank you for your constant support over the last five years.

Many thanks to my wife and my parents.

Lastly, I would like to thank Wake Forest University and the National Cancer Institute

(Grant R01 CA101880) for supporting my Ph.D. research, and Wake Forest Innovations

for protecting my intellectual property.

III

TABLE OF CONTENTS

LIST OF FIGURES AND TABLES XIII

LIST OF ABBREVIATIONS XIX

ABSTRACT XXIV

CHAPTER 1. INTRODUCTION 1

1.1. Platinum Based Cancer Chemotherapy 1

1.2. Platinum(II) Based Anticancer Agents 6

1.2.1. Traditional platinum(II) complexes 6

1.2.2. Non-traditional platinum(II) based anticancer agents 9

1.2.2.1. Transplatin analogues 9

1.2.2.2. Monofunctional platinum complexes 12

1.2.2.3. Polynuclear platinum(II) complexes 14

1.2.2.4. A new “lead”: platinum-acridines 17

1.3. Drug Delivery Systems for Platinum Based Anticancer Drugs 20

1.3.1. Polymeric micelles 21

1.3.2. Polymer-based prodrugs 23

1.3.3. Liposomes 24

1.3.4. Nanotubes 30

1.4. Target-Selective Activation of Platinum-Based Anticancer Agents 31

IV

1.4.1. Controlling solution reactivity 32

1.4.2. Activation by ligand exchange 34

1.4.3. Activation by pH-sensitive release 38

1.4.4. Photochemical activation 41

1.4.5. Activation by bioreduction 46

1.5. Goals of the doctoral research 50

CHAPTER 2. USING A BUILDING-AND-CLICK APPROACH

FOR PRODUCING STRUCTURAL AND FUNCTIONAL DIVERSITY

IN DNA-TARGETED HYBRID ANTICANCER AGENTS 53

2.1. Introduction and Design Rationale 54

2.2. Results and Discussion 56

2.2.1. Design of a combinatorial screen assay: assembly and

characterization 56

2.2.2. Validation of the combinatorial screen assay: biological

evaluation and SAR analysis 59

2.3. Conclusion 66

2.4. Experimental Section 66

2.4.1. Reagents and Instrumentation 66

2.4.2. Synthesis and characterization of acridine building blocks 68

2.4.3. Synthesis and characterization of platinum building blocks 74

V

2.4.4. Experimental procedure for “click” reactions 75

2.4.5. Typical procedure for the preparation of platinum–acridine

derivatives 76

2.4.6. Cell proliferation assay 79

CHAPTER 3. USING FLUORESCENT POST-LABELING TO PROBE

THE SUBCELLULAR LOCALIZATION OF DNA-TARGETED

PLATINUM 81

3.1. Introduction and Design Rationale 82

3.2. Results and Discussion 83

3.2.1. Post-labeling method 84

3.2.2. Post-labeling using copper-free click chemistry 92

3.2.3. Pre-labeling method using copper-free click chemistry 95

3.3. Conclusion 100

3.4. Experimental Section 101

3.4.1. Materials and methods 101

3.4.2. Synthetic procedures and product characterization 102

3.4.3. In-gel Alexa Fluor 488 fluorescence detection 104

3.4.4. Stability of compound P1-A9 in buffers 107

3.4.5. Plasmid unwinding experiments 107

3.4.6. CD experiments 107

VI

3.4.7. Cell-based assays based on post-labeling method 108

3.4.7.1. Cell culture maintenance 108

3.4.7.2. Cell treatment and Cu(I)-catalyzed azide-alkyne

cycloaddition (CuAAC) ligation 109

3.4.7.3. Confocal microscopy and fluorescence intensity analysis 110

3.4.8. Cell based assays based on pre-labeling method 111

3.4.8.1. Labeling of compound P1-A9 with Alexa Fluor

488-DIBO to generate P1-A9* 111

3.4.8.2. Cell culture and incubations 111

3.4.8.3. Confocal microscopy 113

CHAPTER 4. SYNTHESIS AND EVALUATION OF A

PLATINUM-ACRIDINE-ENDOXIFEN CONJUGATE 114

4.1. Introduction and Design Rationale 115

4.2. Results and Discussion 117

4.2.1. Efforts to synthesize platinum-acridine-endoxifen conjugates 118

4.2.2. Biological evaluation of target conjugates 119

4.3. Conclusion 122

4.4. Experimental Section 124

4.4.1. Synthesis of target conjugates 124

VII

4.4.2. Stability study of compound P4-A12 133

4.4.3. Cytotoxicity studies and cell viability assays 133

4.4.3.1. Sample preparation and cell culture 133

4.4.3.2. Cytotoxicity studies 134

CHAPTER 5. DESIGN OF ENZYMATICALLY CLEAVABLE

PRODRUGS OF A POTENT PLATINUM-CONTAINING

ANTICANCER AGENT 136

5.1. Introduction and Design Rationale 137

5.2. Results and Discussion 139

5.2.1. Design and chemistry 139

5.2.2. Metal-assisted ester hydrolysis 143

5.2.3. Ester cleavage by human carboxylesterase-2 (hCES-2) 149

5.2.4. Cytotoxicity studies 152

5.3. Conclusion 156

5.4. Experimental Section 157

5.4.1. Materials, general procedures, and instrumentation 157

5.4.2. Synthesis and characterization of target prodrugs 159

5.4.3. Time-dependent NMR spectroscopy 168

5.4.4. Chemical hydrolysis assay 169

VIII

5.4.5. Enzymatic cleavage assay 169

5.4.6. Determination of partition coefficients (logD) 170

5.4.7. Cell based assays 171

5.4.7.1. Cell culture maintenance 171

5.4.7.2. Cytotoxicity assay 172

CHAPTER 6. DEVELOPMENT OF A CHEMICAL TOOLBOX FOR

THE DISCOVERY OF MULTIFUNCTIONAL PLATINUM-ACRIDINE

AGENTS 173

6.1 Introduction 174

6.2 Results and Discussion 177

6.2.1 Optimization of coupling reaction 177

6.2.2. Assembly of fragment library for multifunctional

platinum-acridine conjugates 182

6.2.3 Aqueous conditions for the synthesis platinum-acridine based

conjugates 194

6.2.4. “One-tube” assembly of multifunctional platinum-acridine

conjugates 197

6.2.5 Platinum mediated ester hydrolysis 201

6.3. Conclusions 206

IX

6.4. Experimental Section 208

6.4.1.Materials, general procedures, and instrumentation 208

6.4.2. Synthetic procedures and product characterization 209

6.4.3. Assembly of multifunctional platinum-acridine agents through

a library approach 211

6.4.4. Chemical hydrolysis assay 212

CHAPTER 7. LIPOSOMAL DELIVERY OF

PLATINUM-ACRIDINE ANTICANCER FOR NON-SMALL

CELL LUNG CANCER 214

7.1. Design Rationale 215

7.2. Results and Disscussion 216

7.2.1. Aqueous stability of platinum-acridine analogues 216

7.2.2. Optimization of conditions for drug loading 218

7.2.2.1. Selection of lipids 218

7.2.2.2. The importance of freeze-thaw cycles for drug loading 222

7.2.2.3. The influence of loading buffers on drug loading 223

7.2.2.4. The influence of drug feeds on drug loading 224

7.2.2.5. The influence of loading methods on drug loading 225

7.2.3. Characterization of liposomes 226

X

7.2.4. Stability of liposomes in PBS and during long-term storage 231

7.2.5. In vitro cellular uptake and localization study 235

7.2.6. Acute toxicity studies of liposomal P8-A1 238

7.2.7. in vivo antitumor efficacy 239

7.3. Conclusion 241

7.4. Experimental Section 245

7.4.1. Materials and instruments 245

7.4.2. Optimization of the encapsulation conditions for P8-A1 245

7.4.3. Preparation of liposomes using ethanol injection method 246

7.4.4. Preparation of liposomes using reverse-phase evaporation method 247

7.4.5. Preparation of liposomes for biological evaluation 247

7.4.6. Characterizations of liposomes 248

7.4.6.1. Size and zeta potential of liposomes 248

7.4.6.2. Transmission electron microscopy (TEM) 248

7.4.6.3. Encapsulation efficiency of liposomes 249

7.4.6.4. Determination of lipid composition 249

7.4.7. In vitro stability assay 250

7.4.7.1. Drug leakage under physiological relevant conditions 250

7.4.7.2. In vitro drug release of P8-A1 at elevated temperature 250

XI

7.4.7.3. Storage stability 250

7.4.8. Cellular uptake study of Lipo-1 in NCI-H460 lung cancer cells 251

7.4.9. Animal studies 215

7.4.9.1. Evaluation of maximum tolerated dose (MTD) 252

7.4.9.2. A549 xenograft study 252

CHAPTER 8. SUMMARY AND OUTLOOK 254

REFERENCES 259

APPENDIX A. NMR SPECTROSCOPY 316

APPENDIX B. LC-MS SPECTROSCOPY 380

APPENDIX B1. LC-MS ANALYSIS OF THE REACTIONS

BETWEEN FRAGMENTS “A”AND “P” 380

APPENDIX B2. LC-ESMS ANALYSIS OF PURIFIED

COMPOUNDS 441

APPENDIX B3. LC-ESMS ANALYSIS OF

STABILITY OF CONJUGATES IN BUFFERS 477

APPENDIX B4. METAL-ASSISTED ESTER HYDROLYSIS 485

APPENDIX B5. hCES-2 MEDIATED ESTER HYDROLYSIS 509

APPENDIX B6. LC-MS ANALYSIS OF THE

COUPLING REACTIONS FOR MULTIFUNCTIONAL

XII

PLATINUM-ACRIDINES 516

APPENDIX C. CD SPECTRA 529

APPENDIX D. CONFOCAL MICROSCOPY IMAGES 530

APPENDIX E. CELL PROLIFERATION ASSAYS 533

APPENDIX F. COMPUTATIONAL STUDIES 535

APPENDIX G. MTD STUDIES 536

APPENDIX H. SIZE DISTRIBUTION OF LIPOSOMES 537

APPENDIX I. COUPLING REAGENTS USED IN CHAPTER 6 539

APPENDIX J. LIPIDS USED IN CHAPTER 7 540

SCHOLASTIC VITA 541

XIII

LIST OF FIGURES, SCHEMES AND TABLES

CHAPTER ONE

Figure 1.1. Clinically relevant platinum-containing anticancer agents. 2

Table 1.1. Platinum Complexes and Formulations in Clinical Use and

Clinical Trials 3

Figure 1.2. Cellular uptake, detoxification and active efflux of cisplatin. 6

Figure 1.3. Examples of traditional cisplatin-type complexes. 10

Figure 1.4. Structures of transplatin and active analogues. 11

Figure 1.5. Examples of inactive and active monofunctional platinum agents. 14

Figure 1.6. Examples of polynuclear platinum agents. 17

Figure 1.7. New lead: “Platinum-Acridine”. 20

Figure 1.8. Examples of polymeric prodrugs for platinum drugs. 24

Table 1.2. FDA Approved Liposomal Formulations 27

Table 1.3. Liposomal Formulations for Platinum Drugs 28

Figure 1.9. Examples of nanotube as delivery vectors for platinum drugs. 31

Figure 1.10. Activation of polymer–platinum conjugates by ligand exchange. 36

Table 1.4. Platinum-Based Formulations Activated by Ligand Exchange 37

Figure 1.11. Representative examples of pH-sensitive platinum complexes. 38

Table 1.5. Platinum-Based Formulations Activated by Lysosomal Degradation 40

Figure 1.12. Examples of photochemically active Pt(IV) complexes. 43

XIV

Figure 1.13. Self-assembly of polymeric photoactivated Pt (IV). 44

Table 1.6. Examples and Characteristics of Photoactivatable Pt(IV) Prodrugs 45

Figure 1.14. Delivery of Pt(IV) prodrug with nanoparticles. 47

Table 1.7. Targeted and Multifunctional Cisplatin-Derived Pt(IV) Prodrugs

Activated by Bioreduction 48

Figure 1.15. The schematic illustration of platinum-acridine based conjugates. 50

Figure 1.16. Strategies to constructe platinum-acridine conjugates. 52

CHAPTER TWO

Figure 2.1. Structures of cisplatin and platinum-acridine derivatives. 55

Figure 2.2. Design of the combinatorial screening assay. 55

Figure 2.3. Fragments library. 57

Figure 2.4. High-throughput LC-ESMS analysis of “click” reaction mixtures. 60

Figure 2.5. Biological activity profiles for the 60 compounds. 63

Figure 2.6. Relative potencies of fragments in hybrid agents. . 64

Table 2.1. Biological Activity for Selected Library Members 64

Scheme 2.1. Synthesis of acridine (A) building blocks. 69

Scheme 2.2. Synthesis of platinum (P) building blocks. 74

Scheme 2.3. Synthesis of platinum-acridine derivatives. 76

XV

CHAPTER THREE

Figure 3.1. Bioorthogonal fluorescent labeling of platinum–acridine agents. 84

Figure 3.2. Unwinding studies. 85

Figure 3.3. In vitro evaluation of post-labeling method. 88

Figure 3.4. Post-labeling of platinum-acridine in NCI-H460 lung cancer

cells using Alexa Fluor 488-alkyne. 89

Figure 3.5. Intracellular distribution of P1-A9. 91

Figure 3.6. Post-labeling of platinum-acridine in NCI-H460 lung cancer

cells using copper-free click chemistry. 93

Figure 3.7. Schematic illustration of simultaneous detection of

platinum–acridine using preassembly and post-labeling methods. 94

Figure 3.8. LC-MS analysis of P1-A9*. 98

Figure 3.9. Simultaneous detection of platinum–acridine using preassembly

and post-labeling methods in NCI-H460 cells. 99

Scheme 3.1. Synthesis of precursors 3-3 and 3-5. 102

Table 3.1. Composition of CuAAC Reactions Analyzed by Gel Electrophoresis 106

CHAPTER FOUR

Figure 4.1. Structures of tamoxifen and platinum–acridine hybrids. 116

Figure 4.2. Synthesis of target molecules. 119

XVI

Table 4.1. Cytotoxicity Data in Human Breast Cancer Cell Lines 121

Figure 4.3. Differential response of compounds in MCF-7 cancer cells. 122

Figure 4.4. Synthesis of precursor 4-3. 124

Figure 4.5. Synthesis of precursor A11. 127

CHAPTER FIVE

Figure 5.1. General structure platinum–acridine hybrid agents. 137

Scheme 5.1. Synthesis of target compounds. 139

Figure 5.2. Cleavage of ester moieties in compounds P9-A1─P15-A1. 145

Figure 5.3. LC-MS analysis of reaction products resulting from ester

cleavage in compound P9-A1 in PB, PBS and by hCES-2. 147

Figure 5.4. Kinetics of ester hydrolysis in compound P9-A1. 148

Figure 5.5. LC-MS analysis of reaction products resulting from ester cleavage

in P13-A1 by hCES-2. 151

Figure 5.6. Results of the chemosensitivity screen of compounds

P3-A1 ─ P15-A1 in A549 and NCI-H1435 cancer cells. 155

Scheme 5.2. The synthesis of precursors 5-7 ─ 5-10. 159

CHAPTER SIX

Figure 6.1. General structure of platinum-acridine hybrid agents. 175

Figure 6.2. Structures of ester based platinum-acridine analogues. 178

XVII

Scheme 6.1. Synthesis of target compounds. 180

Table 6.1. Conditions of Coupling Reactions with Model Amines 183

Figure 6.3. Schematic representation of coupling reaction and library members. 185

Figure 6.4. LC-ESMS analysis for the reaction of P16-A1 and 6-7 187

Figure 6.5. Schematic illustration of high throughput drug discovery screen. 190

Table 6.2. Conversion Yield and HR-ESMS Results of Platinum-Acridine Based

Conjugates 191

Scheme 6.2. “Two-step” synthesis of platinum-acridine conjugates under

aqueous conditions. 194

Figure 6.6. LC-ESMS analysis of the reaction mixture for preparation of 6-27. 195

Figure 6.7. Schematic illustration of “one-tube” assembly of multifunctional

platinum-acridine conjugates. 199

Figure 6.8. LC-ESMS analysis of “one-tube” reaction. 200

Figure 6.9. LC-ESMS analysis of the hydrolytic products of 6-11. 202

Figure 6.10. Cleavage of ester moieties in selected conjugates. 204

Figure 6.11. Proposed mechanism of chemical hydrolysis. 205

CHAPTER SEVEN

Figure 7.1. Aqueous stability analysis of P2-A1 at 60 ºC for 4 hours 218

Table 7.1. Conditions for the Liposomal Encapsulation of P8-A1 221

XVIII

Figure 7.2. Factors that influence the encapsulation efficiency of P8-A1 223

Table 7.2. Characterization of Four Liposomal Formulations Prepared for P8-A1 229

Figure 7.3. TEM images of liposomes. 230

Figure 7.4. TEM image of a single liposome and the bilayer discs. 231

Figure 7.5. Sephadex G-25 elution profiles for the mixture of Lipo-1 and P8-A1. 232

Figure 7.6. In vitro leakage of P8-A1 from liposomes. 234

Figure 7.7. In vitro release of P8-A1 from Lipo-1 at elevated temperatures. 234

Figure 7.8. Confocal microscopy studies of subcellular distribution. 237

Figure 7.9. Colocalization images of cells treated with Lipo-1 and P8–A1

co-stained with LysoTracker Red. 238

Table 7.3. MTD Study of P8-A1. 242

Table 7.4. MTD study of liposomal formulations of P8-A1 243

Figure 7.10. Antitumor efficacy and mouse weights in A549 xenograft mice. 244

Table 7.5. Average ng Pt in Organs. 245

XIX

LIST OF ABBREVIATION

A Aden(os)ine

ADME Adsorption, distribution, metabolism, and excretion

BER Base excision repair

Carboplatin cis-Diammine-1,1-cyclobutanedicarboxylatoplatinum(II)

CBDCA cyclobutane-1,1-dicarboxylate

CD Circular dichroism

CDI 1,1'-Carbonyldiimidazole

CHOL Cholesterol

Cisplatin cis-Diamminedichloroplatinum(II)

CMC Critical micelle concentration

CNTs Carbon nanotubes

COMU (1-Cyano-2-ethoxy-2-oxoethylidenaminooxy)dimethylamino-

morpholino-carbenium hexafluorophosphate

CT Calf thymus

CTR1 Copper transporter 1

DACH trans-1,2-diaminnocyclohexane

DCM Dichloromethane

DCC N,N'-Dicyclohexylcarbodiimide

XX

DIPEA N,N-Diisopropylethylamine

DLS Dynamic light scattering

dsDNA Double-stranded DNA

DMAP 4-Dimethylaminopyridine

DMF Dimethylformamide

DMSO Dimethylsulfoxide

DMPC 1-α-Dimyristoylphosphatidylcholine

DMPG l-α-Dimyristoylphosphatidylglycerol

DNA Deoxyribonucleic acid

DPPG 1,2-Dipalmitoyl-sn-glycero-3-phospho-(1'-rac-glycerol)

DOPE Dioleoylphosphatidylethanolamine

DOPC Dioleoylphosphatidylcholine

DSPG Distearoylphosphatidylglycerol

DSPE-mPEG2k 1,2-Distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy

(polyethylene glycol)-2000]

EDC 1-Ethyl-3-(3-dimethylaminopropyl)carbodiimide

EPC Egg phosphatidylcholine

EPR effect Enhanced permeability and retention effect

ER Endoplasmic reticulum

XXI

FBS Fetal bovine serum

G Guan(os)ine

GMP 5´-Guanosine monophosphate

GSH Glutathione

GST Glutathione S-transferase

HBTU (2-(1H-benzotriazol-1-yl)-1,1,3,3-tetramethyluronium

hexafluorophosphate)

hCES-2 Human carboxylesterase 2

HPLC High-performance liquid chromatography

HPMA N-(2-Hydroxypropyl) methacrylamide

HR-ESMS High-resolution electrospray-ionization mass-spectrometry

HSPC L-α-phosphatidylcholine, hydrogenated (Soy)

i.p. Intraperitoneal

i.v. Intravenous

IC50 Inhibitory concentration necessary to reduce cell proliferation by

50% of control

ICD Induced circular dichroism or immunogenic cell death

ICP-MS Inductively-coupled plasma mass spectrometry

LC-ESMS Liquid chromatography-electrospray mass spectrometry

Lobaplatin cis-[Trans-1,2-cyclobutanebis(methylamine)] [(S) lactatoO¹,O¹]

platinum(II)

XXII

m/z Mass-to-charge ratio

MDR Multifactorial drug resistance

MMR Mismatch repair pathway

MPS Mononuclear phagocyte system

MRI Magnetic resonance imaging

MRP Multidrug resistance protein

MTD Maximum tolerated dose

MTS 3-(4,5-Dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-

sulfophenyl)-2H-tetrazolium

MWCO Molecular weight cutoff

Nedaplatin cis-Diammine(glycolato-O¹,O²)platinum(II)

NER Nucleotide excision repair

NMR Nuclear magnetic resonance

NSCLC Non-small-cell lung cancer

nt Nucleotides

OCTs Organic cation transporters

Oxaliplatin trans-L-Diaminocyclohexaneoxalatoplatinum(II)

PDT Photodynamic therapy

PEG Polyethylene glycol

Picoplatin cis-Amminedichlor(id)o-2-methylpyridineplatinum(II)

PyBOP Benzotriazol-1-yl-oxytripyrrolidinophosphonium

hexafluorophosphate

XXIII

PK Pharmacokinetics

PT-ACRAMTU [Pt(en)(ACRAMTU-S)Cl](NO3)2: en = ethane-1,2-diamine,

ACRAMTU = 1-[2-(acridine-9-ylamino)ethyl]-1,3-

dimethylthiourea, acridinium cation

RES Reticuloendothelial system

RNA pol II RNA polymerase II

SAR Structure–activity relationships

Satraplatin Bisacetatoamminedichlorocyclohexylamineplatinum(IV)

SCLC Small cell lung cancer

SEM Standard error of the mean

SWNTs Single-walled nanotubes

TEM Transmission electron microscopy

TMS Tetramethylsilane

TSTU O-(N-Succinimidyl)-N,N,N′,N′-tetramethyluronium

Tetrafluoroborate

XPF Xerodermapig mentosum complementation group F

XXIV

NOVEL STRATEGIES FOR IMPROVING THE PHARMACOLOGICAL

PROPERTIES OF PLATINUM-ACRIDINE ANTICANCER AGENTS

Song Ding

Dissertation under the direction of Ulrich Bierbach, Ph.D.

Professor of Chemistry

ABSTRACT

Unlike traditional cisplatin-type platinum-based anticancer drugs, platinum-acridine

hybrid agents were designed as dual platinating/intercalating DNA-targeted cytotoxics,

which are able to cause cancer cell death at low-nanomolar concentrations.

Unfortunately, the preclinical development of these agents has been hampered by their

severe systemic toxicity. The goal of this dissertation was to devise strategies that

improve the drug-like properties of platinum-acridines and allow their safe delivery. To

achieve this, several classical and newly developed synthetic methodologies have been

used to generate functionally unique hybrid agents. Several model systems, whole-cell

assays, and animal studies have been used in this dissertation to validate their design and

demonstrate their utility as anticancer agents. A modular screening platform was

developed, based on a platinum-mediated amine-to-nitrile addition reaction, for rapid

identification of functionalized platinum-acridine agents. These pre-screening assays

produced functionalized “warheads” while providing insight into structure–activity

relationships (SAR). Using several library members, we set out to explore synthetic

XXV

approaches to construct platinum-acridine-based conjugates. A chemically robust azide-

modified platinum-acridine was selected to validate the feasibility of copper-mediated

and copper-free click chemistry as platinum-compatible conjugation reactions. This

chemistry was used to attach fluorescent molecules to detect platinum-acridines in cancer

cells by confocal microscopy. Both fluorophore tagging prior to incubation with cells and

post-labeling methods were explored. In addition, a hydroxyl-modified warhead was

conjugated with endoxifen via a chemically stable carbamate bond to produce a highly

active hybrid agent in estrogen receptor positive (ER+) breast cancer. In another study,

lipophilic ester-based prodrugs of platinum–acridines were generated showing improved

drug-like properties (e. g., partition coefficients, logD). Two distinct pathways by which

the target compounds can be activated have been confirmed by LC-ESMS and/or NMR

techniques: (i) a platinum-assisted, self-immolative ester cleavage in a low-chloride

environment, and (ii) enzymatic cleavage by human carboxylesterase-2 (hCES-2). Highly

efficient amide coupling reactions in platinum complexes were also developed. This

modular approach can be used to assemble a diverse library of platinum-acridines

containing other bioactive components, such as molecularly targeted therapies, targeted

ligands, and chemosensitizers. Finally, liposomal encapsulation of platinum-acridine was

achieved, and the formulations were evaluated in A549 lung cancer xenograft models in

mice. Improved anticancer efficacy of one of the liposomal formulations compared with

the free drug was observed in this assay. In conclusion, the research in this dissertation

has laid the foundation for the future preclinical development of platinum-acridines as

oncology drugs by devising new synthetic methodology and providing proof-of-concept

data in clinically relevant models of cancer.

1

CHAPTER 1

INTRODUCTION

1.1 Platinum based cancer chemotherapy

Since the serendipitous discovery of the anticancer activity of cisplatin, or cis-

diamminedichloroplatinum(II), platinum based drugs became the mainstay of

chemotherapy for years. Table 1.1 and Figure 1.1 summarize relevant drugs and their

liposomal formulations in routine clinical use, as well as agents that are currently being,

or have been, evaluated in clinical trials. Cisplatin was the first platinum complex

approved for the treatment of a broad range of cancers including non-small cell lung

cancer, small cell lung cancer, testicular, ovarian and bladder cancer [1]. It is widely

accepted that the intracellular activation of platinum drugs through aquation, which is

sensitive to the concentration of physiological chloride ions, is important for subsequent

DNA platination, with the preference to coordinate at the N7 position of guanine [2]

(Figure 1.2). The induced conformational distortions in duplex DNA are recognized by

many proteins that are involved in replication, transcription, and DNA damage repair,

leading to aberrant signal transduction and consequential cell apoptosis [2, 3]. Aside from

DNA associated cytotoxicity, recent evidence indicates that oxaliplatin, but not cisplatin,

may be involved in modulating immune response by triggering immunogenic cell death

(ICD) [4, 5]. It has been proposed that such a mechanism, in which the cancer cells killed

by oxaliplatin act as cancer vaccines, may eliminate the residual, highly resistant cancer

2

cells and exert a long-lasting anticancer effect. These compelling findings prompted a

new therapeutic paradigm for platinum based anticancer drugs [6, 7].

Figure 1.1. Clinically relevant platinum-containing anticancer agents. For current status,

refer to Table 1.1.

3

Table 1.1. Platinum Complexes and Formulations in Clinical Use and Clinical Trials [1,

8]

Complex/Formulation Status Major Indications

Cisplatin Worldwide clinical use Metastatic testicular and ovarian

tumors, advanced bladder cancer

Carboplatin Worldwide clinical use Advanced ovarian carcinoma

Oxaliplatin Worldwide clinical use Metastatic colorectal cancer

Nedaplatin Clinical use in Japan Small and non-small cell lung

cancer, head and neck tumors,

esophageal and bladder tumors,

cervix carcinomas

Lobaplatin Clinical use in China Breast, testicular, ovarian, small

cell lung and gastric carcinomas,

chronic myeloid leukemia

Heptaplatin Clinical use in South

Korea

Gastric cancer

Picoplatin Phase II clinical trials

Phase III clinical trials

Metastatic colorectal cancer,

metastatic castration-resistant

prostate cancer refractory or

resistant ovarian cancer,

refractory or progressed SCLC

BBR3464 Phase II clinical trials Gastric and esophageal

adenocarcinoma

Satraplatin Phase II clinical trials Metastatic, androgen-

independent prostate cancer

Lipoplatin (cisplatin

liposome)

Phase III clinical trials NSCLC, breast cancer, gastric

cancer

SPI-77 (cisplatin

liposome)

Phase II clinical trials Advanced NSCLC, refractory

ovarian cancer

MBP-426 (oxaliplatin

liposome)

Phase II clinical trials Gastric and esophageal

adenocarcinomas

4

Unfortunately, the efficacy of cisplatin is limited by its notorious systemic toxicities

such as nephrotoxicity, neurotoxicity, ototoxicity and myelosuppression [9]. The reasons

for these symptoms are multifold, but generally can be attributed to a lack of tumor

selectivity of current platinum drugs. Additional limitations of platinum-based drugs are

intrinsic and acquired tumor resistance [3]. Lack of responsiveness of platinum drugs can

be caused by insufficient cellular accumulation [10]. Specific pathways for platinum

drugs to enter into cancer cells include passive diffusion and uptake by cell membrane

transporters [10] (Figure 1.2). Multiple transporters were found to participate in the active

influx of platinum drugs, such as copper transporter 1 (CTR1) [11] and organic cation

transporters (OCTs) [12]. Factors that impede cellular uptake result in reduced levels of

platinum in cancer cells, which is correlated with the ineffectiveness of these drugs [10].

Active efflux, likewise, plays an important role in platinum drug resistance. Cancer cells

with overexpressed efflux proteins, including multidrug resistance 1 (MDR1), multidrug

resistance protein-1 (MRP1, also known as ABCC1), multidrug resistance protein-2

(MRP2, also known as ABCC2) and copper efflux transporters ATP7A and ATP7B, may

lose their sensitivity to platinum drugs [10]. Once platinum complexes have entered into

cancer cells, detoxification by various cytoplasmic thiol-containing nucleophiles is

another barrier for them to overcome before reaching their cellular target, nuclear DNA

[13]. The major scavengers in this category is glutathione, the tripeptide that avidly binds

to platinum center via cysteine sulfur, which is likely facilitated by glutathione S-

transferase (GST) [14]. Likewise, elevated levels of thiol-containing proteins, such as

metallothioneins, may lead to platinum drug resistance as well [3].

5

The resistance to platinum drugs may also be triggered after the formation of platinum-

DNA adducts. In the case of cisplatin, the major adducts are intrastrand 1,2-crosslinks,

which bend the DNA duplex significantly towards the major groove. This distortion

produces a wide minor groove, which is preferentially recognized by DNA damage

recognition proteins [3]. An extensive body of evidence suggests the involvement of

DNA repair machineries in causing increased tolerance to platinum-induced DNA

damage by removing platinum-DNA adducts, including nucleotide-excision repair (NER),

base excision repair (BER), mismatch repair (MMR) and double-strand-break repair.

NER is the major mechanism responsible for the removal of cisplatin induced DNA

lesions. As a result, the levels of the key proteins in this pathway, such as ERCC1

(excision repair cross-complementing-1) and XPF (xerodermapig mentosum

complementation group F), are being used as predicative biomarkers of cisplatin

resistance in the clinic [15, 16]. Cancer cells with deficient DNA repair systems were

found to be hypersensitive to platinum drugs. Abnormal apoptotic signaling pathways,

moreover, which are associated with various proteins such as p53, anti-apoptotic and pro-

apoptotic proteins, may result in platinum drug resistance in cancer cells [17].

To overcome the abovementioned limitations, several strategies have been proposed for

safer and more effective platinum based anticancer agents. The first approach is based on

the assumption that certain types of platinum-DNA adducts, which cause unique

conformational distortions, may be biologically more significant than others. To tune the

structural distortions produced by platinum(II) in dsDNA, variation was made to (i) the

ligand sets (non-leaving groups), (ii) the stereochemistry, (iii) the nuclearity, (iv) and the

ability of the complexes to form cross-links vs. monofunctional adducts (see 1.2.1).

6

Thousands of derivatives have been synthesized and tested since the FDA approval of

cisplatin. In another approach, cisplatin-type agents have been linked to tumor selective

carriers, such as folic acid, polymers and nanoparticles. In addition, multifunctional

platinum agents, which contain other bioactive molecules may help to overcome

resistance mechanisms and sensitize tumors to these drugs. Some examples of these

strategies will be discussed in the later parts of this dissertation.

Figure 1.2. Cellular uptake, detoxification and active efflux of cisplatin.

1.2. Platinum(II) Based Anticancer agents

1.2.1 Traditional platinum (II) complexes

Inspired by cisplatin, a large number of cisplatin analogues have been produced and

evaluated as effective anticancer candidates with reduced systemic toxicity [1]. Like

7

cisplatin, these analogues contain two leaving ligands, which leads to the formation of

similar intrastrand or interstrand crosslinks in DNA, thereby inhibiting DNA replication

and transcription. Modifications in the coordination sphere of cisplatin include variations

of non-leaving ammine groups and leaving ligands. The replacement of leaving ligands

affects the aquation kinetics of platinum complexes (will be discussed in detail in 1.4.2),

which, in turn, has a significant impact on their tissue distributions and side effects. This

idea led to the development of carboplatin (Figure 1.1), a second-generation cisplatin

analogue bearing dicarboxylate as the leaving groups. Carboplatin shows less toxic side

effect than cisplatin, even though it produces identical platinum-DNA adducts. The

success of carboplatin spurred the design of many platinum complexes containing

(di)carboxylate leaving groups [18]. Alternative leaving ligands such as acetylacetonate

(acac) and β-diketonates (Figure 1.3, 1-1, 1-2) have also been reported in platinum

anticancer agents [19].

The anticancer efficacy of platinum drugs is closely associated with the platinum-DNA

adducts they produce. Variations of the non-leaving ammine groups of cisplatin can lead

to platinum-DNA adducts with disparate steric hindrance and conformational changes in

target DNA, which might not be recognized by the proteins that contribute to cisplatin

induced cytotoxicity. Therefore, replacement of one or two of the non-leaving ligand may

help to create platinum complexes that maintain sensitivity in tumors resistant to cisplatin.

One successful example is oxaliplatin (Figure 1.1), in which DACH (trans-1,2-

diaminnocyclohexane) was selected as the non-leaving group. The bulky DACH ring in

oxaliplatin occupies most of the DNA major groove, leaving a narrow groove surface and

thus resulting in platinum-DNA adducts processed differently from cisplatin [3].

8

Oxaliplatin has been approved worldwide for the treatment of colorectal cancer and

showed no cross-resistance with cisplatin. It should be noted that the non-leaving ligands

also influence the lipophilicity [18] and the aquation rate [18, 20] of platinum complexes,

which overall can aid to circumvent platinum drug resistance. In particular, picoplatin

(Figure 1.1) was found to overcome glutathione-mediated detoxification due to its bulky

2-methyl pyridine (picoline) ring. In picoplatin, the methyl group is situated right above

the metal center, which protects platinum from attack from intracellular sulfur

nucleophile. Several examples of traditional cisplatin type complexes are summarized in

Figure 1.3.

Lack of tumor selectivity is a major reason for the severe systemic toxicity of cisplatin.

To produce tumor targeted platinum-based anticancer agents, molecules with a high

affinity for cancer cells have been incorporated into platinum complexes as non-leaving

ligands. These ligands include peptides [21] (Figure 1.3, 1-8), estradiol (Figure 1.3, 1-9

and 1-10) [22-24], and folic acid (Figure 1.3, 1-11) [25]. The presence of these ligands

allows the platinum complexes to be recognized by receptors overexpressed on the cancer

cell surfaces and enhance their uptake into tumors. Targeting moieties can also be

introduced as leaving groups in platinum complexes. Since most of them are bulky, the

dissociation of the platinum moiety from the tumor targeting ligands in cancer cells

assures that the carrier does not interfere with the DNA binding of the platinum

“warheads”. Examples of these platinum complexes are shown in Figure 1.3. Apart from

these tumor selective carriers, bioactive ligands, such as tarmoxifen analogues [26]

(Figure 1.3, 1-12), were installed in platinum complexes, which were expected to act

9

synergistically with platinum complexes to sensitize resistant breast cancers to platinum-

based agents.

1.2.2. Non-traditional platinum(II)-based anticancer agents

1.2.2.1. Transplatin analogues

The reasons for the poor anticancer properties of the trans isomer of cisplatin,

transplatin (Figure 1.4), have been investigated for decades [27]. Unlike its cis congener,

transplatin is unable to form stable intrastrand 1,2-crosslinks in the DNA helix due to

geometric constrains [28], while the formation of intrastrand 1,3-crosslinks has been

observed [29]. The major platinum-DNA adducts produced by transplatin are

monofunctional, and only a small amount of them slowly convert into interstrand

crosslinks [27]. This can be explained by the solution chemistry of transplatin. The

replacement of the first chloride with water in transplatin is kinetically favored due to a

mutual trans effect of the chloride ligands, however, the removal of the second chloride

ligand trans to a ligand, low in the trans-effect series such as water or DNA nitrogen

appears to be difficult [30]. The ineffectiveness of transplatin, despite its ability to

modify DNA, emphasizes the importance of the type of platinum-DNA adduct for the

biological outcome of platinum-based anticancer drugs. The higher chemical reactivity of

transplatin also facilitates the interaction with many intracellular nucleophiles, such as

GSH, which accelerates the deactivation process and results in poor biological activity

[28].

10

Fig

ure

1.3

. E

xam

ple

s of

trad

itio

nal

cis

pla

tin

-type

com

ple

xes

.

11

Active trans-platinum complexes can be generated by substituting one or two NH3

ligands with steric hindered amines [27]. These bulky ligands are able to reduce the

reactivity of transplatin analogues in blood and stabilize the platinum-DNA adducts they

produce. Representative examples of active transplatin analogues are shown in Figure 1.4.

The first cytotoxic transplatin analog reported (trans-[PtCl2(NH3)(pyridine)]) contains a

pyridine ring and exhibits comparable cytotoxicity with cisplatin [31], which inspired the

synthesis of numerous platinum complexes with trans geometry bearing planar

heterocyclic ligands, such as quinoline, isoquinoline, and thiazole [27]. The platinum-

DNA adducts produced by these analogues include monofunctional adducts, interstrand

and intrastrand crosslinks [27, 28], which elicit biological responses distinct from

cisplatin and in cell lines overcome resistance to platinum drugs used in the clinic [28].

Other ligands containing N-donor groups, for instance, iminoethers [32], aliphatic amines

[33] and heterocyclic aliphatic amines [34], have been used in the design of transplatin

analogues as well, most of which display cytotoxicity in the micromolar range even in

cell lines resistant to cisplatin [35]. Another intriguing feature of transplatin analogues

lies in their ability to induce ternary DNA-protein adducts, which also potentially

contribute to their cytotoxicity [35].

12

Figure 1.4. Structures of transplatin and active analogues (ipa = isopropyl amine).

1.2.2.2. Monofunctional platinum complexes

Early empirical structure-activity relationships (SAR) ruled out the possibility of using

monofunctional platinum complexes as clinically useful anticancer agents [36]. Most of

the monofunctional platinum complexes share the general formula cis-[Pt(Am)2(L)Cl]+,

in which one chloride ligand is the only labile ligand, which leads to the formation of

monofunctional adducts with DNA. Early efforts to develop monofunctional platinum

complexes resulted in many complexes unfortunately inactive both in vivo and in vitro

[37-40]. Some of them are listed in Figure 1.5 (complexes 1-20, 1-21, 1-22, 1-23). The

concept of monofunctional platinum complexes, however, was recently revisited. The

13

cationic complex cis-[Pt(NH3)2(Pyridine)Cl]+, also known as pyriplatin, was used in a

study to demonstrate that the cytotoxicity of cationic platinum complexes correlates with

the level of organic cation transporters (OCTs) expressed by colon cancer cells [41].

Although much less potent than cisplatin in a number of cell lines, pyriplatin exhibited a

distinct spectrum of activity [41, 42]. The DNA adduct formed by pyriplatin (1-24),

revealed in a crystal structure of a site-specifically platinated DNA, cause less local

distortions in dsDNA compared with crosslinking platinum anticancer agents. It was

proposed that this feature may protect the adducts from removal by the DNA repair

machinery [41, 43] and overcome tumor resistance. Additional mechanistic studies

revealed a relationship between the cytotoxicity of pyriplatin and its capacity to stall

RNA polymerase II (RNA pol II), subsequently inducing transcription inhibition and cell

death [41]. The crystallographic analysis suggested the increase in the steric bulk of the

amine ligand may increase the transcription inhibition. This finding prompted a drug

screen based on a small library of monofunctional platinum complexes bearing

heterocyclic ammine ligands with various steric hindrance [44]. In this study, the

phenanthridine derivative, cis-[Pt(NH3)2(phenanthridine)Cl]+ (NO3-), (phenanthriplatin

(1-25)), was the most potent analogue identified, and its unique anticancer spectrum was

confirmed in the NCI-60 cancer cell line panel [44-46]. The cell killing ability of

phenanthriplatin, on the basis of IC50 values, is 4-40 times higher than that of cisplatin.

The discovery of phenanthriplatin led to the development of other monofunctional

platinum complexes, such as SA-PT(1-26) [47] shown in Figure 1.5.

14

Figure 1.5. Examples of inactive and active monofunctional platinum agents.

1.2.2.3. Polynuclear platinum(II) complexes

Polynuclear platinum(II) complexes represent a novel class of platinum based drugs

with structural features derived from cisplatin [48, 49]. BBR3464 (Figure 1.1) is one of

the most successful analogues in this class, which is a trinuclear drug composed by two

separated monofunctional trans-[PtCl(NH3)2]+ moieties bridged by a tetra-amine trans-

15

[Pt(NH3)2(NH2(CH2)6NH2)]2+ linker[49]. Long range (Pt,Pt) inter and intrastrand

crosslinks were found in the DNA adducts produced by BBR3464, in which the two

platinated DNA bases are separated by up to 4 base pairs. In particular, the intrastrand

crosslinks are formed in both 3’→3’ and 5’→5’ directions, which are distinctly different

from other known platinum-DNA adducts [50]. This novel DNA binding mode finally

leads to Z-DNA like structure around the platinated sites. This type of distortion is not

recognized by high mobility group HMG1 proteins, which are known to contribute

substantially to the cytotoxicity of cisplatin [50]. BBR3464 showed higher potency than

cisplatin when tested in a broad range of tumor cell lines and overcomes resistance to the

clinical drugs [51]. The effective doses in xenografts in mice ranged from 0.3-0.6 mg/kg,

which caused significant tumor regression in several tumor models [51]. BBR3464 is the

only non-cisplatin platinum complex that entered into phase II clinic trials with

manageable toxicity,however, its narrow therapeutic index prevented it from advancing

to the next stage [48].

Variations in the tetramine platinum linkers in BBR3464 with polyamines gave rise to

exceptionally cytotoxic dinuclear analogues [49, 50]. Notably, the positive charges and

the ability to form hydrogen bonds of these connectors seemed to play important roles in

determining the anticancer efficacy. One lead complex is BBR 3610 (1-27,Figure 1.6)

with a spermine analogue as the linker, which was designed to perform in a similar way

to BBR3464 [52]. Other alternative bridging ligands include steric hindered polyamine

(1-29) [53],aromatic (1-30) [54] and heterocyclic ring systems (1-31 and 1-32,Figure

1.6) [55]. Replacement of the leaving ligand chloride with NH3 or other dangling

amines resulted in a new class of substitution-inert platinum complexes [48]. The lead

16

complex in this class is TriplatinNC, of which the square planar tetraamine platinum

moiety is able to form “bidentate” N-O-N hydrogen bonds with the oxygen in phosphate.

This distinct non-covalent binding mode, known as “phosphate clamps” [56], displays

high affinity with DNA and leads to nucleic acid compaction [57]. The prototype

TriplatinNC (1-33,Figure 1.6) shows comparable cytotoxicity to cisplatin in a panel of

tumor cell lines independent of p53 status and the level of GSH in cells [58, 59]. High

cellular accumulation acts as another parameter contributing to the cytotoxicity of

TriplatinNC [60]. The uptake of the substitution-inert, highly charged platinum

complexes is mediated by heparin sulfate proteoglycan (HSPG) [60, 61], since HSPG is

overexpressed in tumor cells, this unique recognition mechanism may confer tumor

selectivity to TriplatinNC analogues. Moreover, the nucleolar localization of TriplatinNC

was also determined [62, 63], linking its anticancer effect to the inhibition of RNA

polymerase I mediated synthesis of ribosomal RNA (rRNA) [62].

17

Figure 1.6. Examples of polynuclear platinum agents.

1.2.2.4. A new lead: platinum-acridines

The idea to generate DNA adducts distinct from the traditional bifunctional 1,2-

intrastrand cross-links has led to the discovery of many new types of non-classical

platinum anticancer complexes with different DNA binding modes. Inspired by these

designs, attempts to combine DNA intercalation and platination in one molecule led to

the creation of platinum-acridine agents [64-66]. One prototypical complex in this class is

PT-ACRAMTU ([PtCl(en)(ACRAMTU)](NO3)2, en = ethane-1,2-diamine, ACRAMTU

= 1-[2-(acridin-9-ylamino)ethyl]-1,3-dimethylthiourea, Figure 1.7, 1-34) [65], which

18

features a single chloro leaving group and a sulfur donor containing linker that connects

the intercalator to the metal, allowing the formation of monofunctional adducts. The best

biological activity was achieved when N-methylethylenediamine was selected as the

main component of the linker (~ 4 fold more potent than cisplatin in NCI-H460 cells),

which is flexible enough in this case to platinate DNA base pairs adjacent to interaction

sites, and thus favor the dual-binding mode involving both DNA intercalation and

platination [67]. Extension of the linker chain reduces the rate of platination step, and

results in reduced cytotoxicity [68]. Other properties of PT-ACRAMTU, such as the

cationic nature and DNA affinity of the acridinium moiety, seem to allow efficient uptake

into cancer cells and in nuclear accumulation and rapid DNA binding. PT-ACRAMTU

proved to be more active than cisplatin confirmed by clonogenic growth and cell

proliferation assays, with IC50 in the low-micromolar and submicromolar range across a

broad range of cancer cells, including ovarian, lung, colon, breast, pancreatic, and brain

cancers [65, 69-71]. Submicromolar activity of PT–ACRAMTU was also found in cancer

cells with aberrant p53 and k-ras gene status, which are typically resistant to cisplatin

[72].

Unfortunately, PT-ACRAMTU was unable to slow tumor growth in a mouse xenograft

model. This prompted extensive SAR studies to identify second generation platinum-

acridines with enhanced activity in vitro and in vivo. Structural modifications were made

to the platinum–intercalator linker, the acridine moiety itself, the non-leaving groups, and

the donor group through which the intercalator is attached to the metal (Figure 1.7B) [70,

73-75]. The most striking cytotoxic enhancement was observed when the thiourea moiety

in PT-ACRAMTU was replaced by amidine as the donor group, resulting in an up to 500-

19

fold enhanced potency compared with clinical drug cisplatin in NSCLC cell line NCI-

H460 (Figure 1.7C, 1-35) [76]. Low nanomolar activity was also found in other cell

lines[68]. The significantly improved activity of these hybrids in vitro has been attributed

to their ability to produce adducts in double-stranded DNA more rapidly than either PT–

ACRAMTU or cisplatin [77]. In addition, unlike bifunctional cisplatin type complexes,

only minor DNA conformational perturbations were induced by the DNA adducts of

platinum-acridines, as confirmed in an NMR solution structural study [78], suggesting

that these DNA adducts are recognized and processed differently by intracellular proteins.

Although the superior cytotoxicity of the second-generation platinum-acridines finally

translated into inhibition of tumor growth in vivo, like other platinum-based anticancer

agents, they remained quite toxic to test animals [76]. This drawback obviously restricts

the development of platinum-acridines as anticancer drugs. To address this limitation,

tumor-targeted approaches and prodrug concepts will be explored in this dissertation in

an attempt to turn the platinum–acridine hybrid agents into systemically less toxic tumor

selective warheads.

20

Figure 1.7. (A) Structure of PT-ACRAMTU. (B) Variations in platinum-acridine

analogues for structure-activity relationship (SAR) studies. (C) Cytotoxcity enhancement

in the first and second generation of platinum-acridines.

1.3. Drug Delivery Systems for Platinum Based Anticancer Drugs

Like other cytotoxics, the “platinums” often show unfavorable pharmacokinetic

properties, which result in inefficient delivery to the diseased tissue and undesired off-

target effects. To overcome these drawbacks, various drug delivery carriers have been

developed in which platinum drugs are chemically attached to tumor-targeting carriers or

physically encapsulated inside nano-sized particles. Targeted delivery vehicles developed

for platinum-based anticancer drugs have recently been reviewed in the literature [79, 80].

The effectiveness of a drug delivery system depends on the nature of the carrier materials,

which alter the pharmacokinetic properties of the cytotoxic payload and may promote

21

selective accumulation in tumor tissue. A wide range of carriers have been introduced to

deliver platinum-based agents, including peptides, receptor ligands, and monoclonal

antibodies, which are actively recognized by cancer cells. Polymers and nanoparticles

enhance the passive accumulation of platinum drugs at the tumor site by exploiting the

enhanced permeability and retention (EPR) effect. This effect, termed “passively targeted”

in the literature, has been confirmed in animal models [81] and observed in patients who

are treated with intravenous liposomal doxorubicin (Doxil) [82]. It should be noted

though that the exact role of the EPR effect in the treatment in humans is still being

debated [81]. To further improve tumor uptake of cytotoxic platinum, tumor-targeting

ligands have been engineered on the surface of nano-sized drug delivery vectors. The

selective interactions between the ligands and the receptors, which are commonly

overexpressed in tumor sites, are proposed to mediate endocytosis, and improve the

therapeutic index [83]. Nevertheless, the advantages of targeted over non-targeted

nanoparticles in vivo are not conclusive and several studies have generated conflicting

results [84].

1.3.1. Polymeric Micelles

Polymeric micelles are self-assembled core-shell structures composed of amphiphilic

block copolymers in aqueous solution at the concentration above their critical micelle

concentrations (CMC) [85]. Unlike those based on small molecule surfactants, the CMCs

of polymeric micelles, generally, are much lower, which indicates higher stability against

dissociation caused by dilution in a large volume of body fluid in vivo [85, 86].

Polymeric micelles as drug carriers have been employed to deliver many anticancer drugs,

22

such as doxorubicin, paclitaxel, camptothecin, which resulted in improved anticancer

efficacy and reduced adverse effects [87]. In particular, due to hydrophobic interactions

with the core of micelles, drugs with poor water solubility can be encapsulated in

micelles with high loading capacity; as a result, micellar encapsulation became one of the

most useful strategies to solubilize hydrophobic drugs and improve their bioavailability

[87]. By contrast, the micellation of hydrophilic compounds in the hydrophobic core has

been problematic due to the weak interactions between the drugs and lipophic segments

[86, 87]. Stability is another concern when designing a micellar formulation. Accelerated

dissociation of micelles may take place under high ionic strength conditions, such as in

blood, or in the presence of serum proteins, inducing pre-mature release of payloads

before they reach tumor sites. This limitation can be partly overcome through cross-

linking amongst unimers [87]. Notably, the extent of crosslinking not only influences the

stability, but can be also used to tune the rate and profile of drug release.

Several micellar formulations for platinum drugs are in various phases of clinic trials

and show promising anticancer efficacy [8]. Successful micellation of platinum

complexes with conventional amphiphilic block copolymers, such as PEG-b-PCL [88],

requires hydrophobic modification, which can be achieved by incorporating lipophilic

ligands as leaving groups in Pt(II) analogues or as the axial ligands in Pt(IV) prodrugs

[89]. Polymers with polyionic segments [90, 91], as alternative delivery vectors, can

spontaneously form polyionic complex micelles (PICs) with cationic platinum complexes

in aqueous solution as a result of electrostatic interactions and coordination bonds [92,

93]. Due to electrostatic neutralization, platinum complexes were found to localize to the

core of micelles [93, 94]. Moreover, the bifunctional cisplatin type platinum complexes

23

can act as cross-linkers in PICs, which may further stabilize the micelles in circulation

and lead to drug release in a controllable manner [93]. The examples and relevant

chemistry of platinum loaded micelles will be discussed in detail in sections 1.4.2 and

1.4.3.

1.3.2. Polymeric Based Prodrugs

The covalent attachment of anticancer agents to the backbones of polymers has been

demonstrated to be an effective strategy to improve their bioavailability [95, 96]. Several

promising polymer-drug conjugates are currently undergoing clinical studies [8]. The

tissue selectivity of the loaded drugs is dominated by the hydrophilic polymers, which

prolong circulation and are able to accumulate at tumor sites by taking advantage of the

EPR effect [97]. To further improve their anticancer efficacy, tumor specific moieties can

be incorporated into prodrugs, allowing enhanced tumor uptake [97]. The concept of

polymeric prodrugs has already been applied in the delivery of platinum drugs [98].

Platinum complexes can be conjugated to polymers through an extended linker, the

cleavage of which not only determines the stability of the conjugate in circulation, but

also allows selective drug release. In addition, platinum complexes can be attached to

polymers containing carboxylate and amino groups. The most successful polymeric

platinum prodrug so far is Prolindac [99-101], in which DACHPt, the [Pt (DACH)]2+

fragment of oxaliplatin, is attached to N-(2-Hydroxypropyl) methacrylamide (HPMA)

[96] through an amidomalonate chelating group (Figure 1.8) [100]. This conjugate

demonstrated ability to induce tumor regression in several xenograft models [101], and

was well tolerated in phase I clinic studies without significant toxicity [102]. Other

24

polymers exploited for this purpose include polyphosphazenes [103], cyclodextrins [104],

polyamino acids [105] and polysaccharides [106]. Additionally, dendrimers [107],

representing a unique class of dendritic polymers, can be used in a similar way [108-110].

Attractive features of dendrimers that make them promising drug delivery systems

include highly uniform structures, narrow size distribution and sufficient surface

functional groups available for drug tethering.

Figure 1.8. Examples of polymeric prodrugs for platinum drugs.

1.3.3. Liposomes

Liposomes are nano-sized, spherical and enclosed bilayer structures with superior

biocompatibility and low toxicity [84, 111]. They are mainly composed of phospholipids

and cholesterol and can encapsulate hydrophilic small molecules in their aqueous cores

and hydrophobic drugs within their phospholipid bilayers [84, 111]. Biomacromolecules,

25

such as therapeutic proteins [84] and nucleic acids [112], which are susceptible to

enzymatic degradation and poorly penetrate membranes, can also be effectively loaded

and delivered by liposomal formulations, leading to enhanced in vivo efficacy. In addition

to the therapeutics, magnetic resonance imaging (MRI) contrast agents can be

incorporated into liposomes as well for disease diagnosis and MRI-guided drug delivery

[113, 114]. Currently, several liposomal formulations have already been approved by the

US Food and Drug Administration (FDA) (Table 1.2), and many more are in various

phases of clinic trials. The first successful product is Doxil, the liposomal formulation of

doxorubicin used in the treatment of Kaposi's sarcoma, breast cancer, and ovarian cancer

with significantly reduced cardiotoxicity compared to the free drug.

Liposomal formulations of platinum-based agents, such as cisplatin and carboplatin,

have been extensively studied, with several promising candidates currently tested in

clinical trials (Table 1.3). One challenge in generating liposomal formulations of

platinum drugs is to achieve a high drug-to-lipid ratio [115]. Since most platinum drugs

are barely soluble in organic solvents, they cannot be loaded by the common procedure

used for paclitaxel [116], according to which the lipophilic drug is first dissolved along

with the lipids in an organic solvent such as chloroform. Platinum complexes are

normally encapsulated by passive loading [117, 118]. In this process the aqueous solution

containing platinum complex is used to hydrate a pre-formed lipid film to produce

platinum loaded liposomes. However, this method usually suffers from low loading

capacity as the result of limited water solubility (for example, the water solubility of

cisplatin is ~2 mg/ml) and poor membrane permeability of platinum complexes [119]. To

increase their water solubility, platinum complexes can be converted to aqua-species by

26

reacting them with silver nitrate, in which the chloride ligand is replaced by water [120].

This conversion also renders the cisplatin-type complexes cationic, which improves their

ability to interact with negatively charged lipids, such as DPPG, through electrostatic

interactions [115]. Liposomes developed according to this strategy showed an

unprecedented drug-to-lipid ratio of 2.5 and 1000-fold higher cytotoxicity in cell based

assays [115]. Lipophilic modification of platinum complexes is an alternative strategy to

improve drug loading. In this case, lipophilic ligands, including fatty acids [121],

phospholipids [122] and carboxylic acid-modified cholesterol [123], are incorporated

through ligand-exchange reactions, producing amphibious platinum complexes that can

self-assemble into phospholipid bilayers.

Leakage of the entrapped molecules from the liposomes during circulation would

produce undesired systemic toxicity. To avoid premature release in vivo, structural

modifications are made to the lipid bilayers. For example, incorporation of cholesterol

can reduce the fluidity of the membranes, leading to decreased leakage of the payloads

from the liposomes [124]. Alternatively, phospholipids with high phase transition

temperatures, which undergo gel-to-liquid-crystalline phase changes above physiological

temperature, also stabilize the bilayers [117]. Furthermore, drug retention is also affected

by the type of payload. Excellent drug retention properties are observed in Doxil, for

example, in which doxorubicin is encapsulated with a high loading capacity due to

loading technique that takes advantage of a trans-membrane pH/salt gradient [82]. The

exploitation of trapping agents, such as EDTA [125] and dextran sulfate [126], which

form inclusion complexes with the payload inside the liposomes, can improve the drug

retention as well.

27

Tab

le 1

.2. F

DA

Appro

ved

Lip

oso

mal

Fo

rmula

tions.

Bra

nd N

ame

Dru

g

Lip

id c

om

posi

tion (

mola

r

rati

o)

Indic

atio

ns

Ref

eren

ce

Am

bis

om

e A

mphote

rici

n B

H

SP

C,

DS

PG

, ch

ole

ster

ol

(2:0

.8:1

)

Fun

gal

infe

ctio

ns

[12

7]

Dau

noX

om

e D

aunoru

bic

in

DS

PC

and c

hole

ster

ol

(2:1

)

Blo

od t

um

ors

[1

28]

Dox

il

Dox

oru

bic

in

HS

PC

, ch

ole

ster

ol,

and

PE

G2

000-D

SP

E (

56:3

9:5

)

Kap

osi

’s s

arco

ma,

Ovar

ian/b

reas

t ca

nce

r

[82]

Myoce

t D

ox

oru

bic

in

EP

C a

nd c

hole

ster

ol

(55:4

5)

Com

bin

atio

n t

her

apy w

ith

cycl

ophosp

ham

ide

in m

etas

tati

c

bre

ast

cance

r

[129]

Dep

ocyt

Cyta

rabin

e

Chole

ster

ol,

Tri

ole

in, D

OP

C,

and D

PP

G (

11:1

:7:1

)

Neo

pla

stic

men

ingit

is a

nd

lym

phom

atous

men

ingit

is

[130]

Epax

al

Inac

tivat

ed h

epat

itis

A

vir

us

(str

ain R

G-S

B)

DO

PC

and D

OP

E

Hep

atit

is A

[1

31]

Infl

exal

V

Inac

tivat

ed h

emag

luti

nin

e

of

Infl

uen

za v

irus

stra

ins

A

and B

DO

PC

and D

OP

E

Infl

uen

za

[132]

28

Table 1.3. Liposomal Formulations for Platinum Drugs.

Formulation Drug Status Indications Reference

Lipoplatin Cisplatin Phase III NSCLC, breast cancer,

gastric cancer

[133]

SPI-077 Cisplatin Phase II Advanced NSCLC,

refractory ovarian cancer

[134]

Aroplatin

(L-NDDP)

aroplatin Phase II Refractory colorectal

cancer, malignant pleural

mesothelioma

[135]

Lipoxal Oxaliplatin Phase I Advanced gastrointestinal

cancer

[136]

MBP-426 Oxaliplatin Phase II Gastric, gastroesophageal,

esophageal

adenocarcinomas

[137]

Another factor that influences the therapeutic responses of liposomes is their release

properties at the disease sites. Formulations with poor release rates are unable to deliver

adequate amount of drug to the diseased tissues, which limits their bioavailability. This

has led to the discontinuation of clinical trials of SPI-077 (Table 1.3), a liposomal

formulation of cisplatin lacking in vivo antitumor efficacy due to insufficient drug release,

in spite of improved pharmacokinetics and reduced side effects [138]. This issue has been

resolved in lipoplatin, another liposomal cisplatin, which uses DPPG as one of the lipid

components. Because of the fusogenic DPPG, lipoplatin delivers platinum drug into

cancer cells through a membrane-fusion mechanism [133]. In addition, controlled release

from liposomes can be achieved when they are incorporated with various triggers in

response to environmental stimuli, such as heat, light, ultrasound, enzymes, and local pH

[111].

29

Short circulation times are frequently observed in conventional liposomes. This

drawback can be attributed to the rapid recognition of liposomes by the mononuclear

phagocyte system (MPS), such as hepatic Kupffer cells and splenic macrophages, which

leads to rapid clearance and reduced accumulation of drugs in targeted organs [83, 139].

The non-specific uptake by MPS is likely to be associated with opsonin, the serum

protein that can bind with liposomes, causing uptake into liver and spleen [139]. The

half-life of liposomes in circulation is to a large extent determined by the

physicochemical properties of lipid bilayers. For example, small liposomes with sizes

less than 100 nm have been demonstrated to evade scavange by MPS and show

prolonged circulation time [140]. Liposomes bearing negative charges interact more

readily with serum proteins than the cationic or neutral ones, showing rapid recognition

by MPS, and are normally more toxic [141]. Long circulation time can be achieved by

hydrophilic modification with polyethylene glycol (PEG). This leads to so-called “stealth”

liposomes, in which PEG provides a protective, hydrophilic shield that prevents the

interaction of liposomes with proteins in blood [83, 142]. The ability of long-circulating

liposomes to accumulate in tumors has been documented in thousands of publications,

and has been attributed to enhance permeability and retention (EPR) effect. Nevertheless,

it should be noted that rapid blood clearance may occur if the dose of PEGylated

liposomes is very low, which is caused by an undesired immunogenic response, known as

the “accelerated blood clearance (ABC) phenonomon” [142].

30

1.3.4. Nanotubes

Carbon nanotubes (CNTs) have been used as drug delivery systems in an attempt to

overcome the dose-limiting toxicity of platinum-based drugs [143, 144]. They are tubular

structures on the nanometer scale with high surface area, and relatively large interior

cavities. CNTs can be further categorized into single-walled nanotubes (SWNTs) and

multi-walled nanotubes (MWNTs) [144]. Although most nanotubes show little

cytotoxicity, their biocompatibility remains as major concern for in vivo applications

[145]. Ultra-short CNTs, 20–80 nm in length and 1.4 nm in diameter, seemed to be more

bio-compatible than the full-length CNTs [146, 147]. Both the exterior wall surface and

the interior space in CNTs can be used to load various drugs for improved efficacy [143].

However, to avoid cytosolic detoxification by intracellular nucleophiles, the interior

cavity is most likely a better location for platinum complexes, where they can be

encapsulated via capillary action [148]. This approach actually turns out to be

challenging for loading hydrophilic platinum complexes, such as cisplatin, which are

inherently incompatible with the hydrophobic cavities of CNTs. In this context, high

loading capacity would be achieved by selecting the corresponding more lipophilic Pt(IV)

prodrugs as alternatives [149, 150] (Figure 1.9, 1-38). To prevent rapid drug leakage of

nanotubes from occurring in circulation, the openings of CNTs can be blocked by

molecular caps; this also allows the release of platinum complexes in a controlled way at

disease sites [151]. In addition, platinum complexes can be tethered to CNTs through

covalent bonds, which requires surface modification of CNTs with carboxylic acid [151]

(Figure 1.9, 1-39), in which case the release of platinum complexes relies on ligand

exchange. This will be discussed in more detail in section 1.4.2. Apart from their

31

extraordinary capacity to retain payloads, the transmembrane penetration of CNTs was

found to occur via a “nanoneedle” mechanism as a complementary pathway to

endocytosis [152], which may help minimize undesired entrapment of platinum drugs in

the lysosomes. Incorporation of PEG [153] (Figure 1.9, 1-40) as a surface modification

not only improves the biocompatibility of CNTs and prolong their circulation times, by

evading the recognition of MPS, but also may improve their dispersity in biological

media. CNTs containing targeting moieties [154, 155], such as folic acid (Figure 1.9, 1-

41), have been demonstrated to preferentially accumulate at tumor sites, and result in

significant tumor regression in mice xenograft compared with controls.

Figure 1.9. Examples of nanotube as delivery vectors for platinum drugs.

32

1.4. Target-Selective Activation of Platinum-Based Anticancer Agents

A successful drug delivery system requires structural integrity of the formulation in

circulation and selective release of cytotoxic payload in tumor tissue. Several validated

strategies for the targeted release of non-metal-based anticancer agents can also be used

for delivering platinum complexes. These involve installation of pH-sensitive [156, 157],

hypoxia-labile [158, 159], and enzymatically cleavable linkers [160, 161]. In platinum-

containing targeted therapies, additional metal-centered reactivities exist such as ligand

exchange, redox reactions, and electrophilic/Lewis-acidic bond activation, which can be

exploited for drug delivery applications.

1.4.1. Controlling solution reactivity

The reactivity of platinum-based agents in biological matrices depends on the nature

and stereoisomerism of the ligands in the metal coordination sphere and on the relative

nucleophilicity of competing bioligands. As mentioned above, the coordinative Pt–N

bonds between cisplatin and DNA require intracellular aquation. By contrast, undesired

reaction of cisplatin with thiol-containing molecules, such as GSH, which causes off-

target effects and tumor resistance, proceeds without aquation [162, 163]. Relative

ligand affinities can generally be predicted from the hard and soft acids and bases

principle (HSAB) [164]. The bonds formed between Pt(II) and cysteine thiol can be

considered resistant to cleavage by any intracellular nucleophile on a biologically

relevant time scale [162, 165]. On the other hand, coordination to methionine thioether

was found to be slowly reversed by DNA nitrogen [166-169]. Formation of reversible

33

Pt(II)–thioether bonds with methionine-rich proteins, such as copper transporter protein

(CTR1), has been implicated in the cellular trafficking of platinum complexes [170, 171].

Inspired by these findings, thioethers have been proposed as alternative donor groups in

carrier ligands that allow controlled release and transfer of Pt(II) to DNA [170]. Finally,

the stereochemistry of substitution reactions in square-planar platinum complexes and

susceptibility of specific ligands to substitution are modulated by the trans-effect [172].

Sulfur ligands, for example, show a relatively strong trans-effect that may promote

(undesired) substitution of the ammine nonleaving groups in cisplatin [172]. Conversely,

even weakly nucleophilic chloride ion has been shown to substitute sulfur donors, if the

latter has been installed trans to phosphine ligands [173], which are highest in the trans-

effect series.

Antitumor active platinum compounds contain leaving groups of intermediate lability,

such as chloride or dicarboxylates, to assure the (pro)drug remains intact during

circulation but sufficiently reactive in target tissue to produce its biological effect.

Aquation of cisplatin occurs in a chloride-ion dependent manner, which is largely

suppressed in the blood [3], but favored in the cytosol where the concentration of

chloride is 30–50-fold lower. The aquation products of cisplatin rapidly react with DNA

nitrogen and can be considered the therapeutically active form of the drug. Aquation also

enhances the water solubility of cisplatin [174] and confers positive charge to the

complex [115, 164], which allows encapsulation of the activated form of the drug in the

hydrophilic core of nanoliposomes [115, 174] and polymeric nanoparticles [8, 175].

Replacement of the chlorido ligands in cisplatin with CBDCA (cyclobutane-1,1-

dicarboxylate), which produces a relatively inert six-membered O,O-chelate, affords

34

carboplatin, a second-generation cisplatin derivative with improved toxicity profile. Loss

of the leaving group in aqueous media in carboplatin is significantly slower than in

cisplatin (half-lives at 37 ºC and pH 7 of 450 h and 2 h, respectively [176, 177]). Unlike

chlorido and carboxylato ligands, N-donor ligands become leaving groups only in the

presence of trans-activating sulfur donors. While this reactivity is generally undesired

because it causes deactivation of platinum drug, it has been exploited in platinum-

containing linkers used to conjugate kinase inhibitors to a tumor-specific carrier for

improved selectivity. In this case, intracellular thiols allow controlled release of active

inhibitor by cleavage of a Pt–N bond [178].

1.4.2. Activation by ligand exchange

Various platforms have been developed for the targeted delivery of a cytotoxic

platinum payload in which platinum has been attached to carboxylic acid-modified

polymeric materials. Polyaspartic acid and polyglutamic acid contain cis-

diam(m)ineplatinum(II) moieties coordinated to monodentate carboxylate or chelated in a

bidentate fashion (Figure 1.10). Platinum in these formulations decreases the negative

charge and increases the hydrophobicity of the polymers. Because of their amphiphilic

character, the platinum-modified polymers self-assemble into micelles, in which the most

hydrophobic segments of platinum-modified polymer form the core and unmodified

polycarboxylic acid the outer shell of the particle. Bifunctional platinum produces cross-

links between polymer chains, which increases the stability of the micellar structure in

circulation [80, 179-181]. The release of active platinum species from these formulations

is triggered by simple ligand exchange with ubiquitous nucleophiles such as chloride and

35

water [8, 80, 181]. Acidic conditions may accelerate the release of the platinum payload

due to competitive protonation of carboxylate at low pH (discussed in detail in the

following section). Carboxylic acid-modified peptide residues were also introduced as

side chains in polymers to allow modification with platinum complexes. Terminal Ɛ-

carboxylate groups installed adjacent to an amide nitrogen form a stable five-membered

N,O-chelate with cis-diammineplatinum(II) [182].

The release kinetics of platinum complexes from polymer can be tuned by

incorporation of multiple donors with varying affinity for platinum. For instance,

biopolymers “doped” with tellurium release platinum complex triggered by competitive

ligand displacement on a time scale of several days up to a month [183]. Carriers

containing N-donors have been used as macromolecular non-leaving groups to generate

substitution-inert polymer–platinum conjugates [184]. In this case, the inactive polymer

would require degradation to produce freely diffusible, monomeric fragments able to

form cytotoxic DNA adducts similar to cisplatin. However, a mechanism by which intact

polymer enters the nucleus to damage chromatin has also been proposed [185]. As

alternative carriers to polymers, other systems such as dendrimers [186], mesoporous

silica nanoparticles (MSN) [187, 188], carbon nantotubes [146], and gold nanoparticles

[189, 190], as well as hybrid materials, have also been used to deliver platinum

complexes (see also section 1.3). Selective representative systems are summarized in

Table 1.4.

36

Fig

ure

1.1

0.

Act

ivat

ion o

f poly

mer

–pla

tinum

conju

gat

es b

y l

igan

d e

xch

ange.

Note

the

dis

tinct

ion b

etw

een

pla

tinum

-coord

inat

ing c

arbox

yla

te i

n t

he

inte

rior

and s

olu

bil

izin

g c

arbox

yla

te o

n t

he

surf

ace

of

the

par

ticl

e.

37

Tab

le 1

.4.

Pla

tinum

-Bas

ed F

orm

ula

tions

Act

ivat

ed b

y L

igan

d E

xch

ange

Dru

g

Pla

tfo

rm

Co

mponen

ts

Pre

clin

ical

and c

linic

al s

tatu

s R

efer

ence

s

Cis

pla

tin

Mic

ella

r

form

ula

tio

n

(NC

-60

04

)

Po

ly(e

thyle

neg

lyco

l)-b

-

po

ly(g

luta

mic

aci

d)

NC

-6004 s

how

s sa

me

pote

ncy

as

cisp

lati

n i

n m

ou

se x

eno

gra

fts

and r

educe

s oto

toxic

ity a

nd n

ephro

toxic

ity i

n m

ice

and g

uin

ea

pig

; dow

nre

gula

tes

gen

es r

elat

ed t

o t

um

or

invas

ion a

nd

met

asta

sis;

is

curr

entl

y b

eing e

val

uat

ed i

n p

has

e II

cli

nic

al t

rial

s

in t

he

US

agai

nst

pan

crea

tic

cance

r an

d N

SC

LC

.

[19

1-1

93

]

Oxal

ipla

tin

Mic

ella

r

form

ula

tio

n

(NC

-40

16

)

Po

ly(e

thyle

neg

lyco

l)-b

-

po

ly(g

luta

mic

aci

d)

NC

-4016 s

how

s im

pro

ved

eff

icac

y i

n x

eno

gra

ft m

ou

se m

od

els

com

par

ed t

o o

xal

ipla

tin, unli

ke

oxal

ipla

tin

, d

oes

not

ind

uce

acu

te

cold

hyper

sensi

tivit

y i

n r

ats;

sel

ecti

vel

y r

elea

ses

pay

load

wit

hin

late

en

do

som

es n

ear

the

nucl

eus;

fo

rmula

tion i

s cu

rren

tly b

eing

eval

uat

ed i

n p

has

e I

clin

ical

tri

als.

[19

4-1

96

]

Oxal

ipla

tin

Po

lym

eric

pro

dru

g

Po

lyp

hosp

haz

ene

The

pro

dru

g s

how

s hig

h c

yto

toxic

ity i

n a

wid

e ra

nge

of

hu

man

cance

r ce

ll l

ines

and im

pro

ved

eff

icac

y a

nd

lo

wer

tox

icit

y t

han

oxal

ipla

tin i

n x

enogra

ft m

odel

s; i

s bio

deg

radab

le a

nd

accu

mula

tes

in t

um

or

tiss

ue.

[19

7,

198

]

Cis

pla

tin

Cro

ss-l

inked

po

lym

er

ves

icle

s

P(O

NB

An) m

-b-P

(ON

B-

PE

G2

k) n

; te

trak

is(3

-

mer

capto

pro

pio

nat

e)a

The

self

-ass

emble

d v

esic

le a

llow

s co

ntr

oll

ed r

elea

se o

f ci

spla

tin

over

day

s in

vit

ro.

[19

9]

Oxal

ipla

tin

Po

lym

eric

met

allo

som

es

Y-s

hap

ed c

opoly

mer

, -

cho

lest

eroyl-

poly

(L-

glu

tam

ic a

cid)

and

po

ly(e

thyle

neg

lyco

l)

(PE

Gas

us-

PL

GA

-

Ch

ole

)

The

form

ula

tion s

how

s pro

longed

blo

od

cir

cula

tio

n, se

lect

ive

tum

or

accu

mula

tion, an

d i

mpro

ved

eff

icac

y i

n v

itro

an

d i

n v

ivo

.

[20

0]

Cis

pla

tin

Den

dri

mer

P

oly

(am

idoam

ine)

(PA

MA

M)

den

dri

mer

s

Den

dri

mer

ic f

orm

ula

tion s

how

s re

duce

d t

oxic

ity a

nd

co

mp

arab

le

effi

cacy

to c

ispla

tin i

n A

2780 x

enogra

ft m

od

el.

[18

6]

Cis

pla

tin

MS

N

Sil

ica

Nan

opar

ticl

es p

rom

ote

upta

ke

of

cisp

lati

n i

nto

HeL

a ce

lls;

sho

w

hig

her

anti

tum

or

acti

vit

y t

han

fre

e ci

spla

tin

agai

nst

MC

F-7

an

d

HeL

a ce

lls.

[18

8]

aO

NB

An =

ox

anorb

orn

enyl

anh

yd

ride;

ON

B-P

EG

2K =

ox

anorb

orn

enyl

po

ly(e

thyle

neg

lyco

l).

38

1.4.3. Activation by pH-sensitive release

The extracellular matrix of tumor tissue is slightly more acidic than that found in

normal tissue. This has been exploited for the design of several pH-sensitive drug

delivery systems, in which the low pH (6.0–6.5) [201] triggers rapid drug release and

uptake into target cells [202]. Drugs with pH-dependent cytotoxicity profiles reduce

severe toxicities in normal tissues, which may lead to an improved therapeutic index.

Thioplatins [203], a group of platinum xanthate derivatives, are significantly more

cytotoxic if treated cancer cell cultures are maintained at pH 6.8 instead of pH 7.4. They

show potent antitumor activity similar to cisplatin in a xenograft model, but are better

tolerated than the clinical drug by the test animals. In structure–activity relationship

(SAR) studies a correlation was observed between biological activity and the basicity

(based on pKa values of the conjugate acids) of the pH-responsive leaving group.

Likewise, aminoalcohol [201] and 1,3-dihydroxyacetone oxime [204] have been

introduced with similar results (Figure 1.11). Both leaving groups form five-membered

N,O-chelates, which are inert at neutral physiological pH but become reactive under

slightly acidic conditions.

Figure 1.11. Representative examples of pH-sensitive platinum complexes.

39

Targeting acidic intracellular organelles, such as endosome (pH ≈ 5.5) and lysosome

(pH ≈ 5.0) [156], is another strategy intensively investigated for targeted delivery of

anticancer agents. Unlike the acidic environment in the tumor extracellular matrix, the

low pH in these organelles accelerates hydrolysis of acid-labile functional groups,

including hydrazones, acetals, and amides, which are frequently used to conjugate

organic drugs to carrier molecules and materials [205]. Similar to platinum payloads

targeted at the low-pH extracellular matrix, acid-labile platinum–carboxylate

coordination has also been exploited to trigger release of platinum drugs by an

intracellular pH differential. Micellar vehicles [8] composed of polymers bearing

carboxylic acids, such as polypeptides, poly(methacrylic acid), and acid-modified

polysaccharides have been developed. These micelles are stabilized in circulation by

platinum-mediated cross-links between polymer strands, but disintegrate as they

accumulate in acidic intracellular endosomal/lysosomal vesicles [159]. Five-membered

N,O-chelates containing carboxylate show enhanced stability at neutral physiological pH

compared to carboxylate coordination alone. This observation has led to the

development of ProlindacTM, a poly(N-(2-hydroxypropyl)methacrylamide) (pHPMA)

based polymer–platinum conjugate, which is currently undergoing phase II clinical

trials[206].

A major problem with the delivery of anticancer agents whose major biological targets

are located in the cytosol or nucleus is their sequestration into, and degradation by, acidic

organelles [207]. Incorporation of agents capable of disrupting the endosomal membrane,

like pore-forming proteins and peptides, pH-buffering molecules, or fusogenic molecules,

into drug delivery systems may overcome this obstacle. An example of this concept is

40

the use of pH-sensitive liposomes containing fusogenic lipid

dioleoylphosphatidylethanolamine (DOPE), which promotes endosomal escape [208].

The cisplatin-based liposomal formulation showed increased activity in a human lung

cancer cell line compared with free cisplatin and was able to overcome resistance to the

clinical drug. On the other hand, a recent study [209] suggests that the release of

platinum in acidic subcellular compartments may lead to improved potency. When

oxaliplatin-based (1R,2R-diaminocyclohexane) platinum (II) was delivered with

poly(ethyleneglycol)-b-poly(glutamic acid)-based polymer, improved antitumor activity

was observed for the micellar formulation. This effect was ascribed to selective drug

release in the lysosomes, which allows the cytotoxic payload to bypass cytoplasmic

detoxification and accumulate in the perinuclear region. Three representative

formulations activated in a pH-dependent manner intracellularly are summarized in Table

1.5.

Table 1.5. Platinum-Based Formulations Activated by Lysosomal Degradation

Platform Components Preclinical and clinical

observations

References

Nanoparticle Succinic acid-

decorated dextran,

cisplatin,

doxorubicin

Prolonged circulation, decreased

accumulation in normal tissues,

enhanced therapeutic efficacy in

tumor-bearing mice

[210]

Dendrimer mPEGylated

peptide dendrimer

w/ peripheral

carboxyl groups,

oxaliplatin

Formulation suppresses tumor

growth more efficiently than

oxaliplatin without inducing

toxicity in a SKOV-3 human

ovarian tumor xenograft

[174]

Polymeric

prodrug

(ProLindac™)

Poly(hydroxyprop

ylmethacrylamide)

w/triglycine side

chains, oxaliplatin

Safety and efficacy have been

demonstrated in phase I/II clinical

trials, superior efficacy of

ProLindac in several tumor

models, acts synergistically with

other cytotoxic agents in vitro

[100, 206]

41

1.4.4. Photochemical activation

The concept of spatially selective activation of anticancer agents at the tumor site with

light, known as photodynamic therapy, has also been applied to photochemically active

metal complexes. Irradiation of transition metal complexes with ultraviolet or visible

light generates excited electronic states, which lead to altered chemical reactivity [211].

As a consequence, specific metal–ligand bonds are weakened, giving rise to accelerated

bond dissociation and ligand exchange. Octahedral Pt(IV) complexes undergo various

photochemical transformations, including ligand substitution, isomerization, and

reductive elimination to the corresponding square-planar Pt(II) complexes [211]. This

provides a means of caging the reactive Pt(II)-based drugs in their inert and nontoxic

Pt(IV) form and activating them as prodrugs selectively in tumors, which reduces damage

to normal tissue and systemic toxicity [212, 213]. A major advantage of light-activated

Pt(IV) complexes over traditional, singlet oxygen-producing photodynamic therapy (PDT)

is that the former mechanism is oxygen independent [214], which has advantages in

treating hypoxic tumors. Photochemical activation of Pt(IV) complexes typically

requires short-wavelength (UVA) radiation, which does not penetrate tissue deeply

enough for treatment of non-topical lesions [215]. (Two-photon excitation has been

proposed as a technique to overcome this drawback [216], see Table 1.6.) Representative

Pt(IV) complexes explored for applications in PDT are presented in Figure 1.12.

One of the first light-sensitive Pt(IV) complexes studied was trans,cis-[PtCl2I2(en)] (1-

20), which readily loses its photolabile iodido ligands. The photoreductive loss of the

iodido ligands generates Pt(II) complexes, which are able to form DNA cross-links [214,

217, 218]. Unfortunately, complex 1-42 suffers from unexpected instability under cell

42

culture conditions even in the absence of light, presumably due to intracellular reduction

by glutathione [214, 217, 218]. As a consequence, light had no effect on the inhibition of

cancer cell proliferation by this complex [217]. Structural modifications were made to

reduce the dark reactivity of the prototypical Pt(IV) complex in the presence of

intracellular reducing agents. This was achieved by introducing axial ligands (Y, see

Figure 1.12) that allow tuning the lipophilicity and redox potential of the metal. For

example, cis,trans-[PtI2(OAc)2(en)] (1-43) [217], which contains axial acetato instead of

chlorido ligands, behaves inert when incubated with calf-thymus DNA for 6 h in the dark,

but readily forms DNA adducts in the presence of light. By contrast, use of hydroxide

ligands in cis,trans-[PtI2(OH)2(en)] (1-44) gave a complex that was relatively unreactive

with DNA in the presence of light [219].

Regardless of the nature of axial ligands none of the iodide complexes showed

sufficient dark stability in the presence of biological reducing agents. To address this

problem, the iodido ligands were replaced with photolabile azido ligands. The Pt(IV)–

azido complexes undergo two-electron reductions to generate Pt(II) species via release of

azide radicals, which further decompose into N2 [220]. Pt(IV)–diazidodiam(m)ine

complexes 1-45 and 1-46 are stable in serum for several weeks in the dark, and only their

photoreduction products react with DNA. The Pt(IV)–azido complexes overcome

cisplatin resistance in 5637 human bladder cancer cells [221].

The complex trans,trans,trans-[Pt(N3)2(OH)2(NH3)2] (1-47), a geometric isomer of

compound 1-45, was as cytotoxic as cisplatin in a number of cell lines [222, 223].

Replacement of one NH3 ligand with pyridine (py) to produce trans,trans,trans-

[Pt(N3)2(OH)2(NH3)(py)] (1-48) resulted in markedly enhanced phototoxicity [224]. The

43

compound showed an advantage over cisplatin in cisplatin-sensitive and cisplatin-

resistant A2780 ovarian cancer cells of 80-fold and 15-fold, respectively [225]. NH3 in

1-48 undergoes undesired photodissociation leading to the species with reduced

cytotoxicity [226, 227]. This reactivity is not observed in trans,trans,trans-

[Pt(N3)2(OH)2(py)2] (1-49), which rapidly induces DNA cross-links and shows greatly

enhanced cytotoxicity [228-230].

Figure 1.12. Examples of photochemically active Pt(IV) complexes.\

Finally, to further improve the pharmacokinetic properties of the Pt(IV)–azido

complexes, the prodrugs were attached to an amphiphilic block copolymer through amide

linkages formed between the amine groups of the polymer and the dangling carboxylate

groups of succinato ligands installed in the axial positions of Pt(IV) [231]. The

polymeric prodrug self-assembles into micellar nanoparticles in aqueous solution, which

are stable in the dark but spontaneously release active Pt(II) species photoreductively

(Figure 1.13).

44

Fig

ure

1.1

3.

Sel

f-as

sem

bly

of

poly

mer

ic p

rodru

g t

hat

rel

ease

s ac

tive

Pt(

II)

spec

ies

trig

ger

ed b

y U

VA

irr

adia

tion.

45

Tab

le 1

.6. E

xam

ple

s an

d C

har

acte

rist

ics

of

Photo

acti

vat

able

Pt(

IV)

Pro

dru

gs

Pla

tinum

der

ivat

ive

Pla

tform

P

recl

inic

al a

nd c

linic

al o

bse

rvat

ions

Ref

eren

ces

1-4

5

Pro

dru

g

Form

s ci

spla

tin-t

ype

guan

ine–

gu

anin

e cr

oss

-lin

ks

wit

h D

NA

up

on

irra

dia

tion;

is a

s cyto

tox

ic a

s ci

spla

tin i

n a

cel

l m

odel

when

photo

acti

vat

ed.

[23

2, 233]

1-4

6,1

-47

P

rod

rug

T

he

com

ple

xes

wer

e m

ore

photo

cyto

tox

ic t

han

the

corr

espondin

g

cis-

isom

ers.

[230, 234, 235]

1-4

8

Pro

dru

g

Form

s in

tras

tran

d D

NA

cro

ss-l

inks;

is

up

to 8

0-f

old

more

cyto

tox

ic

than

cis

pla

tin;

is a

ctiv

e in

cis

pla

tin

-res

ista

nt

hum

an o

var

ian c

ance

r

cell

s; a

nti

tum

or

acti

vit

y i

n x

enogra

ft i

s en

han

ced b

y v

isib

le l

ight.

[222, 225]

1-4

9

Pro

dru

g

Com

ple

x c

an b

e ac

tivat

ed o

ver

a w

ide

wav

elen

gth

ran

ge,

incl

udin

g

vis

ible

lig

ht;

pro

duce

s ce

ll k

ill

pote

nti

ally

thro

ugh

pla

tinat

ion o

f

DN

A a

nd/o

r pro

tein

s.

[236]

trans,

tran

s,tr

ans-

[Pt(

N3) 2

(OH

) 2(N

H2M

e)(p

y)]

Pro

dru

g

Ox

idiz

es g

uan

ine

wh

en i

rrad

iate

d d

ue

to f

orm

atio

n o

f si

ngle

t

ox

ygen

.

[237]

cis-

[PtC

l 2(M

OP

EP

a) 2

]

Pro

dru

g

The

firs

t P

t(IV

) co

mple

x t

hat

can

be

acti

vat

ed b

y t

wo

-photo

n

abso

rpti

on;

ligan

d e

xch

ange

can b

e tr

igger

ed b

y f

emto

seco

nd

las

er

irra

dia

tion a

t 600–740 n

m.

[216]

Pt(

IV)–

azid

e co

mple

xes

b;

Mic

elle

Ir

radia

tion o

f m

icel

lar

form

ula

tion r

apid

ly l

iber

ates

act

ive

Pt(

II)

spec

ies

that

pro

duce

DN

A c

ross

-lin

ks

in v

itro

; in

viv

o s

tud

y s

how

ed

impro

ved

eff

icac

y a

nd r

educe

d s

yst

emic

tox

icit

y.

[231]

aM

OP

EP

= 4

-[2-(

4-m

ethox

yphen

yl)

- et

hyn

yl]

pyri

din

e. b

Co

-form

ula

ted w

ith m

ethox

yl-

poly

(eth

yle

ne

gly

col)

-blo

ck-p

oly

(ε-

capro

lact

one)

-poly

(L-l

ysi

ne)

46

1.4.5. Activation by bioreduction

Reactions of square-planar Pt(II) anticancer agents with biological nucleophiles other

than DNA are likely responsible for their severe systemic toxicity. To overcome this

drawback, Pt(IV) complexes have been proposed early on as non-toxic prodrugs that can

be activated by reduction to the corresponding Pt(II) species [238]. Because octahedral

Pt(IV) complexes are coordinatively saturated and kinetically inert, ligand exchange

reactions with bionucleophiles in the blood can be suppressed to minimize off-target

toxicity. As another potential benefit, Pt(IV) complexes show pharmacokinetic and

toxicity profiles favorable for oral delivery [239, 240].

Satraplatin [218] (cis,trans,cis-[PtCl2(OAc)2(NH3)(cyclohexylamine)], also known as

JM216), is the paradigm of a clinically evaluated oral platinum complex. Satraplatin

loses its axial acetato ligands in cells by reductive elimination. Modification of the axial

ligands with lipophilic moieties is an important means of tuning the

lipophilicity/hydrophilicity balance of the metal complexes [241], which affords

molecules with both favorable solubility and membrane permeability[242]. Lipophilic

groups are also beneficial for nanoencapsulation of the complexes (Figure 1.14A) [89,

243].

The rate of reduction of Pt(IV) prodrugs mainly depends on the nature of the axial

ligands and correlates well with Pt(IV)/Pt(II) reduction potentials [242, 244, 245].

Carboxylato and hydroxido ligands typically give complexes with lower reduction

potentials and slower reduction rates than chlorido ligands [246, 247]. Strongly electron-

withdrawing ligands may render the Pt(IV) complexes too reactive. Mitaplatin [248], for

example, a Pt(IV) compound containing dichloroacetate (DCA) as axial ligands, which

47

was designed to target both nuclear DNA and the mitochondria, undergoes rapid ester

hydrolysis (t1/2 = 2 h) when incubated in biological relevant media in the absence of

reducing agents [249]. This observation suggests that the complex may not enter cancer

cells intact, which would affect the synergistic action of its components.

Importantly, the axial ligands can be turned into chemical handles for conjugating the

inert Pt(IV) prodrugs to macromolecules [250, 251] and nano-sized particles [150, 154,

252-254] (Figure 1.14B). Other bioactive molecules [255, 256], such as valproic acid [89,

257, 258], estrogen [258] and peptides [259-261], have also been introduced as axial

ligands, yielding multifunctional Pt(IV) complexes that damage cancer cells by

synergistic mechanisms or selectively target tumor tissue. Table 1.7 summarizes

technologies developed for the delivery of Pt(IV) prodrugs and their incorporation into

multifunctional therapies.

Figure 1.14. (A) Encapsulation of lipophilic Pt(IV) prodrug within nanoparticles; (B)

covalent attachment of Pt(IV) prodrug to nanoparticles.

48

Tab

le 1

.7.

Tar

get

ed a

nd M

ult

ifunct

ion

al C

ispla

tin-D

eriv

ed P

t(IV

) P

rodru

gs

Act

ivat

ed b

y B

iore

duct

ion.

Pla

tform

C

om

ponen

ts

Pre

clin

ical

and c

linic

al o

bse

rvat

ions

Ref

eren

ces

Tar

get

ed c

onju

gat

e

Mit

och

ondri

a-pen

etra

tin

g

pep

tide,

Pt(

IV)

com

ple

x

The

conju

gat

e is

equ

ally

eff

ecti

ve

in w

ild

-typ

e an

d

cisp

lati

n-r

esis

tant

ovar

ian c

ance

r ce

lls;

cyto

tox

icit

y

corr

elat

es w

ith d

amag

e to

mit

och

ondri

al D

NA

(m

tDN

A).

[262]

Dual

-funct

ion

conju

gat

e

α-T

oco

ph

erol

succ

inat

e,

Pt(

IV)

com

ple

x

The

conju

gat

e ca

use

s both

dam

age

to n

ucl

ear

DN

A a

nd

mit

och

ondri

al d

ysf

un

ctio

n;

the

conju

gat

e is

more

cyto

tox

ic

than

cis

pla

tin.

[263]

Nan

opar

ticl

e P

LG

A-b

-PE

G,

Pt(

IV)

pro

dru

g,

pro

stat

e-sp

ecif

ic m

embra

ne

anti

gen

Pro

longed

syst

emic

blo

od c

ircu

lati

on a

nd d

ecre

ased

accu

mula

tion o

f P

t in

the

kid

ney

s le

ad t

o a

rem

ark

able

impro

vem

ent

in t

her

apeu

tic

index

.

[264]

Nan

otu

be

Sin

gle

-wal

led

car

bon

nan

otu

be,

Pt(

IV)

pro

dru

g,

foli

c ac

id

The

nan

otu

bes

incr

ease

cel

lula

r upta

ke

of

pla

tinum

and

show

enhan

ced c

yto

tox

icit

y.

[150, 154]

Gold

nan

opar

ticl

e A

min

e-fu

nct

ional

ized

,

poly

val

ent

oli

gonucl

eoti

de

gold

nan

op

arti

cles

,

Pt(

IV)

pro

dru

g

The

cyto

tox

icit

y o

f th

e P

t(IV

) co

mp

lex

incr

ease

s

signif

ican

tly i

n s

ever

al c

ance

r ce

ll l

ines

.

[252]

Dual

-funct

ion

conju

gat

e

Dic

hlo

roac

etat

e,

Pt(

IV)

pro

dru

g

The

conju

gat

e kil

ls c

ance

r ce

lls

by d

amag

ing n

ucl

ear

DN

A

and d

isru

pti

ng m

itoch

on

dri

al f

unct

ion.

[265]

Du

al-f

un

ctio

n,

targ

eted

conju

gat

e

Pt(

IV)

pro

dru

g,

FP

R1/2

-

targ

eted

pep

tide

Sel

ecti

vel

y k

ills

the

cance

r ce

lls

and

in

du

ces

imm

un

e

resp

onse

in c

ell

model

s.

[26

6]

49

Tab

le 1

.7 (

conti

nued

)

Mic

elle

P

OE

GM

EM

A3

5-b

-PM

AA

200

and P

OE

GM

EM

A2

6-b

-

PM

AA

90, P

t(IV

) p

rodru

g,

foli

c ac

id

Mic

elle

s en

ter

cance

r ce

lls

by f

ola

te-r

ecep

tor-

med

iate

d e

ndocyto

sis.

The

cyto

tox

icit

y i

n c

ance

r

cell

lin

es i

s in

feri

or

to t

he

dru

g a

lone.

[267]

Mic

elle

P

t(IV

) co

mple

xes

as

bu

ild

ing

blo

cks

for

coord

inat

ion

poly

mer

s

Red

uct

ion o

f P

t(IV

) le

ads

to d

egra

dat

ion o

f

poly

mer

; fo

rmula

tion s

how

s pro

longed

blo

od

circ

ula

tion a

nd i

mpro

ved

tum

or

accu

mula

tion i

n

mic

e.

[250]

Mic

elle

M

PE

G5

000-b

-P(L

A1

400-c

o-

MC

C100

0),

pac

lita

xel

,

pla

tinu

m(I

V)

pro

dru

g

Form

ula

tion s

how

s sy

ner

gis

tic

acti

on i

n v

itro

and

reduce

d s

yst

emic

tox

icit

y a

nd e

nhan

ced a

nti

tum

or

effi

cacy i

n v

ivo.

[268]

50

1.5. Goals of Doctoral Research

While platinum–acridine agents outperform cisplatin in cultured cells, they suffer from

poor pharmacokinetic properties, which result in limited efficacy and dose-dependent

systemic toxicity. The overall goal of this research is to develop chemistries that will

enable new technologies to accelerate the preclinical development of platinum-acridines

with improved drug-like properties. Toward this end, we set out to establish the

methodologies that allow the chemical conjugation of platinum-acridine derivatives to

other bioactive molecules, such as tumor-targeted ligands, so that the consequent

conjugates are rendered with tumor selectivity, enhanced efficacy and reduced side

effects. To achieve this goal, conjugatable platinum-acridines with excellent cytotoxicity

should be firstly identified (Figure 1.15).

Figure 1.15. The schematic illustration of platinum-acridine based conjugates.

In chapter 2, we have designed a high-throughput library approach to search for potent

functionalized platinum-acridines by taking advantage of a highly efficient platinum-

mediated amine-nitrile addition reaction. This approach enables us to rapidly assemble

51

platinum-acridines in a fragment-based manner and test them in cytotoxic assays. In

addition, the structure-activity relationship established in this study suggests that the

optimal sites for further chemical conjugations are the side chains in nitrile moiety (R1,

Figure 1.15) and acridine fragment (R2, Figure 1.15), the modications of which cause

minimal interference as platinum-acridines bind to DNA.

Next, we explored the conjugation chemistries for the attachment of functionalized

platinum-acridines to other molecules. Two synthetic strategies have been developed in

this dissertation (Figure 1.16). In chapter 4, the estrogen receptor (ER) inhibitor

endoxifen was tethered to hydroxyl-modified platinum-acridine to sensitize ER positive

breast cancer to platinum drugs. The synthesis of the conjugates were achieved via the

“pre-assembly” method, in which the platinum moiety is introduced after all other

functional fragments were assembled into organic ligands. However, synthesis of

platinum complexes according to this method appears to be highly customized, which

prompted us to develop “post-modification” method for platinum-acridines. In this

approach, the conjugatable platinum-acridine analogues were ligated to other components

through “platinum-compatible” conjugation reactions. The copper-mediated click

chemistry was used in chapter 3 to map the subcellular distribution of an azide-modified

platinum-acridine derivative with fluorescent molecular probe containing an alkyne

handle, which allows us to gain more mechnastic information about platinum-acridines.

In addition, we have developed an efficient coupling reaction in chapter 6 to generate an

amide bond between carboxylic acid modified platinum-acridines and other bioactive

ligands containing amino group. The modularity of this approach enables the parallel

synthesis of a library of multifunctional platinum-acridine conjugates for drug screening.

52

For platinum-acridine conjugates, we reasoned that the controlled release of cytotoxic

warheads in tumor cancer is also important to their efficacy. A class of ester-based

prodrugs of platinum-acridine were produced in chapter 5. The installation of ester group

is supposed to improve the lipophilicity of the hydrophilic platinum complexes, and thus

presenting improved drug-like properties. These prodrugs were found to be activated

through two pathways: (1) metal-assisted, chloride-triggered ester cleavage and (2)

hCES-2 mediated enzymatic cleavage.

Different from the approaches relying on platinum-acridine conjugates, another strategy

in this dissertation is to deliver platinum-acridines using nanoliposomes, which have been

exploited to improve the therapeutical indices of many other conventional

chemotherapeutic agents. We optimized the procedures to encapsulate the most cytotoxic

derivative and expected this delivery vehicle for platinum-acridine analogues to able to

overcome the drawbacks they currently experienced.

Figure 1.16. (A) Construction of platinum-acridine conjugates with “pre-assembly”

method. (B) Construction of platinum-acridine conjugates with “post-modification”

method.

53

CHAPTER 2

USING A BUILDING-AND-CLICK APPROACH FOR PRODUCING

STRUCTURAL AND FUNCTIONAL DIVERSITY IN DNA-TARGETED

HYBRID ANTICANCER AGENTS

The work contained in this chapter was initially published in Journal of Medicinal

Chemistry in 2012. The manuscript, including figures and schemes, was drafted by Song

Ding and Dr. U. Bierbach and edited by Dr. U. Bierbach before submission to the journal.

Since the publication, changes in both format and content have been made to maintain a

consistent format throughout this work. The experiments described were performed by

Song Ding, except for the the cytotoxicity assays which were contributed by Dr. Xin

Qiao.

54

2.1. Introduction and Design Rationale

Platinum-based anticancer drugs continue to be the cornerstone of many chemotherapy

regimens despite their clinical drawbacks and the recent advent of more tolerable

molecularly targeted therapies[3] [269]. DNA-targeted agents acting by novel

mechanisms at the molecular level show considerable promise as potential treatments for

intractable tumors [270]. In particular, we and others have demonstrated that DNA

adducts that do not mimic the cytotoxic lesions produced by the anticancer agent cisplatin

(cis-diamminedichloridoplatinum(II), Figure 2.1) may overcome tumor resistance to the

clinical platinum drugs [271-273]. Platinum–acridine agents, represented by the

prototypical structures in Figure 2.1, unlike cisplatin, do not induce DNA cross-links but

form monofunctional–intercalative adducts [274]. The unique structural and

thermodynamic features of this dual binding mode and the rapid formation of the hybrid

adducts in cellular DNA result in an up to 500-fold higher cytotoxic potency compared to

cisplatin in aggressive, rapidly proliferating cancers [275-279]. While the cytotoxic

properties of the platinum–acridines translate into promising tumor growth inhibition in

vivo, the dose-limiting systemic toxicity observed in mice suggests that modifications of

these agents are necessary to improve their pharmacological properties [279].

Specifically, we are interested in attaching the platinum–acridines as cytotoxic

“warheads” to receptor- or tumor tissue-targeted vehicles. To achieve this goal, it is

necessary to install chemically or enzymatically reversible linkers in the original

structures. This is not a trivial task as many modifications made to a cytotoxic agent for

the purpose of delivering it may compromise its target interactions and decrease its

potency [280].

55

Figure 2.1. Structures of cisplatin and platinum-acridine derivatives.

Figure 2.2. Design of the combinatorial screening assay. (A) Synthesis of fragment

libraries containing variable and functionalizable groups R1–R3. For the color-coding of

residues, Figure 2.3. (B) Microscale coupling (“click”) reactions. (C) Qualitative and

quantitative LC-ESMS analysis. (D) High-throughput screening in cancer cells on 96-

well plates. (E) SAR data analysis and hit identification.

56

To accelerate the preclinical development of our technology, we have developed new

methodology based on efficient platinum-mediated click chemistry to generate and screen

a modular library of functionalized, conjugatable platinum–acridines. Fragment-based

approaches using simple condensation and click chemistries have demonstrated great

utility for the design of enzyme-targeted inhibitors [281, 282]. Here, we demonstrate, for

the first time, that modular library screening can be applied to DNA-interacting

anticancer agents. Proof-of-concept data is presented demonstrating that the new assay

provides a powerful tool for establishing structure–activity relationships (SAR) and

identifying candidates for the desired biological application.

2.2. Results and Discussion

2.2.1. Design of the combinatorial screen assay: assembly and characterization

To generate the desired compound library, sets of suitably modified platinum

complexes and 9-aminoacridine derivatives, the two building blocks of the hybrid agents,

were synthesized. We then used highly efficient amine addition to platinum-coordinated

nitrile ligand, a classical metal-catalyzed reaction for forming CN bonds [283], to

generate the amidine-linked hybrids (based on the prototype with X = NH in Figure 2.1).

The clean conversion to hybrid agent without the formation of stoichiometric by-

products, which is reminiscent of organic click chemistry [284], provides an ideal platform

for generating diversity within the library of platinum–acridines (for details see the

Experimental Section).

57

Figure 2.3. Library of platinum (P) and acridine (A) building blocks with design

elements highlighted.

A total of 6 nitrile-modified platinum complexes and 10 acridines were synthesized as

building blocks for the modular click assembly of 60 hybrids (Figure 2.3). Structural

diversity in this small library was achieved by varying the nonleaving group(s) (R1) and

nitrile ligands (R2) in the platinum moieties, as well as the geometry of the acridine side

chain (R3). Several of the unmodified parent compounds studied previously were also

58

generated. Conjugatable functional groups were placed strategically within the platinum

and acridine moieties as attachment points for targeted carriers. These include hydroxyl

(in P5, P6, A3, and A4) and carboxylic acid (in A5–A7) groups, which lend themselves

as coupling partners in (reversible) ester, carbamate, and amide linkages [285]. In

addition, terminal azide groups were installed in extended side chains using amide (A8,

A9) and carbamate (A10) coupling chemistry. These derivatives are also of interest for

their use as imaging tools in bioorthogonal reactions[286] with alkyne-modified

fluorophores to study the subcellular distribution of the hybrid agents.

To demonstrate the utility of this approach, 60 micro-scale reactions were assembled

containing stoichiometric amounts of platinum (P) and acridine (A) and allowed to

incubate in DMF under conditions that produce hybrid agent without any detectable side

products (see experimental details and Appendix B1). Automated in-line high-

performance liquid chromatography–electrospray mass spectrometry (LC-ESMS) was

used to monitor the progress of the reactions and to calculate the yield of each hybrid.

This was possible by integrating the HPLC traces recorded at an acridine-specific

wavelength assuming the chromophores in the acridine precursors (A) and in the newly

formed hybrids show the same absorptivity. An example of an LC profile along with

ESMS characterization of the reaction mixture is given in Figure 2.4 for hybrid P6-A1

(for a complete set of 60 LC-MS profiles see APPENDIX B1). Note the absence of side

products and high yield (80%) of the conversion. Baseline separation of the two fractions

corresponding to unreacted acridine and product platinum–acridines was observed for all

60 samples analysed, which greatly facilitated quantification and mass spectrometric

characterization of the hybrids.

59

Species in solution were identified by their molecular-ion peaks and fragments

generated by in-source collision-induced dissociation (CID). The reaction mixtures were

diluted with appropriate amounts of media so that each sample contained the same

concentration of hybrid agent prior to incubations with cancer cells. NCI-H460 cells

were then exposed to a fixed concentration of 50 nM of each platinum–acridine and the

viability of the cancer cells relative to untreated control was assessed after 72 h of

incubation using a colorimetric cell proliferation assay. The assay was also performed

with the platinum and acridine precursors, which proved to be at least two orders of

magnitude less cytotoxic than the hybrids. These control experiments were necessary to

demonstrate that unreacted precursors in reactions that did not go to completion did not

significantly contribute to the inhibition of cell proliferation (see APPENDIX E. Figure

E.2. for an example). Finally, to further validate the pre-screening approach, selected

compounds of interest were resynthesized and their IC50 values determined in the same

cell line (Figure 2.2).

2.2.2. Validation of the combinatorial screen assay: biological evaluation and SAR

analysis.

To validate our combinatorial approach as a tool for establishing SAR in platinum–

acridines, the data set generated was examined by plotting cell viabilities for the 60

samples (P1-A1 to P6-A10). To gain insight into the relative importance of the

fragments P and A in each compound, samples were numerically sorted into groups of

hybrids containing the same platinum moiety and groups sharing the same acridine ligand

(Figure 2.5). The former alignment demonstrates that for hybrids containing the same

60

Figure 2.4. High-throughput LC-ESMS analysis of “click” reaction mixtures. (A)

Reverse-phase HPLC trace for the reaction of platinum precursor P6 with acridine A1.

(B) ESMS spectrum of A1 (HPLC fraction with retention time 8.8 min) recorded in

positive-ion mode. Characteristic molecular and fragment ions: [M+H]+ (m/z 252.2) (C)

ESMS spectrum of P6-A1 (HPLC fraction with retention time 9.6 min) recorded in

positive-ion mode. Characteristic molecular and fragment ions: [M]+ (m/z 613.2),

[M+H]2+ (m/z 307.1, z = 2), [M-OH+H]+ (m/z 596.2), [M-[PtCl(pn2-OH)]]+ (free acridine–

amidine ligand, m/z 293.3) [M-2H-Cl]+ (m/z 575.3), [M-H-Cl]2+ (m/z 288.3, z = 2), [M-

pn2-OH-Cl]+ (m/z 485.2). The structures of A1 and P6-A1 are shown as insets in (B) and

(C).

61

platinum moiety (P1–P6) the biological activity usually decreases significantly with

increasing degree of modification and length of the (functionalized) 9-aminoacridine side

chain (Figure 2.5A). Analysis of the global activity profiles for P1–P6 also shows that

variation of the nitrile ligand (R2 = Et in P1, P3, and P5 vs. R2 = Me in P2, P4, and P6)

leads to little variation in activity and produces pairs of most similar compounds. By

contrast, replacement of the ethane-1,2-diaminoethane (en) non-leaving group with

ammine (NH3) ligands (P1/P2 vs. P3/P4) enhances the cytotoxicity of hybrids containing

acridines A5–A7. For several of the less active derivatives introduction of rac-1,3-

diaminopropan-2-ol (pn2-OH) as a nonleaving group (in P5 and P6) resulted in improved

cancer cell kill (average cell viabilities in Figure 2.5A were 55% for P1/P2, 45% for

P3/P4, and 42% for P5/P6, relative to control cells). Likewise, a plot of cell viabilities

for the 10 types of hybrids defined by common acridine moieties (Figure 2.5B) gives

important clues about SAR in this library of compounds. In general, the derivatives

containing ethylene groups in the acridine side chain performed better than those

containing extended propylene chains, based on average cell viabilities of 15% and 30%

calculated for the pairs A1/A3 and A2/A4, respectively. Derivatives A6 and A7, which

contain extended, carboxylic acid-modified side chains that only differ in the positioning

of the secondary amino function (Figure 2.3) produce an average reduction in cell

viability of merely 50% while showing similar activity profiles. Extension of the side

chains to contain azide functionalities (A8–A10) follows the same trend and leads to a

further decrease in cytotoxicity of the corresponding hybrids, which form a cluster of

highly similar profiles (Figure 2.5B). On the basis of mean cell viabilities calculated

62

across the entire set of 60 library members, hybrids containing A1, A3, and P6 resulted in

the strongest cell kill effect (Figure 2.6).

To assess the utility of the library screen as a tool for target compound identification, a

structurally diverse subset of four analogues of interest was resynthesized and tested in

NCI-H460 cells. IC50 values were calculated from the corresponding dose–response

curves (see Appendix E, Figure E.1) and are summarized along with data acquired

previously for the parent compounds in Table 2.1. Of the hybrids chosen, the azide-

functionalized compound P1-A8 required the highest concentration (110 ± 8 nM) to

inhibit cell proliferation by 50%, consistent with its relatively poor performance in the

prescreen. (It should be noted, however, that this derivative is still an order of magnitude

more potent than cisplatin in this cell line [276].) Compounds that reduced cell viability

to levels of 20% of control in the high-throughput assay resulted in significantly lower

IC50 values. The hydroxy-modified hybrids P4-A3 and P6-A1 showed the highest cell-

kill potential with the latter reaching cytotoxicity levels in the low-nanomolar

concentration range similar to the unmodified platinum–acridines P1-A1 and P2-A1

[277, 279].

63

Fig

ure

2.5

. B

iolo

gic

al a

ctiv

ity p

rofi

les

for

the

60 c

om

pounds

bas

ed o

n t

he

via

bil

itie

s of

trea

ted

cel

ls r

elat

ive

to u

ntr

eate

d

contr

ol

det

erm

ined

in a

colo

rim

etri

c ce

ll p

roli

fera

tion a

ssay.

The

test

com

pounds

are

sort

ed a

nd c

olo

r-co

ded

by c

om

mon

pla

tinum

moie

ties

in (

A)

and b

y c

om

mon a

crid

ine

moie

ties

in (

B).

E

rror

bar

s in

dic

ate

± s

tandar

d d

evia

tions

for

sets

of

thre

e

dat

a poin

ts f

or

each

com

pound.

64

Figure 2.6. Relative potencies of acridine (A) and platinum (B) fragments as building

blocks in hybrid agents expressed as ± % deviations from the mean cell viability (M.V.)

determined across the entire set of 60 library members.

Table 2.1. Summary of structural elements and biological activity for selected library

members.

Compound R1 R2 Functional group attached to R1 or R3 IC50 (nM)c

P1-A1a en Et b 12 ± 2

P1-A8 en Et N3 110 ± 8

P2-A1a en Me b 2.8 ± 0.3

P3-A7 (NH3)2 Et COOH 90 ± 5

P4-A3 (NH3)2 Me OH 45 ± 9

P6-A1 pn2-OH Me OH 8.6 ± 0.5

a References [277] and [278]. b Unmodified derivative. c IC50 values are reported as the

mean ± standard deviation for at least 2 individual experiments performed in triplicate.

65

The pre-screening method provides insight into the relative importance of the two

functionally interdependent components, based on the similarity of activity profiles and

clustering of individual library members. Because the method is based on testing

reaction mixtures without purification steps at an arbitrarily fixed concentration of test

compound, the cell viabilities extracted from it do not correlate with IC50 values, which

have to be calculated from the corresponding drug–response curves. Potential

complications in evaluating the biological activity of library members may arise in cases

in which incomplete conversion of the building blocks to hybrid is observed. Under such

circumstances, unreacted platinum or acridine precursors may contribute to the cell kill

observed in the pre-screen. Furthermore, synergistic effects between multiple

components, which might lead to enhanced cell kill, cannot be completely ruled out.

Generally these mixtures resulted in poor inhibition of cancer cell proliferation,

indicating that both the precursors and the hybrid agent were only marginally cytotoxic.

However, in some cases, such as P3-A7, which showed only 50% conversion but reduced

cell viability by more than 70%, it had to be confirmed that the hybrid and not unreacted

precursor caused the cell kill, which proved to be the case (see the Appendix E, Figure

E.2). Finally, in addition to demonstrating the feasibility of our library design and its

validity as a prescreening tool, we have identified promising hydroxyl-modified hybrids

P4-A3 and P6-A1 as potential “warheads” for the desired application of targeted prodrug

delivery.

66

2.3. Conclusion

In summary, we have demonstrated the potential of library screening for delineating

SAR in a class of DNA-targeted hybrid agents and for lead discovery. The modular click

approach in conjunction with the facile preparation of building blocks provides a

powerful tool for efficient screening of second-generation derivatives of these promising

cytotoxics. The use of metal-based coupling chemistry as an efficient screening tool for

the discovery of metal-containing (hybrid) pharmacophores is a previously unexplored

approach. Here, we provided proof-of-concept data by generating 60 hybrid agents and

testing them in one cancer cell line. We envision that this assay can easily be extended to

multi-step library synthesis, for instance by combining the platinum-mediated additions

with other forms of click or coupling chemistries. Likewise, parallel testing in multiple

cell lines would add another dimension to this assay. Such an approach would ultimately

lend itself to cluster analysis of extended databases with the ultimate goal of designing

personalized oncology drugs that can be targeted to specific cancer types.

2.4 Experimental Section

2.4.1. Reagents and Instrumentation.

All reagents were used as obtained from commercial sources without further

purification unless indicated otherwise. 1H NMR spectra of the target compounds and

intermediates were recorded on a Bruker Advance 300 MHz instrument. Proton-

decoupled 13C NMR spectra were recorded on a Bruker Advance 300 instrument

67

operating at 75 MHz. Chemical shifts (δ) are given in parts per million (ppm) relative to

tetramethylsilane (TMS). 1H NMR data are reported in the conventional form including

chemical shifts (δ, ppm), multiplicities (s = singlet, d = doublet, t = triplet, q = quartet, m

= multiplet, br = broad), coupling constants (Hz), and integral intensities). 13C{1H} NMR

data are reported as chemical shifts (δ, ppm). HPLC-grade solvents were used for all

HPLC and mass spectrometry experiments. LC-ESMS analysis was performed on an

Agilent 1100LC/MSD ion trap mass spectrometer equipped with an atmospheric pressure

electrospray ionization source. Eluent nebulization was achieved with a N2 pressure of 50

psi, and solvent evaporation was assisted by a flow of N2 drying gas (350 °C). Positive-

ion mass spectra were recorded with a capillary voltage of 2800 V and a mass-to-charge

scan range of 150 to 2200 m/z. To establish the purity of target compounds, samples

were diluted in methanol containing 0.1 % formic acid and analyzed using a 4.6 mm

150 mm reverse-phase Agilent ZORBAX SB-C18 (5 µm) analytical column at 25 °C.

The following solvent system was used for the characterization of “click” reactions:

solvent A, optima water, and solvent B, methanol/0.1% formic acid, with a flow rate of

0.5 mL/min and a gradient of 95% A to 5% A over 12 minutes. The purity of the

resynthesized compounds (> 95%) was confirmed using the following solvent system:

solvent A, optima water, and solvent B, methanol/0.1% formic acid, with a flow rate of

0.5 mL/min and a gradient of 95% A to 5% A over 25 minutes. A wavelength range of

363–463 nm was used to monitor the reactions and to determine reaction yields. Peak

integration was done using the Agilent LC/MSD Trap Control 4.0 data analysis software.

The synthetic precursors [PtCl2(en)], cisplatin , and rac-1,3-diaminopropan-2-ol [287], 9-

phenoxyacridine (2-1) [65], and building blocks N1-(acridin-9-yl)-N2-methylethane-1,2-

68

diamine (A1) [65], and N1-(acridin-9-yl)-N3-methylpropane-1,3-diamine (A2) [288] were

synthesized according to the cited methods.

2.4.2. Synthesis and characterization of acridine building blocks.

The synthesis of 2-((2-(acridin-9-ylamino)ethyl)amino)ethan-1-ol (A3). A mixture

of phenoxyacridine (2-1) (2.71 g, 0.01 mol) and 2-(2-aminoethylamino)ethanol (2-4)

(1.14 g, 0.011 mol) in 15 mL of anhydrous THF was refluxed for 16 h. The solvent was

evaporated off and the residue was dissolved in 30 mL of acetone. To this solution were

added 5 mL of concentrated HCl and the mixture was stirred at 4 C for 3 hours. A

yellow precipitate formed which was recovered by filtration, resuspended in 50 mL of 2

M ammonium hydroxide, and stirred at room temperature for 30 min. The aqueous phase

was extracted with CH2Cl2 and the organic phase was collected, dried over Na2SO4, and

concentrated using rotary evaporation, affording 2.57 g of the free base as an yellow solid

(Yield: 92%). 1H NMR (300 MHz, CDCl3) δ 8.27 (d, J = 8.5 Hz, 2H), 7.69 (d, J = 7.6 Hz,

2H), 7.59 (t, J = 7.3 Hz, 1H), 7.26 (t, J = 7.5 Hz, 2H), 4.50 (s, 1H), 3.88 (t, J = 6.2 Hz,

2H), 3.47 (t, J = 5.7 Hz, 1H), 2.90 (t, J = 6.2 Hz, 2H), 2.63 (t, J = 5.7 Hz, 2H). 13C NMR

(75 MHz, CDCl3) δ 151.70, 129.74, 124.84, 121.30, 60.39, 51.27, 50.18, 40.90. MS (ESI,

positive-ion mode): calculated for C17H20N3O ([M+H]+), 282.15; found: 282.3.

69

Sch

eme

2.1

. S

ynth

esis

of

acri

din

e (A

) buil

din

g b

lock

s.

70

N1-(acridin-9-yl)-N3-methylpropane-1,3-diamine (A2) was prepared using the

procedure described for A3. Yield: 94%. 1H NMR (300 MHz, CDCl3) δ 8.12 (d, J = 8.7

,2H), 8.02 (d, J = 8.7, 2H), 7.61 (t, J = 6.7 Hz, 2H), 7.25 (t, J = 7.5 Hz,2H), 4.01 (t, J

= 5.9 Hz, 2H), 2.96 - 2.75 (m, 2H), 2.53 (s, 3H), 1.94-1.68 (p, J = 5.9 Hz , 2H). MS (ESI,

positive-ion mode): calculated for C17H20N3 ([M+H]+), 266.36; found: 266.2.

3-((2-(acridin-9-ylamino)ethyl)amino)propan-1-ol (A4) was prepared using the

procedure described for A3. Yield: 86%. 1H NMR (300 MHz, CDCl3) δ 8.11 (d, J = 8.5

Hz, 2H), 8.02 (d, J = 8.6 Hz, 2H), 7.72 - 7.54 (m, 2H), 7.35-7.29 (m, 2H), 3.87-3.82 (m,

4H), 3.16 - 2.64 (m, 4H), 1.79 (p, J = 6.5 Hz, 2H). 13C-NMR (75 MHz, CDCl3) δ 151.97,

147.93, 130.25, 127.65, 123.25, 122.70, 116.33, 62.15, 49.48, 48.91, 47.68, 31.98. MS

(ESI, positive-ion mode): calculated for C18H22N3O ([M+H]+), 296.38; found: 296.3.

The synthesis of (2-(acridin-9-ylamino)ethyl)glycine (A5). A mixture of

phenoxyacridine (2-1) (2.71 g, 0.01 mol) and 2-((2-aminoethyl)glycine (2-6) (1.3 g,

0.011 mol) in 20 mL of dry MeOH was refluxed for 3 h. The yellow solid that

precipitated during the reaction was collected by filtration, washed with hot THF and

ether, and dried in a vacuum, affording 2.55 g of the product as a yellow solid (Yield: 86

%). 1H NMR (300 MHz, MeOH-d4) δ 8.22 (m, 2H), 7.55 (t, J = 7.6 Hz, 2H), 7.43 (m,

2H), 7.15 (t, J = 7.6 Hz, 2H), 4.08 (d, J = 6.0 Hz, 2H), 3.22 (s, 2H), 3.16 (t, J = 5.9 Hz,

2H). 13C NMR spectrum of this compound was not obtained due to limited solubility of

the compound. MS (ESI, positive-ion mode): calculated for C17H18N3O2 ([M+H]+),

296.34; found: 296.3.

(3-(acridin-9-ylamino)propyl)glycine (A6) was prepared using the procedure

described for A5. Yield: 91%. 1H NMR (300 MHz, CDCl3) δ 8.21 (d, J = 8.4 Hz, 2H),

71

7.54 (m, 3H), 7.12-7.18 (m, 4H), 6.93-6.62 (m, 2H), 3.91 (t, J = 6.4 Hz, 2H), 3.21 (s,

2H), 3.01 (t, J = 7.2 Hz, 2H), 2.03 (p, J = 6.5 Hz, 2H). 13C NMR (75 MHz, CDCl3) δ

166.95, 157.31, 152.15, 130.13, 129.21, 125.84, 120.79, 118.56, 115.13, 49.87, 48.74,

45.42, 27.92. MS (ESI, positive-ion mode): calculated for C18H20N3O2 ([M+H]+), 310.37;

found: 310.2.

3-((2-(acridin-9-ylamino)ethyl)amino)propanoic acid (A7) was prepared using the

procedure described for A5. Yield: 84%. 1H NMR (300 MHz, D2O) δ 8.04 (d, J = 8.7

Hz, 2H), 7.83 (dd, J = 8.4, 7.0 Hz, 2H), 7.58 - 7.34 (m, 4H), 4.33 (t, J = 6.1 Hz, 2H), 3.60

(t, J = 5.8 Hz, 1H), 3.33 (t, J = 6.3 Hz, 2H), 3.33 (t, J = 6.3 Hz, 2H). 13C NMR (300

MHz, D2O) δ 166.95, 157.31, 152.15, 130.13, 129.21, 125.84, 118.56, 115.13, 49.87,

48.80, 45.42, 27.92. MS (ESI, positive-ion mode): calculated for C18H20N3O2 ([M+H]+),

310.37; found: 310.3.

The Boc-protected acridine derivative (2-9) (1.36 g, 4.6 mmol) was synthesized as

follows. A5 was suspended in 30 mL of anhydrous methanol, to which was added Boc2O

(1.3 g, 6 mmol) in 5 mL of anhydrous MeOH at 0-5 ºC maintained with an ice bath. The

mixture was then stirred at room temperature for 4 h. The solvent was removed by rotary

evaporation and residue was dissolved in 10 mL of dichloromethane and precipitated

with 200 mL of anhydrous diethyl ether. The solid was recovered by filtration and dried

in a vacuum affording 1.79 g (99%) of the product as a yellow solid, which was used in

the next step without further purification.

Compound 2-9 (1 g, 2.52 mmol) and 1,1'-carbonyldiimidazole (CDI, 533 mg, 3.28

mmol) were combined in 20 mL of anhydrous DMF. The mixture was heated to 40-50 ºC

and stirred for 6 h. Then the solution was cooled to 0-5 º C in an ice bath and 264 mg of

72

2-azidoethanamine dissolved in 3 mL of anhydrous DMF were added. The mixture was

stirred at 0-5 º C for 4 h. DMF was removed by vacuum distillation at 35-40 ºC, and the

residue was redissolved in 40 mL of dichloromethane and washed with 1 M HCl (3 × 20

mL). The organic phase was collected, dried with anhydrous Na2SO4, and concentrated to

afford an orange oil. To remove the Boc group, the residue was dissolved in 6 mL of a

1:1 mixture of anhydrous dichloromethane and trifluoroacetic acid and stirred at room

temperature for 3 h. The reaction was quenched by adding 10 mL of 1 M NaOH solution.

The crude product was extracted from NaOH solution with DCM, dried over anhydrous

Na2SO4, and concentrated. The product was purified by flash chromatography (Al2O3,

DCM:MeOH, 30:1). Yield: 0.59 g (64 %). 1H NMR (300 MHz, CDCl3) δ 8.10 (d, J = 8.7

Hz, 2H), 7.97 (d, J = 8.6 Hz, 2H), 7.60 (t, J = 8.3, 6.8 Hz, 2H), 7.40 - 7.14 (m, 3H), 3.89

(t, J = 5.6 Hz, 2H), 3.50 - 3.23 (m, 6H), 2.99 (t, J = 5.6 Hz, 2H). 13C NMR (75 MHz,

CDCl3) δ 172.56, 152.98, 146.04, 131.21, 125.75, 123.56, 122.97, 115.48, 50.72, 48.69,

48.47, 44.85, 38.82, 36.22. MS (ESI, positive-ion mode): calculated for C19H22N7O

([M+H]+), 364.42; found: 364.3.

3-((2-(acridin-9-ylamino)ethyl)amino)-N-(2-azidoethyl)propanamide (A9) was

prepared using the procedure described for A8. Yield: 64%. 1H NMR (300 MHz, CDCl3)

δ 8.09 (d, J = 8.7 Hz, 2H), 8.02 (d, J = 8.7 Hz, 2H), 7.59 (t, J = 7.6 Hz, 2H), 7.41 (s, 1H),

7.27 (t, J = 7.5 Hz, 2H), 4.95 (brs, 2H), 3.92 (t, J = 5.6 Hz, 2H), 3.44 (s, 4H), 2.91-3.14

(m, 4H), 2.53 (t, J = 6.0 Hz, 2H). 13C NMR (75 MHz, CDCl3) δ 152.09, 149.60, 129.65,

129.23, 123.60, 121.91, 115.75, 51.37, 51.32, 36.51, 29.15. MS (ESI, positive-ion mode):

calculated for C20H24N7O ([M+H]+), 378.45; found: 378.3.

73

The synthesis of 2-((2-(acridin-9-ylamino)ethyl)amino)ethyl (2-

azidoethyl)carbamate (A10). The Boc-protected acridine derivative (2-11) was prepared

as described for 2-9 starting with compound A3. Compound 2-11 (1 g, 2.62 mmol), TEA

(793 mg, 7.85 mmol) and 4-nitrobenzyl chloroformate (732 mg, 3.4 mmol) were

dissolved in 20 mL of anhydrous DCM. The mixture was stirred at room temperature for

16 h. Then 271 mg of 2-azidoethanamine dissolved in 5 mL of anhydrous DCM was

added and the reaction was stirred for another 8 h. The solvent was removed using

vacuum distillation and the residue was redissolved in 40 mL of DCM and washed with 1

M HCl (3 × 20 mL). The organic phase was collected, dried with anhydrous Na2SO4, and

concentrated to afford an orange oil. To remove the Boc group, the orange oil was

dissolved in 6 mL of a 1:1 mixture of anhydrous dichloromethane and trifluoroacetic acid

and stirred at room temperature for 3 h. The reaction was quenched by adding 10 mL of 1

M NaOH solution. The crude product was extracted from NaOH solution with DCM,

dried over anhydrous Na2SO4, and concentrated. The product was further purified by

flash chromatography (Al2O3, DCM:MeOH, 30:1). Yield: 0.73 g (71 %). 1H NMR (300

MHz, CDCl3) δ 8.12 - 7.96 (m, 4H), 7.59 (t, J = 7.8 Hz, 2H), 7.23 (t, J = 7.6 Hz, 2H),

5.65 (brs, 1H), 4.27 (t, J = 4.8 Hz, 2H), 3.92 (t, J = 5.7 Hz, 1H), 3.61 - 3.20 (m, 4H), 3.01

(t, J = 5.7 Hz, 2H), 2.96 (t, J = 4.9 Hz, 2H). 13C NMR (75 MHz, CDCl3) δ 156.53,

152.69, 145.93, 131.22, 125.98, 123.3, 123.00, 115.07, 64.48, 50.97, 48.35, 48.17, 47.83,

40.45. MS (ESI, positive-ion mode): calculated for C20H24N7O2 ([M+H]+), 394.45; found:

394.3.

74

Scheme 2.2. Synthesis of platinum (P) building blocks. en = ethylenediamine; pn2-OH = 2-

hydroxy-1,3-propanediamine.

2.4.3. Synthesis and characterization of platinum building blocks.

Synthesis of precursor [PtCl(EtCN)(en)]Cl (P1) [279, 289]. The complex [PtCl2(en)]

(2-12) (0.50 g, 1.54 mmol) was heated under reflux in 25 mL of dilute HCl (pH 4) with

propionitrile (6.85 mL, 98.5 mmol) until the yellow suspension turned into a colorless

solution (~2 h). Solvent was removed by rotary evaporation, and the pale-yellow residue

was redissolved in 10 mL of dry methanol. A small amount of an insoluble yellow solid

was removed by membrane filtration and the colorless filtrate was added directly into 250

mL of vigorously stirred dry diethyl ether, affording P1 as an off-white, extremely

hygroscopic microcrystalline precipitate. Yield: 0.48 g (83%). 1H NMR (300 MHz, D2O)

δ 5.72, 5.64 (2 br s,0.7 H, HD exchange), 2.87(q, J=7.5 Hz, 2H), 2.48 - 2.75 (m, Pt

satellites, 4 H), 1.3 (t, J=7.5 Hz, 3H,).

Synthesis of precursor [PtCl(MeCN)(en)]Cl (P2). P2 was prepared using the similar

procedure as described for P1 as the white solid with the yield 89%. 1H NMR (300 MHz,

75

D2O) δ 5.80, 5.65 (2 br s,1H, HD exchange), 2.87(q, J=7.5 Hz, 2H), 2.55 - 2.65 (m, Pt

satellites, 4 H), 2.53 (s, 3H).

Synthesis of precursor [PtCl(EtCN)(NH3)2]Cl (P3). P3 was prepared using the

procedure described for P1. Yield: 74%. 1H NMR (300 MHz, D2O) δ 2.89 (t, J = 7.6,

2H), 1.30 (t, J = 7.5, 3H).

Synthesis of precursor [PtCl(MeCN)(NH3)2]Cl (P4). P4 was prepared using the

procedure described for P1. Yield: 63%. 1H NMR (300 MHz, D2O) δ 2.53 (3 H, m, Pt

satellites), 4.35, 4.48 (4 H, HD exchange, 2 br s).

Synthesis of precursor [PtCl(EtCN)(pn2-OH)]Cl (P5). P5 was prepared using the

procedure described for P1. Yield: 91%. 1H NMR (300 MHz, D2O) δ 4.26 (m, 1H), 2.99 -

2.54 (m, 6H), 1.39 - 0.96 (t, J = 7.5 Hz, 3H).

Synthesis of precursor [PtCl(MeCN)(pn2-OH)]Cl (P6). P6 was prepared using the

procedure described for P1. Yield: 77%. 1H NMR (300 MHz, D2O) δ 4.28 (m, 1H), 2.85 -

2.53 (m, 4H), 2.53 (s, 3H).

2.4.4. Experimental procedure for “click” reactions.

Stock solutions of the platinum–nitrile complexes P1–P6 (20 mM) and acridine

derivatives A1–A10 (20 mM) were prepared in anhydrous DMF. Concentrations of A1–

A10 were determined spectrophotometrically (λmax = 413 nm, ε = 10,000 M-1 cm-1). The

reactions between platinum complexes and acridine derivatives were carried out in 1.5-

mL Eppendorf tubes by mixing equal volumes (100 μL) of platinum–nitrile complex and

acridine derivative. The reaction mixtures were placed in a shaker and incubated at 4 C

76

for 5 days. To detect and characterize the hybrid agents and to determine conversion

yields, 1-μL samples were removed from each reaction and diluted with 1000 μL of

methanol containing 0.1% formic acid prior to in-line LC-ESMS analysis. Sample

injections into the LC unit were accomplished via a thermostatted (4 C) autosampler.

Chromatographic separations were performed with a 4.6 mm 150 mm reverse-phase

Agilent ZORBAX SB-C18 (5 µm) analytical column with the column temperature

maintained at 25 °C. The binary mobile phase consisted of: solvent A, optima water, and

solvent B, methanol/0.1% formic acid delivered at a gradient of 95% A to 5% A over 12

minutes and a flow rate of 0.5 mL/min. The formation and the extent of conversion of the

platinum–acridines was monitored in the corresponding chromatograms using the

LC/MSD Trap Control 4.0 data analysis software.

Scheme 2.3. Synthesis of platinum-acridine derivatives. en = ethylenediamine; pn2-OH =

2-hydroxy-1,3-propanediamine.

77

2.3.5. Typical Procedure for the Preparation of Platinum–Acridine Derivatives.

Synthesis of [PtCl(en)C20H25N4O](NO3)2 (P1-A3). Complex P1 (381 mg, 1.00 mmol)

was converted to its nitrate salt by reaction with AgNO3 (162 mg, 0.950 mmol) in 7 mL

of anhydrous DMF. AgCl was removed by syringe filtration, and the filtrate was cooled

to -10 °C. Acridine precursor A3 (282 mg, 0.100 mmol) was added to the solution, and

the suspension was stirred at -10 °C for 24 h. After treatment with activated carbon, the

reaction mixture was added into 300 mL of vigorously stirred diethyl ether. The yellow

precipitate was stirred for approximately 30 min and then recovered by membrane

filtration and dried in a vacuum overnight. The solid was dissolved in anhydrous

methanol containing one equivalent of HNO3 (1 M) and stirred at room temperature for

30 min, and the crude dinitrate salt was precipitated with 300 mL of anhydrous diethyl

ether. The product was further purified by recrystallization from hot ethanol to give 606

mg of the dinitrate salt as a yellow microcrystalline solid (Yield: 82%). 1H NMR (300

MHz, MeOH-d4) δ 8.42 (d, J = 8.4 Hz, 2 H), 7.88 (t d, J = 6.9, 1.2 Hz, 2 H), 7.74 (d, J =

8.4 Hz, 2 H), 7.51 (t d, J = 7.2, 1.2 Hz, 2H 1 H), 6.12 (br s, 1 H), 5.32 (s, 2 H), 5.11 (s, 2

H), 4.32 (t, J = 6.7 Hz, 2 H), 3.89 (t, J = 6.7 Hz, 2H), 3.63 (t, J = 5.0 Hz, 2 H), 3.47 (t, J =

5.1 Hz, 2 H), 2.94 (q, J = 7.2 Hz, 1 H), 2.47 (m, 4 H), 1.18 (t, J = 7.2 Hz, 3 H). 13C NMR

(75 MHz, MeOH-d4) δ 171.73, 160.21, 141.38, 136.63, 126.48, 125.35, 119.73, 114.13,

60.70, 29.26, 11.85. MS (ESI, positive-ion mode): m/z for C22H32ClN6OPt ([M]+),

627.06; found: 627.3.

Synthesis of [PtCl(en)C22H27N8O](NO3)2 (P1-A8). P1-A8 was prepared using the

same procedure as described for P1-A3 and was recovered with a yield of 81%.1H NMR

(300MHz, DMF-d7) δ 13.90 (s, 1H), 9.99 (s, 1H), 8.83 - 8.52 (m, 3H), 8.28 - 7.92 (m,

78

4H), 7.62 (td, J = 6.6, 1.5 Hz 2H), 6.73 (s, 1H), 5.84 (s, 2H), 5.51 (s, 2H), 4.54 (q, J = 5.7

Hz, 2H), 4.44 (s, 2H), 4.15 (t, J = 5.9 Hz, 2H), 3.44 - 3.52 (m , 4H), 3.06 (q, J = 7.9 Hz,

2H), 2.67 (s, 4H), 1.33 (t, J = 7.5 Hz, 3H). 13C NMR (75 MHz, DMF-d7) δ 170.37,

169.56, 158.62, 140.17, 135.27, 125.92, 123.80, 118.90, 112.94, 50.32, 48.97, 48.79,

46.34, 38.78, 28.67, 11.00. MS (ESI, positive-ion mode): for C24H34ClN10OPt ([M]+),

709.13; found: 708.5.

Synthesis of [PtCl(NH3)2C19H23N4O](NO3)2 (P4-A3). P4-A3 was prepared using the

same procedure as described for P1-A3 and was recovered with a yield of 87%. 1H NMR

(300 MHz, MeOH-d4) δ 8.40 (d, J = 8.8 Hz, 2H), 7.87 (t, J = 6.8, 2H), 7.72 (dd, J = 8.7,

1.2 Hz, 1H), 7.49 (d, J = 8.4 Hz, 2H), 4.31 (t, J = 7.0 Hz, 2H), 4.10 (brs, 2H), 3.91 (t, J =

6.8 Hz, 1H), 3.75 (s, 2H), 3.65 (t, J = 4.9 Hz, 2H), 3.49 (t, J = 4.9 Hz, 2H), 2.58 (s, 3H).

13C NMR (75 MHz, MeOH-d4) δ 167.49, 160.02, 141.40, 136.55, 126.49, 125.30, 119.77,

114.14, 60.73, 49.86, 48.16, 47.39, 23.02. MS (ESI, positive-ion mode): for

C19H28ClN6OPt ([M]+), 587.00; found: 585.2.

Synthesis of [PtCl(pn2-OH)C18H21N4](NO3)2 (P6-A1). P6-A1 was prepared using the

same procedure as described for P1-A3 and was recovered with a yield of 92%. 1H NMR

(300 MHz, DMF-d7) δ 13.92 (s, 1H), 9.89 (s, 1H), 8.68 (d, J = 8.7 Hz, 2H), 7.9 -8.11 (m,

4H), 7.62 (td, J = 6.4, 3.1 Hz, 2H), 6.21 (s, 1H), 5.65 (s, 1H), 5.60 (s, 1H), 5.13 (s, 2H),

4.87 (s, 1H), 4.50 (s, 2H),4.12 (t, J = 6.3 Hz 2H), 4.06 (s, 1H), 3.50 (s, 4H), 3.18 (s, 3H),

2.59-2.97 (m, 4H). 13C NMR (75 MHz, DMF-d7) δ 165.94, 158.64, 139.96, 135.33,

125.29, 123.91, 118.98, 112.80, 65.84, 48.38, 47.80, 33.79, 28.66. MS (ESI, positive-ion

mode): for C21H30ClN6OPt ([M]+), 613.04; found: 612.3.

79

Synthesis of [PtCl(NH3)2C21H24N4O2](NO3) (P3-A7). Platinum complex P3 (354 mg,

1 mmol) was converted to its nitrate salt by reaction with AgNO3 (162 mg, 0.95 mmol) in

7 mL of anhydrous DMF. AgCl was removed by syringe filtration, and the filtrate was

cooled to -10 °C. Acridine precursor A7 (310 mg, 0.1 mmol) was added to the solution,

and the suspension was stirred at 4 °C for 5 days. The mixture was poured into 300 mL

of vigorously stirred diethyl ether, and the precipitate was recovered by membrane

filtration and dried in a vacuum overnight. The product was further purified by

recrystallization from hot methanol to give 461.7 mg of the product as a yellow solid

(Yield: 67%). 1H NMR (300 MHz, MeOH-d4) δ 8.29 (d, J = 8.6 Hz, 2H), 7.92 - 7.58 (m,

4H), 7.41 (t, J = 7.7 Hz, 2H), 4.06 (t, J = 6.5 Hz, 2H), 3.74 (t, J = 6.4 Hz, 2H), 3.63 (t, J =

6.5 Hz, 2H), 3.02 (t, J = 7.9 Hz, 2H), 2.35 (t, J = 6.4 Hz, 2H), 1.23 (t, J = 8.0 Hz, 2H).

MS (ESI, positive-ion mode): for C21H30ClN6O2Pt ([M]+), 629.04; found: 629.2.

2.4.6. Cell proliferation assay

The human non-small cell lung cancer cell line, NCI-H460, was obtained from the

American Type Culture Collection (Rockville, MD, USA) and was cultured in RPMI-

1640 media (HyClone) containing 4.5 g/L glucose, 1.5 g/L sodium bicarbonate, 10 mM

HEPES, and 110 mg/L sodium pyruvate supplemented with 10% fetal bovine serum

(FBS), 10% penstrep (P&S), and 10% L-glutamine. Cells were incubated at a constant

temperature at 37 °C in a humidified atmosphere containing 5% CO2 and were

subcultured every 2 to 3 days in order to maintain cells in logarithmic growth.

80

Cell viability was assessed using the CellTiter 96 Aqueous Non-Radioactive Cell

Proliferation Assay (Promega, Madison, WI, USA) as described previously [277-278].

Briefly, NCI-H460 cell suspensions were harvested and seeded into 96-well microplates

at a density of 750 cells/well. The cells were preincubated at 37 °C overnight and then

treated with 50 nM compound from the library reactions or serial dilutions of selected

purified compounds (the DMF content in the mixtures was less than 1% at the highest

incubation concentration), as well as control samples containing platinum and acridine

precursors and samples containing media/DMF. After an incubation period of 72 h, 20

μL of MTS/PMS solution was added to each well and incubated at 37 °C for 4 h. The

absorbance of tetrazolium dye was measured at 490 nm using a plate reader. The fraction

of viable cells was calculated as a percentage of untreated control and is reported as the

mean ± standard deviation for 3 incubations of each compound. IC50 values were

calculated from non-linear curve fits using a sigmoidal dose–response equation in

GraphPad Prism (GraphPad Software, La Jolla, CA) and are averages of two individual

experiments performed in triplicate.

81

CHAPTER 3

USING FLUORESCENT POST-LABELING TO PROBE THE SUBCELLULAR

LOCALIZATION OF DNA-TARGETED PLATINUM

The work contained in this chapter was initially published in Angew.Chem.Int.Ed. in

2013 and J.Biol.Inorg.Chem. in 2014. The manuscript, including figures and schemes,

was drafted by Song Ding and Dr. U. Bierbach and edited by Dr. U. Bierbach before

submission to the journal. Since the publication, changes in both format and content have

been made to maintain a consistent format throughout this work. The experiments

described were performed by Song Ding, except for the the confocal microscopy and

cytotoxicity assays which were contributed by Dr. Xin Qiao and Dr. Fang Liu, and the

DNA unwinding experments was performed by Dr. Jimmy Suryadi.

82

3.1. Introduction and Design Rationale

Adducts in the nuclear DNA are the major cause of cancer-cell death triggered by

platinum-based anticancer drugs [290]. Thus, cellular uptake and accumulation,

distribution, and trafficking between subcellular compartments and, ultimately,

localization to the nucleus are crucial parameters in the mechanism of these agents.

Several techniques have been used to monitor the intracellular distribution of platinum

drugs. These include element-specific analytical methods and nondestructive absorption-

or emission-based imaging techniques, as well as electron microscopy [291, 292].

Fluorophore-tagged derivatives have provided insight into the uptake, distribution, and

intracellular transformation of platinum drugs [291]. Such an approach has to take into

consideration the organelle selectivity of the fluorophore, which may vary widely

depending on parameters such as molecular weight, partition coefficient (log P),

amphiphilic character, and pKa value [293]. Thus, one drawback of modifying platinum

drugs with organic fluorophores is that such conjugates may, at least in part, mimic the

properties of the reporter molecule [294]. This would be an undesired feature unless the

fluorescent group itself is a functionally important part of the bioactive molecule.

Likewise, bulky fluorophores may interfere with the DNA-binding mechanism of

platinum drugs. To circumvent these problems, we have developed a method based on

bioorthogonal ligation chemistry, which allowed us to fluorescently label platinum-

acridine hybrid agents in lung cancer cells. Herein, we report the development of this

technique and demonstrate, for the first time, that post-labeling is a powerful tool for

detecting DNA-targeted platinum drugs in subcellular and subnuclear structures.

83

Platinum-acridines, represented by the prototypical P1-A1 (Figure 3.1), are up to 500-

times more cytotoxic to non-small-cell lung cancer cells than the drug cisplatin [295].

Studies in NCI-H460 cells suggest that the cytotoxicity produced by P1-A1 is triggered

by high levels of monofunctional-intercalative hybrid adducts in cellular DNA [275],

which is supported by recent results from chemogenomic screening in S. cerevisiae [296].

Unfortunately, direct fluorescent imaging of platinum-acridines in cancer cells is difficult

because the fluorescence of the 9-aminoacridine is severely quenched when intercalated

into DNA (Appendix D, Figure D.1) [297]. We therefore developed a method that allows

for the labeling of platinum-acridines, in particular their DNA adducts, intracellularly

using a high-performance fluorescent dye in conjunction with a copper-catalyzed azide–

alkyne cycloaddition (CuAAC, click chemistry) [298, 299].

3.2. Results and Discussion

3.2.1. Post-labeling method

This strategy required modification of P1-A1 with an alkyne or azide group in a way

that would not interfere with the DNA- binding mode of the agent. Molecular models

derived from the NMR solution structure of the hybrid adduct [274] (Appendix F, Figure

F.1) suggest that the addition of a clickable group at the amidine N1-methyl group in P1-

A1 (Figure 3.1), which is protruding out of the DNA major groove, should satisfy this

requirement. While direct attachment of either functional group to this position using

polymethylene linkers was incompatible with the metal and resulted in complex

decomposition, our search identified a stable amide-linked azide derivative (P1-A9,

84

Figure 3.1), which also showed excellent stability in aqueous media (see Appendix B3).

Using gel-shift experiments to monitor DNA unwinding (Figure 3.2) and CD

spectropolarimetry, we confirmed that the azide-containing linker does not affect the

geometry of the hybrid adduct (APPENDIX C). Thus, P1-A9 was chosen for our labeling

technique in combination with the alkyne-modified form of the bright green-fluorescent

dye Alexa Fluor 488 (Figure 3.1).

Figure 3.1. Structures of compounds and reaction scheme for bioorthogonal fluorescent

labeling of DNA-targeted platinum–acridine hybrid agents.

85

Fig

ure

3.2

. G

el el

ectr

ophore

sis

of

pla

tinum

–ac

ridin

e-m

odif

ied,

neg

ativ

ely su

per

coil

ed pla

smid

D

NA

. pU

C19 m

odif

ied w

ith

com

pound P

1-A

9 (

A)

wit

h c

om

pound P

1-A

1 (

B).

Eac

h g

el i

s sh

ow

n w

ith t

he

corr

espond

ing p

lot

of

rela

xed

DN

A b

and

inte

nsi

ties

vs.

rb.

The

unw

indin

g a

ngle

s (ϕ

) ca

lcula

ted f

rom

ϕ =

18σ

/ r b

(c),

wher

e σ

is

the

super

hel

ical

den

sity

of

the

pla

smid

(

0.0

50)

and

r b(c

) is

the

max

imum

of

the

quad

rati

c poly

nom

ial

curv

e fi

ts, fo

r co

mpound

s 1 a

nd

2 a

re 1

7 ±

1

and 1

6 ±

1,

res

pec

tivel

y.

86

We established conditions for labeling the DNA adducts of P1-A9 using a CuAAC

reaction in a randomly platinated linearized plasmid. The procedure involved incubation

of the DNA with P1-A9 at varying platinum-to-nucleotide ratios (rb) and reaction of the

platinated DNA with Alexa Fluor 488-alkyne in the presence of a copper-based click-

reaction buffer, followed by purification of the samples and analysis of the labeled DNA

on agarose gels. A fluorescent image of the gel (at 520 nm) along with the ethidium

bromide-stained gel is shown in Figure 3.3 A. The bands on the gel show a pronounced

increase in Alexa Fluor 488 fluorescence intensity with an increasing platinum-to-

nucleotide ratio (lanes 3–5, top gel). To test whether the Alexa Fluor 488-alkyne was

ligated to covalently bound platinum-acridine on the plasmid, a sample of fluorescently

labeled DNA was incubated with NaCN, which reverses Pt-DNA adducts [300]. The

complete disappearance of Alexa Fluor fluorescence in lane 6 (Figure 3.3A, top gel) and

conversion of the platinated DNA into unmodified plasmid (Figure 3.3 A, bottom gel,

band of highest mobility) confirm that the platinum adducts were modified by the green-

fluorescent dye.

CD spectra were recorded of native DNA globally modified with P1-A9 before and

after undergoing ligation with the Alexa Fluor dye (Figure 3.3 B). The adducts formed by

P1-A9 give rise to positive Cotton effects in the 300–500 nm range, characteristic of the

induced circular dichroism (ICD) of the intercalated acridine chromophore.[301] After

labeling with Alexa Fluor 488-alkyne, the CD spectrum shows an additional negative

ICD band at 498 nm, which mimics the absorbance spectrum of the dye, a consequence

of the DNA chirality experienced by the ligated fluorophore. The hypsochromic shift of

four nanometers observed for the acridine-based long-wavelength ICD bands in the

87

labeled adduct relative to the unlabeled adduct of P1-A1 suggests that ligation of the

Alexa Fluor dye perturbs the hybrid binding mode, leading to partial unstacking of the

acridine moiety [274]. By contrast, a comparison of the ICD signatures of DNA-bound

compounds P1-A1 and P1-A9 shows that addition of the azide-containing side chain does

not compromise acridine intercalation. This observation corroborates the notion that P1-

A9, but not its fluorophore-tagged form, is a faithful mimic of P1-A1, which validates the

post-labeling approach.

88

Figure 3.3. (A) Electrophoretic analysis of copper-catalyzed ligation reactions between

linearized (BamH I), platinated pUC19 plasmid DNA and Alexa Fluor 488-alkyne with

reaction conditions indicated for each incubation (for details see Experimental Section).

The top gel shows Alexa Fluor fluorescence monitored at 520 nm. Note the increase in

band intensities and mobility shift with increasing platinum content. The latter band shift

is caused by accumulation of positive charge on the plasmid. The bottom panel depicts

the same gel stained with ethidium bromide. Saturation of DNA binding sites by Alexa

Fluor-tagged platinum–acridine causes inefficient staining by ethidium at rb = 0.2. (B)

CD spectra recorded in the ICD region for unmodified DNA (green trace), DNA

modified with P9-A1 (blue trace), and DNA modified with P9-A1 after ligation with

Alexa Fluor 488-alkyne (red trace).

89

Figure 3. 4. Imaging of P1-A9 (5 µM) in fixed NCI-H460 lung cancer cells using Alexa

Fluor 488-alkyne. Cells were co-stained with Hoechst 33342. (A) Single confocal image

planes of Alexa Fluor 488 (green), Hoechst 33342 (blue), the green and blue channels

merged, and corresponding bright field image are shown for P1-A9 treatment and control

conditions. Control groups were treated with Alexa Fluor 488-alkyne and copper-based

ligation buffer. No-copper controls were treated with Alexa Fluor 488-alkyne in the

absence of copper. (B) Single confocal image plane of the cellular distribution of

platinum–azide-related fluorescence in NCI-H460 cells during mitosis (labeled a) and

interphase (example labeled b).

90

Next, we tested if the labeling strategy developed in cell-free DNA could be applied to

whole cancer cells. For confocal fluorescence microscopy studies, NCI-H460 cells were

incubated with P1-A9 under conditions that were expected to produce high intracellular

levels of platinum drug without killing the cells, as confirmed by the absence of the

morphological hallmarks of apoptotic cell death [302]. Briefly, cultured cells were treated

with three different doses (1, 5, or 25 μm) of P1-A9 for three hours, washed, fixed in

formalin, permeabilized, and incubated with CuAAC reaction buffer in the presence of

Alexa Fluor 488-alkyne. Control experiments without platinum drugs were treated with

fluorescent dye in the presence or absence of copper ions. Following exhaustive washing

to remove excess labeling mixture and co-staining with a nuclear dye (Hoechst 33342),

cells were imaged by confocal fluorescence microscopy.

Images of platinum-treated NCI-H460 cells incubated with Alexa Fluor 488-alkyne and

copper ions clearly demonstrate that this labeling method can be applied to whole cells

(Figure 3.4 A). Cells treated with P1-A9 show a pronounced intensity increase in

fluorescence emission relative to condition-matched controls, as a result of selective

intracellular azide–alkyne ligation chemistry. Alexa Fluor 488 emission intensities in the

nuclei of the treated cells (n > 50) show a clear correlation with increasing incubation

concentrations of P1-A9, confirming that intracellular reaction with the azide-modified

platinum/DNA adduct is the source of the fluorescence (see Appendix D).

91

Fig

ure

3.5

. D

eter

min

atio

n o

f in

trac

ellu

lar

dis

trib

uti

on o

f P

1-A

9 i

n a

NC

I-H

460 c

ell

in i

nte

rphas

e b

y c

o-s

tain

ing w

ith n

ucl

ear

(Hoec

hst

33342)

and m

itoch

ondri

al (

Mit

oT

rack

er D

eep R

ed)

dyes

. T

he

nucl

eolu

s, w

hic

h i

s re

adil

y i

den

tifi

ed i

n t

he

bri

ght-

fiel

d

imag

e an

d a

s a

nu

clea

r re

gio

n u

nst

ained

by H

oec

hst

dye,

appea

rs a

s an

are

a of

hig

h f

luo

resc

ent

inte

nsi

ty i

n t

he

Ale

xa

Flu

or

488

and m

erged

chan

nel

im

ages

.

92

Platinum-azide-associated fluorescence was observed at high intensity within the nuclei

and at a much lower intensity within the cytoplasm (for a detailed analysis of the distinct

cellular regions, see the Appendix D). Figure 3.4 B shows co-localization images of

Alexa Fluor 488 and the nuclear stain Hoechst 33342 within NCI-H460 cells during

different stages of the cell cycle. During mitosis, the highest levels of green fluorescence

were observed in the condensed, replicated chromatin, whereas cells in interphase show

the strongest fluorescence signal across the entire nucleus. The highest fluorescence

intensity in the nucleus during interphase was observed in the nucleoli, which are the

sites of ribosomal RNA (rRNA) transcription during interphase [303]. An NCI-H460 cell

imaged at high magnification shows the relative distribution of Alexa Fluor within the

cell and confirms that P1-A9 accumulates in the nucleus and nucleolus (Figure 3.4).

Analysis of a large number of cells (n = 40–50) shows that the emission in the 488 nm

excitation channel has 50 % higher relative intensity within the nucleolar regions

compared to the surrounding chromatin (see Appendix D, Figure D.3).

3.2.2. Post-labeling using copper-free click chemistry

To investigate if the detection of intracellular platinum–acridine can also be achieved

by copper-free click chemistry, similar labeling experiments were performed with Alexa

Fluor 488-DIBO, which contains an azide-reactive dibenzocyclooctyne group (Figure.

3.6B) [304]. The confocal images obtained from these experiments (Figure. 3.7) show

unfavorable high background fluorescence in all cytosolic organelles of cells not treated

with compound P1-A9. In addition, the DIBO-based fluorophore, unlike simple

Cu/alkyne, was unable to label platinum adducts in the DNA of dividing cells. This

93

observation suggests that the steric bulk of DIBO may prohibit reaction with platinum-

based azide in the compacted chromatin. Thus, copper-free click chemistry using DIBO-

modified dye was dismissed as a post-labeling technique in experiments requiring

orthogonal reactions with two fluorophores.

Figure 3.6. (A) Structures of probes used in this study. (B) Detection of azide-modified

platinum–acridine (DNA adducts) with complementary alkyne-modified fluorescent dye.

The star indicates Alexa Fluor 488 (green) and Alexa Fluor 647 (red) fluorophore. (C)

Simultaneous detection of platinum–acridine in cells incubated with a mixture of

compound P1-A9 and its Alexa Fluor 488-DIBO conjugate, P1-A9*. Post-labeling of P1-

A9 was performed with the alkyne form of Alexa Fluor 647 (red star).

94

Fig

ure

3.7

. C

onfo

cal

imag

es o

f untr

eate

d N

CI-

H460 l

ung c

ance

r ce

lls

(co

ntr

ol)

and c

ells

tre

ated

wit

h c

om

pound P

1-A

9 (

5 µ

M).

Both

pla

tinum

-tre

ated

and c

ontr

ol

cell

s w

ere

trea

ted w

ith A

lex

a F

luor

488

-DIB

O (

gre

en).

Hoec

hst

33342 (

blu

e),

mer

ged

, an

d

bri

ght

fiel

d ch

ann

el im

ages

ar

e al

so sh

ow

n.

T

he

arro

ws

indic

ate

com

pac

ted ch

rom

atin

in

div

idin

g ce

lls,

w

hic

h is

st

ained

effi

cien

tly b

y n

ucl

ear

dye

but

not

by t

he

DIB

O-c

onta

inin

g f

luoro

phore

. S

cale

bar

= 1

0 µ

M.

95

3.2.3. Pre-labeling method

In section 3.2.1, we have developed and validated the post-labeling technique using

quantitative sampling of fluorescence intensities in treated cells and appropriate

condition-matched controls. We have also used this technique to compare the subcellular

localization of post-labeled platinum with that of fluorescently (pre-)labeled compound

P1-A9* ( Figure 3.6A) in NCI-H460 lung cancer cells during interphase and mitosis. We

have also combined the detection of platinum with a second orthogonal click reaction.

These experiments represent the first successful case of orthogonal post-labeling

chemistry in whole cells involving a platinum-based pharmacophore.

To determine the effects of tagging the azide-modified complex P1-A9 with a

fluorophore prior to treatment of cancer cells, we attempted to label this compound with

Alexa Fluor-alkyne dye using CuAAC chemistry. Unfortunately, these reactions, which

had to be performed in (partly) aqueous media, led to substantial decomposition of

compound P1-A9 and products indicating aquation of complex and undesired substitution

of platinum-coordinated chloride by the alkyne moiety (data not shown). By contrast,

Alexa Fluor 488-DIBO reacted cleanly with compound P1-A9 in DMF solution to

produce the desired conjugate P1-A9* (Figure. 3.6A). When the cyclooctyne-modified

dye was reacted with excess platinum (1:10) at room temperature, complete conversion to

cycloadduct was observed within 3 h, as confirmed by in-line LC–MS analysis (Figure

3.8). Reaction mixtures generated this way, in which no unreacted Alexa Fluor 488-

DIBO was detectable, were directly used for colocalization studies in NCI-H460 cells.

Because conjugate P1-A9* was generated in situ on a microgram scale, it was not

96

possible to isolate sufficient amounts of the pure conjugate for imaging and cell

proliferation assays.

Cells exposed to a mixture of P1-A9 and P1-A9* (~9:1) were imaged after

permeabilization and treatment with Alexa Fluor 647-alkyne in the absence or presence

of copper. The former conditions were chosen to selectively detect green-fluorescent P1-

A9* and as a negative control of non-platinum associated Alexa Fluor-647 (red)

fluorescence. Copper-based click chemistry with the red-fluorescent dye was used to

detect Pt-azide (P1-A9) and P1-A9* simultaneously. Confocal images taken under both

conditions show the highest level of green fluorescence in the nucleolar regions of the

cells, confirming that the fluorophore-labeled platinum–acridine (P1-A9*) accumulates

in this subnuclear structure (Figure 3.9). This is an important observation, since Alexa

Fluor 488-DIBO itself shows no significant affinity for the nucleoli in cells in interphase

(see controls for DIBO post-labeling experiments in APPENDIX D). By comparison,

nuclear and cytosolic levels of P1-A9* are considerably lower. As expected, under

copper-free conditions only insignificant background levels of red fluorescence are

observed in the Alexa Fluor 647 channel (Figure 3.9, top panel). When copper is

introduced for the purpose of post-labeling intracellular compound P1-A9, high levels of

red fluorescence are observed across the entire nuclear region of the cells (Figure 3.9,

bottom panel). This observation is consistent with the previously observed (Figure 3.4)

accumulation of the hybrid agent in the nuclear DNA of both cells in interphase and

mitosis. A high degree of co-localization is observed for compounds P1-A9 and P1-A9*

in the merged-channel image, suggesting that labeled and unlabeled platinum–acridine

share qualitatively similar subcellular distribution patterns. However, closer inspection

97

of individual cells reveals that the fluorescence level in the chromatin (relative to that

observed in the nucleolus and cytosol) is significantly higher in the red channel than in

the green channel (see Figure D.4 in APPENDIX D for an enlarged image of a single

cell). A possible explanation for this effect would be the increased steric bulk in P1-A9*

introduced by the Alexa Fluor-DIBO moiety, which can be expected to affect the ability

of the conjugate to associate and form adducts with double-stranded DNA of the

chromatin.

98

Figure 3.8. Labeling of compound P1-A9 with Alexa Fluor 488-DIBO in DMF solution

monitored by in-line LC-MS. HPLC traces are shown for an Alexa Fluor-specific (λmax =

488 nm) and an acridine-specific (λmax = 413 nm) wavelength range. HPLC fractions and

the corresponding negative-ion electrospray mass spectra are labeled ‘1’ and ‘2’ for the

two constitutional isomers of cycloadduct P1-A9* and ‘3’ for compound P1-A9,

respectively. Molecular ions ([M–nH]-) are observed for each species at m/z values of

1556, 1557, and 721, respectively. Note the absence of unreacted Alexa Fluor 488-DIBO

under these reaction conditions (10-fold excess compound P1-A9).

99

Fig

ure

3.9

. C

olo

cali

zati

on s

tud

y o

f pla

tinum

–ac

ridin

e P

1-A

9,

post

-lab

eled

wit

h A

lex

a F

luor

64

7-a

lkyn

e (r

ed c

han

nel

), a

nd i

ts

Ale

xa

Flu

or

488

-DIB

O-l

abel

ed d

eriv

ativ

e, P

1-A

9*,

in N

CI-

H46

0 c

ells

. C

ells

wer

e co

-sta

ined

wit

h H

oec

hst

33342 n

ucl

ear

dye

(blu

e ch

ann

el).

N

ote

the

hig

hes

t deg

ree

of

colo

cali

zati

on o

f th

e tw

o s

pec

ies

in t

he

nucl

eola

r re

gio

ns

of

post

-lab

eled

cel

ls,

whic

h

appea

r as

bri

ght-

yel

low

spots

in t

he

mer

ged

-ch

ann

el i

mag

e. S

cale

bar

s re

pre

sent

a le

ngth

of

10 µ

m.

100

3.3. Conclusions

In conclusion, the current study demonstrates that copper-mediated click chemistry

based post-labeling strategy is a powerful method for mapping the subcellular

localization of a platinum-containing pharmacophore in fixed cancer cells. This was

achieved by modifying a potent platinum-acridine anticancer agent so that a fluorophore

could be ligated to it within cells. The post-labeling approach is likely to truly reflect the

subcellular distribution of platinum-acridines due to less steric hindrance may involve in

the technology, which outperformed other strategies in this study, including post-labeling

with copper-free click chemistry and pre-assembly method. P1-A9 represents the first

example of a platinum-based anticancer agent compatible with alkyne–azide ligation

chemistry. The rapid accumulation of P1-A9 in chromatin in the nucleus after only three

hours of incubation is consistent with the high levels of platinum-DNA adducts seen

previously in NCI-H460 cells by ICP-MS [275]. The imaging results support the notion

that damage to nuclear DNA by platinum drugs is a major trigger of S-phase arrest and

apoptosis in NCI-H460 cells [296, 305]. The biological consequences of platinum-

acridine localization to the nucleoli, on the other hand, remain elusive. Potential targets

for platinum adduct formation in this dynamic, non-membrane-bound structure include

transcriptionally active rRNA genes (rDNA) and nascent rRNA itself [303]. Nucleolar

rRNA synthesis is an emerging target in anticancer drug development, and certain DNA

intercalators have shown selective inhibition of rRNA transcription [306, 307]. Notably,

in deletion strains of S. cerevisiae, platinum-acridines elicit responses from genes

involved in RNA metabolism, which suggests that inhibition of rRNA synthesis may

contribute to the cell death produced by these agents [296]. Finally, we anticipate that this

101

new imaging method could be extended to more biocompatible and ultrafast copper-free

click chemistry, which could monitor platinum drug levels and localization in real-time

within live cells [308].

3.4 Experimental Section

3.4.1. Materials and Methods.

The synthetic precursors 3-2 and 3-4 were synthesized according to the methods

described previously. All reagents were used as obtained from commercial sources

without further purification unless indicated otherwise. 1H NMR spectra of the target

compounds and intermediates were recorded on a Bruker Advance 300 MHz instrument.

Chemical shifts (δ) are given in parts per million (ppm) relative to internal standard

tetramethylsilane (TMS). 1H NMR data is reported in the conventional form including

chemical shift (δ, ppm), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m =

multiplet, br = broad), coupling constants (Hz), and signal integrations. 13C{1H} NMR

data are reported as chemical shift listings (δ, ppm). The NMR spectra were processed

and analyzed using the MestReNova software package. HPLC-grade solvents were used

for all HPLC and mass spectrometry experiments. LC-ESMS analysis was performed on

an Agilent 1100LC/MSD ion trap mass spectrometer equipped with an atmospheric

pressure electrospray ionization source. Eluent nebulization was achieved with a N2

pressure of 50 psi and solvent evaporation was assisted by a flow of N2 drying gas

(350 °C). Positive-ion mass spectra were recorded with a capillary voltage of 2800 V

and a mass-to-charge scan range of 150 to 2200 m/z. To establish the purity of target

102

compounds, samples were diluted in methanol containing 0.1 % formic acid and

separated using a 4.6 mm x 150 mm reverse-phase Agilent ZORBAX SB-C18 (5 µm)

analytical column at 25 °C. The purities of compound S3 and compound S5 were

assessed using the following solvent system: solvent A, optima water, and solvent B,

methanol/0.1% formic acid, with a flow rate of 0.5 mL/min and a gradient of 95% A to

5%. A over 15 minutes. The purity of compound P1-A9 was assessed using the

following solvent system: solvent A, optima water, and solvent B, methanol/0.1% formic

acid, with a flow rate of 0.5 mL/min and a gradient of 95% A to 5% A over 30 minutes.

HPLC traces were recorded with a monitoring wavelength range of 363-463 nm. Peak

integration was done using the LC/MSD Trap Control 4.0 data analysis software.

Scheme 3.1. Synthesis of precursor 3-3 and 3-5.

3.4.2. Synthetic procedures and product characterization

The synthesis of N1-(acridin-9-yl)-N2-(2-azidoethyl)ethane-1,2-diamine (3-3). A

mixture of phenoxyacridine (3-1) (2.71 g, 0.01 mol) and N1-(2-azidoethyl) ethane-1,2-

diamine (3-2) (1.42 g, 0.011 mol) in 25 mL of anhydrous THF was refluxed for 16 h. The

solvent was evaporated off and the residue was dissolved in 30 mL of acetone. To this

103

solution were added 5 mL of conc. HCl, and the mixture was stirred at 4 ˚C for 3 hours.

A yellow precipitate formed which was recovered by filtration, resuspended in 50 mL of

2 M ammonium hydroxide, and stirred at room temperature for 30 min. The aqueous

phase was extracted with CH2Cl2 and the organic phase was collected, dried over Na2SO4,

and concentrated using rotary evaporation, affording 2.57 g of the free base (3-3) as an

yellow solid (Yield: 87%). 1H NMR (300 MHz, CDCl3) δ = 8.16 (d, J = 7.2 Hz, 2H),

8.06 (d, J = 8.4 Hz, 2H), 7.65 (t, J = 7.8 Hz, 2H), 7.35 (t, J = 8.1 Hz, 2H), 6.24 (br s, 1H),

3.82 (t, J = 5.6 Hz, 2H), 3.46 (t, J = 5.6 Hz, 2H), 2.93 (t, J = 6.4 Hz, 2H), 2.87 (t, J = 6.3

Hz, 2H), 1.39 (br s, 1H). 13C NMR (75 MHz, CDCl3) δ 151.65, 149.51, 129.80, 122.97,

117.11, 51.57, 49.19, 49.09, 47.91. MS (ESI, positive-ion mode): calculated for

C17H18N6 ([M+H]+), 306.16; found: 307.2.

The synthesis of N1-(acridin-9-yl)-N2-(3-azidopropyl)ethane-1,2-diamine (3-5). 3-5

was prepared using the procedure described for 3-3 starting from phenoxyacridine (3-1)

(2.71 g, 0.01 mol) and N1-(3-azidopropyl)ethane-1,2-diamine (3-4). 1H NMR (300 MHz,

CDCl3) δ 8.15 (d, J = 8.7 Hz, 2H), 8. 06 (d, J = 8.9 Hz, 2H), 7.65 (t, J = 8.5 Hz, 2H),

7.34 (d, J = 8.4 Hz, 2H), 6.26 (brs, 1H), 3.81 (t, J = 5.4 Hz, 2H), 3.40 (t, J = 6.6, 2H),

2.90 (t, J = 5.7 Hz, 2H), 2.75 (t, J = 6.8, 2H), 1.80 (p, J = 6.7, 2H), 1.12 (brs, 1H). 13C

NMR (75 MHz, CDCl3) δ 151.70, 149.47, 129.82, 129.78, 123.00, 122.84, 116.87,

49.50, 49.43, 49.09, 46.28, 29.42. MS (ESI, positive-ion mode): calculated for C18H20N6

([M+H]+), 320.17; found: 321.2.

The synthesis of [PtCl(NH3)2C23H28N8O](NO3)2 (P1-A9). Platinum complex A1 (381

mg, 1 mmol) was converted to its nitrate salt by reaction with AgNO3 (162 mg, 0.95

mmol) in 7 mL of anhydrous DMF. AgCl was removed by syringe filtration, and the

104

filtrate was cooled to -10 °C. Acridine precursor A9 (378 mg, 1 mmol) was added to the

solution, and the suspension was stirred at -10 °C for 24 h. After treatment with activated

carbon, the reaction mixture was added into 300 mL of diethyl ether. The yellow slurry

was stirred for 30 min, then the precipitate was recovered by membrane filtration and

dried in a vacuum overnight. The solid was dissolved in anhydrous methanol containing

one equivalent of 1 M HNO3, stirred at room temperature for 30 minutes and precipitated

with 300 mL of anhydrous diethyl ether. The product was further purified by

recrystallization from hot ethanol to give 432 mg of a yellow microcrystalline solid

(Yield: 51%). 1H NMR (300 MHz, MeOH-d4) δ 8.45 - 8.39 (m, 2H), 7.89 (ddd, J=8.3,

6.9, 1.2 Hz, 2H), 7.78 - 7.70 (m, 2H), 7.52 (ddd, J=8.8, 6.9, 1.3 Hz, 1H), 5.37 - 5.15 (m,

4H), 4.27 (t, J=6.8 Hz, 2H), 3.84 (t, J=6.8 Hz, 2H), 3.59 (t, J=6.6 Hz, 2H), 3.30 - 3.22 (m,

4H), 2.94 (q, J=7.6, 2H), 2.58 - 2.26 (m, 7H), 1.18 (t, J=7.6, 3H). 13C NMR (75 MHz,

MeOH-d4) δ 173.27, 171.13, 160.17, 141.38, 136.67, 125.45, 119.80, 114.17, 51.38,

49.86, 39.96, 29.19, 11.99.

3.4.3. In-gel Alexa Fluor 488 fluorescence detection

Plasmid DNA (pUC19, 2686 bp) was generated by transforming a chemically

competent E. coli (NovaBlue, Novagen) with pUC19 and extracted from cells using

Qiagen’s MegaPrep Plasmid kit (Qiagen, Valencia, CA). The plasmid was linearized

with restriction endonuclease BamH I (New England BioLabs, Ipswich, MA) at 37 °C

overnight. A small aliquot was removed and analyzed on an agarose gel to confirm

complete conversion to the linear form. After heat-inactivation of enzyme at 65 °C, the

DNA was purified with the QIAquick PCR Purification kit (Qiagen, Valencia, CA).

105

DNA concentrations and purity were determined spectrophotometrically at 260 nm with

= 6500 M1 cm1 (nucleotides, n.t.). The linearized pUC19 was incubated with

compound P1-A9 at drug-to-nucleotide ratios (rb) of 0.2, 0.1, and 0.05 in 5 µL of 10 mM

PBS at 37 °C for 24 hours. To the samples were then added 45 µL of click reaction buffer

(CuSO4, TBTA (tris-[(1-benzyl-1H-1,2,3-triazol-4-yl)methyl]amine), Alexa Fluor 488-

alkyne (Invitrogen, Carlsbad, CA) and sodium ascorbate in 10 mM PBS, pH 7.4) to

produce the final concentrations listed in Table 3.1. These mixtures were incubated for 3

hours at 37 °C in the dark. To remove the low-molecular-weight compounds in the click

reaction mixture, the clicked plasmid DNA were purified by using the MinElute PCR

purification columns (Qiagen, Valencia, CA). To further support that the fluorescent dye

was ligated with to platinum complexes covalently bound to the DNA, 3 mM NaCN was

incubated with the purified Alexa Fluor 488-modified plasmid at 37 °C for 12 hours to

remove the platinum complex from DNA, followed by purification on MinElute PCR

purification column. The samples and control plasmid were analyzed by agarose gel (1 %)

electrophoresis and visualized through in-gel fluorescence using a Typhoon Trio Variable

Mode imager (GE Healthcare, Piscataway, NJ). Images were collected with a 488 nm

excitation laser and a 520 bp40 emission filter at 600 V PMT. The same gels were further

stained with ethidium bromide and documented using the Kodak Electrophoresis

Documentation and Analysis System 290 (Rochester, NY).

106

Tab

le 3

.1. C

om

posi

tion o

f C

uA

AC

Rea

ctio

ns

An

alyze

d b

y G

el E

lect

ropho

resi

s

Lan

e r b

P

uc1

9

(nucl

eoti

des

, µ

M)

P1-A

9

(µM

)

CuS

O4

(µM

)

TB

TA

(µM

)

Sodiu

m A

scorb

ate

(µM

)

Ale

xa

Flu

or

488

(µM

)

NaC

N

(mM

)

1

0

1.8

2

0

0

0

0

0

0

2

0

1.8

2

0

0.1

82

0.1

82

9.1

0.9

1

0

3

0.0

5

1.8

2

0.0

91

0.0

91

0.0

91

4.5

5

0.4

55

0

4

0.1

1.8

2

0.1

82

0.1

82

0.1

82

9.1

0.9

1

0

5

0.2

1.8

2

0.3

64

0.3

64

0.3

64

18.2

1.8

2

0

6

0.1

1.8

2

0.1

82

0.1

82

0.1

82

9.1

0.9

1

3

107

3.4.4. Stability of Compound P1-A9 in buffers

10 mM solutions of compound P1-A9 were prepared in the following buffers: 10 mM

phosphate-buffered saline (PBS), pH 7.4; 10 mM phosphate (PB), pH 7.4; acetate, pH 5.0.

Each sample was heated at 37 °C in the dark for 24 h and then analyzed by LC-ESMS.

The chromatographic separations were performed with a 4.6 mm x 150 mm reverse-

phase Agilent ZORBAX SB-C18 (5 µm) analytical column with the column temperature

maintained at 25 °C and a binary mobile phase system consisting of: solvent A, optima

water, and solvent B, methanol/0.1% formic acid, a gradient of 95% A to 5% A over 30

minutes and a flow rate of 0.5 mL/min.

3.4.5. Plasmid Unwinding Experiments.

Unwinding experiments were performed according to a previously published procedure

[309].

3.4.6. CD experiments

Solutions of calf thymus DNA (lyophilized powder, Sigma, St. Louis, MO) were

prepared in 10 mM phosphate buffer (pH 7.4). The DNA samples were annealed by slow

cooling from 90 ˚C to room temperature, and the final concentration was determined

spectrophotometrically at 260 nm with ε = 6500 M−1·cm−1 (nucleotides, n.t.).

To modify the DNA with platinum agent, calf thymus DNA (1 mM, n.t.) was incubated

in phosphate buffer (10 mM, pH 7.4) overnight at 37 ºC with P1-A1 and P1-A9 at a

platinum-to-nucleotide ratio (rb) of 0.05. To perform the click reaction, calf thymus DNA

108

(1 mM n.t.) was incubated with platinum complex P1-A9 with rb 0.05 at 37 ºC overnight.

The sample was then incubated with click reaction buffer at 37 °C for 2 hour with the

following final concentrations: CuSO4 (50 µM), TBTA (50 µM), Alexa Fluor 488 Alkyne

(50 µM) and sodium ascorbate (500 µM)).

CD experiments were performed on an AVIV Model 215 spectrophotometer equipped

with a thermoelectrically-controlled cell-holder. All the experiments were performed in

quartz cells with 0.5 cm path length. Spectra were recorded in the range 200-600 nm at

25 °C in 1-nm increments with an averaging time of 1 s. CD profiles were base-line

adjusted by subtracting buffer background.

3.4.7. Cell based assays based on post-labeling method

3.4.7.1. Cell culture maintenance

The human non-small cell lung cancer cell line, NCI-H460, was obtained from the

American Type Culture Collection (Rockville, MD, USA) and was cultured in RPMI-

1640 media (HyClone) containing 4.5 gL1glucose, 1.5 gL1 sodium bicarbonate, 10

mM HEPES, and 110 mgL1sodium pyruvate supplemented with 10% fetal bovine

serum (FBS), 10% penstrep (P&S), and 10% L-glutamine. Cells were incubated at a

constant temperature at 37 °C in a humidified atmosphere containing 5% CO2 and were

subcultured every 2 to 3 days in order to maintain cells in logarithmic growth.

109

3.4.7.2. Cell treatment and Cu(I)-catalyzed azide-alkyne cycloaddition (CuAAC)

ligation

NCI-H460 cells were seeded into poly-D-lysine coated glass bottom Petri dishes

(MatTeck Corporation, Ashland, MD, USA) with 105 cellsmL1 suspended in 2 mL of

medium per dish. Cells were incubated overnight and then treated with compound P1-A9

(5 μM) (or media for controls) for 3 h. After rinsing with chilled PBS (3), cells were

fixed by treatment with 3.7% formaldehyde solution (in PBS, pH 7.4) for 15 min at room

temperature and subsequently washed with a 3% solution of bovine serum albumin (BSA;

Sigma) in PBS (pH 7.4) twice for 10 min. For colocalization experiments using

mitochondrial stain, cells were treated with 100 nM MitoTracker Deep Red TM

(Invitrogen) in 2 mL pre-warmed phenol red-free media at 37 °C for 30 min prior to

fixing them using the abovementioned procedure. Cells were then permeabilized by

treatment with 0.5% Triton X-100 (in PBS, pH 7.4) at room temperature for 20 min.

Permeabilization buffer was quenched with 3% BSA in PBS (2 10 min) and the cells

were incubated with 250 μL of click reaction mixture (1 mM CuSO4; 0.5 μM Alexa Fluor

488-alkyne (Invitrogen); 10 mM sodium ascorbate; 50 mM Tris-HCl, pH 7.3) at room

temperature for 30 min. For the no-copper control, the copper solution in the click buffer

was replaced with water. Before co-staining with organelle-specific dyes, cells were

subjected to extensive washes with gentle agitation: 1) 3% BSA in PBS (5 min); 2) 0.5%

Triton X-100 in PBS (2 10 min); 3) PBS for (3 10 min). Nuclei were stained by 5

μgmL1 Hoechst 33342 (Sigma) in PBS for 5 min. Three final PBS washes were

performed immediately prior to image capture.

110

3.4.7.3. Confocal microscopy and fluorescence intensity analysis

Images were collected using a Zeiss LSM 710 confocal microscope (Carl Zeiss

MicroImaging, Thornwood, NY) using either a 40x (PLAN APO, 0.95 NA) or a 63x

(PLAN APO, 1.2 NA) objective lens. All images were acquired in multi-track

configuration mode to minimize excitation cross talk and emission bleed-through. We

utilized a 405 nm laser line (for Hoechst 33342) with an emission range of 424466 nm,

a 488 nm laser line (for Alexa Fluor 488) with an emission range of 489553 nm, and a

633 nm laser line (for MitoTracker Deep Red TM) with an emission range of 645740

nm. For comparative fluorescence intensity analysis, great care was taken to equalize

excitation power, pinhole settings, PMT gain, and offset values across and within

imaging sessions for each respective channel. For all images, the pinhole value was kept

at or below 1.2 airy units, and images were acquired with 8x line averaging at 1024 x

1024 pixels. Images were collected at 12-bit sampling to provide a wide dynamic

intensity range for analysis (0–4096). Zen software was used for image acquisition.

The intensity of platinum–azide-related Alexa Fluor 488 fluorescence in nuclei and

nucleoli was compared quantitatively using Zen software. The fluorescence signals from

the nuclei of labeled cells were measured by drawing a region of interest (ROI) around

each nucleus, with the Hoechst 33342 image used to verify nucleus location. The

fluorescence intensity of each object, Fobj, was calculated using the following formula:

Fobj = IROI Ibkgd,

111

where IROI and Ibkgd are the averaged signal intensities from all the pixels in the ROI and

background, respectively. Background was determined by drawing an ROI of the same

size in an empty region of the image. The fluorescence intensity in the nucleoli was

determined in the same manner. A total of 40–50 cells in 8 different fields were analyzed

in this manner for each treatment (cell counts were analyzed by one-way ANOVA,

indicating there was no statistical difference between fields; P = 0.54). The relative

enhancement of Pt-Alexa Fluor 488 fluorescence in the nucleoli was determined from the

ratio of intensities in nucleoli and in the nuclei for each group. Statistical analysis was

done using one-way ANOVA (GraphPad Prism). P < 0.05 was considered significant.

3.4.8. Cell based assays based on pre-labeling method

3.4.8.1. Labeling of compound P1-A9 with Alexa Fluor 488-DIBO to generate P1-

A9*

The copper-free click reaction was performed by combining solutions of compound

P1-A9 (2.5 µL, 25 mM) and Alexa Fluor 488-DIBO (2.5 µL, 2.5 mM, purchased from

Invitrogen) in DMF, followed by adjustment of the volume to 200 µL, vortexing, and

incubation of the mixture at room temperature for 3 h. The progress of the reaction was

monitored by reverse-phase high-performance liquid chromatography (HPLC) using the

LC module of an Agilent Technologies 1100 LC/MSD trap system equipped with a

multi-wavelength diode-array detector. Analytical separations were performed on a 4.6

mm × 150 mm reverse-phase Agilent ZORBAX SB-C18 (5 μm) column at 25° C using

112

the following conditions: solvent A: optima water/0.1% ammonium formate, solvent B:

methanol/acetonitrile (1:1, v/v), flow rate – 0.5 mL/min, gradient 0-5 min, 95% A; 5-20

min 95% A to 5% A; 20-30 min, 5% A. HPLC traces were recorded over a wavelength

range of 363–463 nm (for compound P1-A9) and 480–496 nm (for labeled P1-A9*).

Electrospray mass spectra (ESMS) of the LC fractions were recorded in negative-ion

mode. All data were processed with the LC/MSD Trap Control 4.0 data analysis

software.

3.4.8.2. Cell culture and incubations

The human non-small cell lung cancer cell line NCI-H460 (American Type Culture

Collection, Rockville, MD) was maintained as described previously. Cells were seeded

into poly-D-lysine coated glass bottom cell culture dishes (MatTeck Corporation,

Ashland, MD) with 105 cells·mL1 suspended in 2 mL of medium per dish. Cells were

incubated overnight and then treated with compound P1-A9 (5 μM, synthesized

according to a published procedure [[310]]), compound P1-A9 in the presence of 10% of

its labeled form (P1-A9*), or medium for controls, at 37 °C for 3 h. After incubation,

cells were washed with PBS (3 2 mL). Finally, cells were fixed by treatment with 3.7%

formaldehyde solution (in PBS, pH 7.4) for 15 min at room temperature, washed with a

solution of bovine serum albumin (3 % BSA in PBS; Sigma) twice for 10 min, and

permeabilized by treatment with 0.5% Triton X-100 (in PBS, pH 7.4) at room

temperature for 20 min. The reaction was quenched with 3% BSA in PBS (2 10 min)

and the cells were incubated with 250 μL of click reaction mixture (1 mM CuSO4; 0.5

μM Alexa Fluor 488-alkyne or Alexa Fluor 647-alkyne (Invitrogen); 10 mM sodium

113

ascorbate; 50 mM Tris-HCl, pH 7.3) at room temperature for 30 min. Before adding

nuclear stain in colocalization experiments, cells were subjected to extensive washes: 1)

3% BSA in PBS (5 min); 2) 0.5% Triton X-100 in PBS (2 10 min); 3) PBS for (3 10

min). Nuclei were stained with 5 μg·mL1 Hoechst 33342 (Sigma) in PBS (5 min),

followed by three additional PBS washes prior to image acquisition.

3.4.8.3. Confocal microscopy

Images were recorded on a Zeiss LSM 710 confocal microscope (Carl Zeiss,

Thornwood, NY) using either a 40 (PLAN APO, 0.95 NA) or a 63 (PLAN APO, 1.2

NA) objective lens. All images were acquired in multi-track configuration mode to

minimize excitation cross talk and emission bleed-through. A 405 nm laser line with an

emission range of 410–478 nm was used for Hoechst 33342, a 488 nm laser line

(emission: 489–553 nm) for Alexa Fluor 488. For all images, the pinhole value was kept

at or below 1.2 airy units. Images were acquired with 8 line averaging at 1024 1024

pixel resolution and 12-bit sampling to provide a wide dynamic intensity range for

analysis (0–4096). Zen software was used for image acquisition and processing.

Confocal image planes for each channel were not contrast-adjusted or otherwise changed

and assembled in Adobe Photoshop CS without further manipulation.

114

CHAPTER 4

SYNTHESIS AND EVALUATION OF A PLATINUM-ACRIDINE-ENDOXIFEN

CONJUGATE

The work contained in this chapter was initially published in Chemical Communications

in 2013. The manuscript, including figures and schemes, was drafted by Song Ding and

Dr. U. Bierbach and edited by Dr. U. Bierbach before submission to the journal. Since the

publication, changes in both format and content have been made to maintain a consistent

format throughout this work. The experiments described were performed by Song Ding,

except for the the cytotoxicity assays which were contributed by Dr. Xin Qiao.

115

4.1. Introduction and design rationale

Breast cancer continues to be a major cause of cancer-related mortality in women

[311]. Anti-estrogens such as tamoxifen (Figure 4.1) have become the mainstay in the

management of the early-stage hormone-dependent form of the disease [311].

Unfortunately, a significant population of patients does not respond to this drug or

develops resistance to endocrine therapy during the course of treatment, and systemic

chemotherapies often become the only armamentarium to combat recurrent disease [312].

Current platinum-based drugs have shown limited utility as first-line or salvage therapies

and no survival benefits have been reported for regimens based on these agents in

combination with targeted therapies [313]. The general lack of responsiveness of breast

cancers to platinum has been attributed to unusually high levels of acquired resistance to

this treatment including detoxification by elevated levels of cytosolic glutathione and

repair of DNA adducts [313]. One common strategy to overcome several of these

problems and provide more effective drugs for hormone-dependent breast cancer is to

combine an estrogen receptor (ER)-binding molecule with a classical cytotoxic agent.

The rationale behind this approach is to improve the pharmacological properties of

common cytotoxics by taking advantage of the targeting, antiestrogenic, and

chemosensitizing effects of ER-interacting ligands [314]. Various conjugates containing

ER-affinic carriers have been designed with the goal of targeting ER-positive breast

cancers more efficiently with the drug cisplatin [258, 315-317] and other cytotoxic agents

[318-321].

While several of the cis-platinum-based compounds have shown promising activity in

vitro and in mouse models [317], no attempts of extending this concept to the newer

116

nonclassical platinum anticancer agents have been reported. One such class of

compounds are platinum–acridine agents (Figure 4.1), which, unlike cisplatin, do not

induce cross-links but target DNA through a dual binding mode involving intercalation

and monofunctional platination of nucleobase nitrogen [289]. In non-small cell lung

cancer (NSCLC), the hybrid agents produce more severe DNA damage than cisplatin and

are significantly more cytotoxic than the clinical agent [275, 277, 301, 305]. Given the

superior cell kill of these agents and their ability to circumvent resistance to cisplatin, we

embarked on generating a breast cancer targeted agent containing our platinum–acridine

hybrid as a cytotoxic component. Here, we report on the design of the first conjugates

featuring a suitably functionalized platinum–acridine warhead coupled to an active

metabolite of the ER modulator and anticancer drug tamoxifen. The synthetic

methodology developed to generate these molecules is presented as well as screening

results in breast cancer cells.

Figure 4.1. Structures of tamoxifen and platinum–acridine hybrids.

117

4.2. Results and Discussion

4.2.1. Efforts to synthesize platinum-acridine-endoxifen conjugates

The strategy for generating the conjugates was to modify our platinum–acridine at the

N1 position (Figure 4.1) with a hydroxyl-functionalized side chain and attach it to the

secondary amino group of (E/Z)-4-hydroxy-N-desmethyltamoxifen [(E/Z)-endoxifen, 4-3,

Figure 4.2], the monohydroxylated/N-demethylated metabolite of tamoxifen [322].

Molecular models (not shown) were used to confirm that linking the two pharmacophores

at this position does not interfere with the binding of the components to their

pharmacological targets [274, 323]. The 4-hydroxo form of the pharmacophore was

chosen because of its superior binding affinity for ER and exquisite potency as an

inhibitor of hormone-dependent breast cancer cell proliferation [322] (confirmed in this

study). In addition, since the E- and Z-isomers of 4-hydroxytamoxifen (but not those of

tamoxifen) interconvert under physiologically relevant conditions [324], it was not

necessary to generate the stereochemically pure forms of the target compounds.

First, we designed the conjugatable platinum–acridine (P4-A3, Figure 4.2) based on the

potent derivatives identified previously (Figure 4.1). The target cytotoxic warhead

features simple ammine (NH3) nonleaving groups to maximize the aqueous solubility

[289] of the conjugate and an acetamidine donor group (R = Me), which has proven to

produce a 3-4 times more cytotoxic hybrid than propionamidine (R = Et) [277]. The

synthesis of P4-A3 involved addition of the linker secondary amino group in 2-((2-

(acridin-9- ylamino)ethyl)amino)ethanol (A3) and its ethylene glycol extended derivative

(A11) to the nitrile ligand in [PtCl(NH3)2(MeCN)]NO3 (Figure 4.2). Attempts to directly

118

attach the endoxifen moiety to the hydroxyl group in P4-A3 using carbamate coupling

chemistry failed because the reaction conditions were incompatible with the metal-

containing fragment. Instead, it was necessary to introduce platinum as the last step of

the reaction sequence after preassembling the entire organic scaffold. This was achieved

by reacting N-Boc-protected and 4-nitrophenyl carbonate-activated acridine–amines 4-1

and 4-2 with endoxifen (4-3) to generate the carbamate-coupled conjugates A12 and A13,

followed by addition of platinum nitrile complex to install the amidine-linked platinum

moiety (Figure 4.2). Conjugates P4-A12 and P4-A13 were isolated as 1:1 mixtures of the

E- and Z-isomers (for details of the product characterization, see the Experimental section

and Appendix A, Figure A.35-Figure A.37).

The carbamate linkage in the conjugates was chosen for its chemical robustness to

assure stability in circulation and delivery of the intact conjugate to the target site [325].

To test the reactivity of the carbamate group in aqueous media, conjugate P4-A12 was

incubated at 37 ˚C for 48 h in buffers mimicking the environments in serum, cytosol, and

the lysosomes. In-line high-performance liquid chromatography–mass spectrometry

(LC-MS) analysis of the samples confirmed that the carbamate linkage in compound P4-

A12 resists hydrolytic cleavage in all cases (see Appendix B3).

119

Figure 4.2. Synthesis of target molecules. Reagents and conditions: (a) 1. For P4-A3 :

[PtCl(NH3)2(MeCN)]NO3, DMF, -20 ˚C; 2. 1 M HNO3, MeOH. (b) Boc2O, DCM, rt. (c)

4-nitrophenyl chloroformate, TEA, DCM, rt. (d) 1. TEA, DCM, rt; 2. TFA, DCM, rt; 3. 1

M NaOH, DCM, rt. (e) 1. [PtCl(NH3)2(MeCN)]NO3, DMF, 4 ˚C; 2. 1 M HNO3, MeOH.

120

4.2.2. Biological evaluation of target conjugates

To assess the cytotoxic potency of the target conjugates, a colorimetric cell

proliferation assay was performed in the human breast adenocarcinoma cell lines MCF-7,

which overexpresses estrogen receptor (ER+), and the estrogen receptor-negative (ER-)

cell line MDA-MB-231.[326] Compounds P4-A12 and P4-A13 were tested along with

the two precursors, hydroxyl-modified platinum–acridine P4-A3, and (E/Z)-4-hydroxy-

N-desmethyltamoxifen (4-3). For comparison, the clinical drugs tamoxifen and cisplatin

were also included in this assay (Table 4.1). The tamoxifen metabolite, 4-3, showed the

strongest antiproliferative effect in MCF-7 with an IC50 value of 1.6±0.4 μM determined

after exposure of cells to the drug for 72 h. Under the same conditions, a 2.4-fold higher

concentration of compound P4-A3 is required to produce the same cell kill effect in this

cell line. Conjugate P4-A12 shows virtually the same potency as its cytotoxic component,

P4-A3, whereas the extended conjugate P4-A13 was the least active compound in this

assay. All compounds were significantly more cytotoxic in MCF-7 cells than in MDA-

MB-231 cells. For compounds P4-A3, 4-3 and P4-A12, for example, the cell line-

specific cytotoxic enhancements were 3-fold, 8-fold, and 6-fold, respectively. While

compounds P4-A3 and P4-A12 are equipotent in MCF-7, the conjugate proves to be

relatively less cytotoxic in MDA-MB-231. Tamoxifen, the mainstay of breast cancer

treatment, and cisplatin were the least active compounds tested in the ER-positive cell

line.

121

Table 4.1. Summary of cytotoxicity data (IC50, μM)a in human breast cancer cell lines

Compound MCF-7 MDA-MB-231

Tamoxifen 16.2 ± 1.2 20.1 ± 1.0

cisplatin 12.1 ± 1.1 40.6 ± 4.2

P4-A3 3.8 ± 0.4 11.3 ± 0.2

4-3 1.6 ± 0.4 13.2 ± 0.1

P4-A12 3.5 ± 0.1 21.4 ± 1.0

P4-A13 27.5 ± 1.4 46.5 ± 0.7 a The IC50 values are reported as the mean ± standard deviation for two individual 72h

incubations performed in triplicate.

Modification of compound P4-A3 with endoxifen (4-3) produces a conjugate (P4-A12)

that shows essentially the same cytotoxic response (based on IC50 values) as the

unmodified platinum–acridine hybrid in MCF-7 cells. While equipotent, both

compounds can be expected to induce cancer cell kill by different mechanisms. A

comparison of the drug–response profiles for compounds 5 and 10 after 48 h of

incubation with MCF-7 cells confirms this supposition. While the platinum–acridine

agent P4-A3 clearly shows an antiproliferative advantage over compound P4-A12 at low

concentrations, it is unable to kill a major fraction of the cells at the highest

concentrations tested (Figure 4.3). These findings suggest that a subpopulation of cells

may exist that are relatively insensitive to this agent, as has been observed previously for

MCF-7 cells exposed to cytotoxics.[327] Importantly, compound P4-A12 overcomes this

form of resistance, based on the considerably lower percentage of cells that remain viable

after incubation with 20 and 40 µM conjugate for 48 h.

122

Figure 4.3. Differential response of MCF-7 cancer cells to compounds P4-A3 and P4-

A12.

4.3. Conclusion

Here, we have presented the first targeted conjugate based on a nonclassical platinum-

containing hybrid structure. The synthetic approach provides a versatile platform for

incorporating a wide range of small-molecule receptor- and cancer cell-affinic ligands, as

well as enzymatically cleavable and self-immolative linkers.[328] Such modifications

may help target cancers more selectively and reduce the undesired high systemic

toxicity[289] observed for the simple platinum–acridines. Using a hormone-dependent

cancer model, we demonstrated that the ER-targeted form of platinum–acridine maintains

123

the same cytotoxicity as compound P4-A3 in MCF-7 cells. A comparison of the relative

cytotoxic enhancements observed for P4-A3 and P4-A12 in ER+ MCF-7 vs. ER- MDA-

MB-231 (Table 4.1) and the individual drug responses for each derivative suggest that

conjugate P4-A12 acts by a distinct mechanism potentially involving interactions with

ER. Further investigation into the molecular mechanism of the conjugate is necessary to

substantiate this notion. Likewise, it will be important to determine why introduction of

an extended linker in compound P4-A13 results in a greatly diminished cytotoxic

response.

In conclusion, the novel conjugate P4-A12 was generated in an attempt to marry the

cytotoxic, DNA-damaging properties of a nonclassical platinum-acridine agent with the

targeting potential of an antiestrogen. Initial tests show promising cell kill effects, which

warrant further studies into structure–activity relationships in derivatives of the

prototypical agent and their in vivo efficacy against hormone-dependent breast cancer.

Multifunctional agents like compound P4-A12 may have applications in patients who do

not respond to tamoxifen due to altered expression levels of ER or an intrinsic inability to

metabolize tamoxifen [329].

124

Figure 4.4. Synthesis of precursor 4-3. Reagents and conditions: (a) Zn dust, TiCl4,

THF, reflux, 82%. (b) Me2NCH2CH2Cl, Cs2CO3, DMF, 120 °C, 67%. (c) (1)ACE-Cl,

DCM, reflux; (2)MeOH, 6 M HCl, reflux, 81%.

4.4. Experimental Section

4.4.1 Synthesis of target conjugates

The synthesis of 1, 1-Bis(4-hydroxyphenyl)-2-phenylbut-1-ene (4-6). This

compound was synthesized according to the method reported previously[330] with

modifications. To a stirred suspension of zinc powder (8 g, 0.124 mol) in 80 mL of

anhydrous THF, TiCl4 (6 mL, 0.056 mol) was added dropwise under argon at -20 °C.

When the addition was complete, the mixture was warmed to room temperature and

refluxed for 3 h. The reaction was cooled in an ice bath and a solution of 4,4’-

hydroxybenzophenone (4-4, 2 g, 0.0092 mol) and propiophenone (4.0 g, 0.0296 mol) in

10 mL of anhydrous THF was added. The reaction mixture was refluxed in the dark for

125

12 h. The reaction was quenched with 100 mL of 10% aqueous K2CO3 and extracted

with ethyl acetate. The organic layers were collected, washed with saturated NaCl, dried

over MgSO4, and concentrated by rotary evaporation, affording a pink oil. To purify the

crude product, the oil was dissolved in 10 mL of ethanol and precipitated by adding 40

mL of water. The white precipitate was recovered by filtration and recrystallized from 35

mL of hot DCM to afford 4-6 (2.4 g, 82%) as a white solid. 1H NMR (300 MHz,

DMSO-d6) δ 9.35 (s, 1H), 9.10 (s, 1H), 7.21 – 7.04 (m, 5H), 6.97 (d, J = 8.4 Hz, 2H),

6.74 (d, J = 8.5 Hz, 2H), 6.60 (d, J = 8.5 Hz, 2H), 6.39 (d, J = 8.5 Hz, 2H), 2.38 (q, J =

7.4 Hz, 2H), 0.84 (t, J = 7.3 Hz, 3H).

The synthesis of (E,Z)-1-[4-(2-Dimethylaminoethoxy)phenyl]-1-(4-hydrox-

yphenyl)-2-phenylbut-1-ene (4-7). This compound was synthesized according to the

method[330] reported previously with slight modifications. A solution of 1,1-bis(4-

hydroxyphenyl)-2-phenylbut-1-ene (4-6, 1 g, 3.2 mmol) in DMF (15 mL) was mixed

with Cs2CO3 (3.1 g, 9.5 mmol) and refluxed for 30 min. To this mixture was added 2-

(dimethylamino)ethylchloride hydrochloride (0.56 g, 3.8 mmol) in three portions over 1 h.

The mixture was heated at 120 °C for another 6 h. After cooling to room temperature,

the mixture was poured into 20 mL of saturated NH4Cl and extracted with ethyl acetate

(3 × 20 mL). The organic phases were combined, washed with saturated NaCl (2 × 20

mL), dried over anhydrous Na2SO4, and concentrated to give brown oil. The crude

material was purified by recrystallization from 10 mL of hot ethanol to afford 0.82 g

white crystals (4-7) with a 1:1 Z/E ratio (yield: 67%). 1H NMR (300 MHz, DMSO-d6) δ

9.38 (s, 0.5H), 9.13 (s, 0.5H), 7.23 - 7.04 (m, 5H), 6.98 (d, J = 8.4 Hz, 1H), 6.92 (d, J =

8.7 Hz, 1H), 6.75 (d, J = 8.5 Hz, 1H), 6.71 (d, J = 8.7 Hz, 1H), 6.63 - 6.53 (m, 2H), 6.40

126

(d, J = 8.6 Hz, 1H), 4.05 (t, J = 5.8 Hz, 1H), 3.88 (t, J = 5.8 Hz, 1H), 2.62 (t, J = 5.8 Hz,

1H), 2.56 - 2.49 (m, 1H), 2.44 - 2.33 (m, 2H), 2.22 (s, 3H), 2.15 (s, 3H), 0.84 (t, J = 7.3

Hz, 3H).

The synthesis of (E,Z)-Endoxifen Hydrochloride (4-3). This compound was

synthesized according to the method[331] reported previously with modifications. 1-

Chloroethyl chloroformate (ACE-Cl, 668 μL, 61.9 mmol) was added dropwise to a

solution of 4-7 and 1,8-bis(dimethylamino)naphthalene (proton sponge, 249 mg, 11.6

mmol) in 9 mL of DCM at 0 ºC. The resulting solution was stirred for 30 min. at 0 ºC and

then refluxed for 12 h. The reaction mixture was cooled to room temperature and

removal of the solvent under reduced pressure gave a brown crude oil. This residue was

dissolved in a mixture of MeOH (20 mL) and 6 M HCl (5 mL) and refluxed for 3 h and

then cooled to room temperature. The MeOH was evaporated off and the residue was

dissolved in 50 mL of ethyl acetate, washed with saturated NaCl and water, dried over

anhydrous Na2SO4, and concentrated to give 0.260 g of 4-3 as white solid, (yield: 81%),

as 1:1 mixture of Z/E Endoxifen isomers, which was used in the next step without further

purification. 1H NMR (300 MHz, DMSO-d6) δ 9.30 (br s, 2H), 7.25 - 7.06 (m, 5H), 7.00

(d, J = 2.7 Hz, 1H), 6.96 (d, J = 2.7 Hz, 1H), 6.79 - 6.72 (m, 2H), 6.65 (d, J = 8.5 Hz, 1H),

6.60 (d, J = 8.5 Hz, 1H), 6.42 (d, J = 8.5 Hz, 1H), 4.26 (t, J = 4.6 Hz, 1H), 4.08 (d, J =

5.7 Hz, 1H), 3.19 (d, J = 5.7 Hz, 1H), 2.60 (s, 1.5H), 2.54 (s, 1.5H), 2.40 (m, 2H), 0.85 (t,

J = 7.3 Hz, 3H).

127

F

igu

re 4

.5.

Synth

esis

of

pre

curs

or

A11.

Rea

gen

ts a

nd c

ondit

ions:

(a)

K2C

O3,

TH

F,

refl

ux

, 87%

. (

b)

9-p

hen

ox

yac

ridin

e, T

HF

,

refl

ux

, 86%

.

128

Synthesis of 2-(2-(2-((2-aminoethyl)amino)ethoxy)ethoxy)ethanol (4-9). The

mixture of ethylene diamine (3.05g, 0.05 mol) and anhydrous potassium carbonate (1.15

g, 0.0083 mol) in 30 ml anhydrous THF was heated under reflux, to which was added

monotosylated triethylene glycol (4-8, 1.71g, 0.0056 mol) in 15 ml anhydrous THF

dropwisely over 1 h, then the reaction was refluxed for another 16 hours. When the

reaction was completed, the white solids were filtered off after the mixture was cooled to

room temperature. To remove THF and unreacted ethylene diamine, the filtrate was then

concentrated under reduced pressure followed by dried under vacume at 45 ºC overnight

to give a pale yellow crude oil (1.34 g, yield 87%), which can be used in the next step

without any purification. 1H NMR (300 MHz, CDCl3) δ 3.85 - 3.51 (m, 10H), 2.92 -

2.63 (m, 10H), 2.31 (s, 4H). 13C NMR (75 MHz, CDCl3) δ 72.86, 70.46, 70.29, 70.12,

61.30, 52.17, 48.94, 41.46.

Synthesis of 2-(2-(2-((2-(acridin-9-ylamino)ethyl)amino)ethoxy)ethoxy)ethan-1-ol

(A11). A mixture of 9-phenoxyacridine (2.71 g, 0.01 mol) and 2-(2-(2-((2-aminoethyl)

amino) ethoxy) ethoxy) ethanol (4-9, 2.31 g, 0.011 mol) in 30 mL of anhydrous THF was

refluxed for 16 h. The solvent was evaporated off and the residue was dissolved in 30 mL

of acetone. To this solution were added 5 mL of concentrated HCl and the mixture was

stirred at 4 ºC for 3 hours. A yellow precipitate formed which was recovered by filtration,

resuspended in 50 mL of 1 M sodinum hydroxide, and stirred at room temperature for 30

min. The aqueous phase was extracted with CH2Cl2 and the organic phase was collected,

dried over Na2SO4, and concentrated using rotary evaporation, affording 3.18 g of the

free base as an orange oil (Yield: 86%). 1H NMR (300 MHz, CDCl3) δ 8.17 (d, J = 8.8

Hz, 1H), 8.05 (d, J = 8.7 Hz, 1H), 7.69 – 7.59 (m, 1H), 3.84 (t, J = 5.5 Hz, 1H), 3.74 (d, J

129

= 4.7 Hz, 1H), 2.93 (t, J = 5.5 Hz, 1H), 2.84 (t, J = 5.0 Hz, 1H). 13C NMR (75 MHz,

CDCl3) δ 151.98, 149.25, 141.24, 133.06, 129.86, 129.11, 126.91, 123.21, 122.70, 121.31,

120.94, 116.86, 116.77, 72.58, 70.50, 70.34, 70.23, 61.51, 49.32, 49.02, 48.45. MS (ESI,

positive mode): for C26H35N3O5 ([M+H]+), 370.26; found: 370.5.

Synthesis of tert-butyl (2-(acridin-9-ylamino)ethyl)(2-hydroxyethyl)carbamate

(Boc protected A3). The Boc protected A3 was synthesized as follows. 2-((2-(Acridin-

9-ylamino) ethyl)amino)ethanol (A3) (2.82 g, 0.01 mol) was suspended in 30 mL of

anhydrous dichloromethane, to which was added Boc2O (2.18 g, 0.01 mol) in 5 mL of

DCM at 0-5 ºC, maintained with an ice bath. The mixture was then stirred at room

temperature for 4 h. The solvent was removed by rotary evaporation and the residue was

dissolved in 10 mL of DCM and precipitated with 200 mL of anhydrous diethyl ether.

The solid was recovered by filtration and dried in a vacuum, affording Boc protected A3

as yellow solid 3.57 g (yield: 93%), which was used in the next step without purification.

1H NMR (300 MHz, CDCl3) δ 8.08 (d, J = 8.7 Hz, 2H), 7.78 (d, J = 7.8 Hz, 2H), 7.45 (t,

J = 7.8 Hz, 2H), 7.12 (t, J = 7.8 Hz, 2H), 4.15 - 3.95 (m, 2H), 3.78 - 3.54 (m, 4H), 3.42 -

3.24 (m, 2H), 1.43 - 1.10 (m, 9H).

Synthesis of 2-((2-(acridin-9-ylamino)ethyl)amino)ethyl (2-(4-(1-(4-

hydroxyphenyl)-2-phenylbut-1-en-1-yl)phenoxy)ethyl)(methyl)carbamate (A12).

Boc protected A3 (1 g, 3.55 mmol), TEA (1.79 g, 17.7 mmol) and 4-nitrophenyl

chloroformate (0.99 g, 4.6 mmol) were dissolved in 20 mL of anhydrous

dichloromethane. The mixture was stirred at room temperature for 16 h. To the resulting

solution of the 4-nitrophenyl-activated form of the acridine (4-1) was added 4-3 (1.74 g,

4.23 mmol) in 5 mL of anhydrous DCM and the reaction was stirred for another 8 h. The

130

solvent was removed in a high vacuum and the residue was redissolved in 40 mL of

DCM and washed with 1 M HCl (3 × 20 mL). The organic phase was collected, dried

over anhydrous Na2SO4, and concentrated to afford orange oil. To remove the Boc group,

the orange oil was dissolved in 6 mL of a 1:1 mixture of anhydrous DCM and TFA and

stirred at room temperature for 3 h. The reaction was quenched by adding 10 mL of 1 M

NaOH solution. The free base was extracted from NaOH solution with DCM, dried over

anhydrous Na2SO4, and concentrated. The product was purified by flash chromatography

(Al2O3, DCM:MeOH, 30:1) to give 1.35 g of A12 as an orange oil with a 1:1 Z/E ratio

(yield: 56%). 1H NMR (500 MHz, CDCl3, 60 ºC) δ 8.13 - 8.05 (m, 4H), 7.51 (s, 2H), 7.27

- 6.41 (m, 16H), 4.34 - 4.29 (t, J = 5.4 Hz, 1H), 4.28 (t, J = 5.3 Hz, 1H), 4.17 - 3.94 (m,

4H), 3.69 (t, J = 5.4 Hz, 1H), 3.58 (t, J = 5.3 Hz, 1H), 3.11 - 2.94 (m, 7H), 2.55 - 2.40 (m,

2H), 1.00 – 0.84 (m, 3H). MS (ESI, positive mode): for C43H44N4O4 ([M+H]+), 681.3;

found: 681.5.

Compound 2-(2-(2-((2-(acridin-9-ylamino)ethyl)amino)ethoxy)ethoxy)ethyl (2-(4-(1-

(4-hydroxyphenyl)-2-phenylbut-1-en-1-yl)phenoxy)ethyl)(methyl)carbamate (A13) was

prepared according to the procedure described for A12, by using A11 as the precursor

with the yield of 37%. 1H NMR (500 MHz, CDCl3, 60 ºC ) δ 8.38 - 8.01 (m, 3H), 7.77 -

7.24 (m, 3H), 7.22 - 6.98 (m, 6H), 6.86 - 6.25 (m, 6H), 4.33 - 3.39 (m, 16H), 3.07 - 2.79

(m, 6H), 2.62 - 2.37 (m, 2H), 1.79 - 1.66 (m, 2H), 1.31 (s, 1H), 0.95 (m, 3H). 13C NMR

(126 MHz, CDCl3, 25 ºC) δ 156.33, 152.38, 148.61, 141.17, 137.80, 132.09, 132.06,

130.67, 130.67, 130.30, 129.73, 127.86, 125.90, 122.83, 122.83, 116.48, 115.36, 115.28,

114.62, 114.56, 113.96, 113.79, 113.23, 113.04, 70.62, 70.36, 69.58, 48.91, 48.70, 48.58,

131

48.21, 47.91, 35.98, 28.99, 26.40, 13.65, 1.03. MS (ESI, positive mode): for C47H52N4O6

([M+H]+), 769.39; found: 769.8.

Synthesis of conjugate [PtCl(NH3)2C19H22N4O](NO3)2 (P4-A3). Platinum complex

P4 (341 mg, 1 mmol) was converted to its nitrate salt by reaction with AgNO3 (162 mg,

0.95 mmol) in 7 mL of anhydrous DMF. AgCl was removed by syringe filtration, and the

filtrate was cooled to -10 °C. Acridine precursor A3 (282 mg, 0.1 mmol) was added to

the solution, and the suspension was stirred at -10 °C for 24 h. After treatment with

activated carbon, the reaction mixture was added into 300 mL of diethyl ether. The

yellow slurry was stirred for 30 min, and the precipitate was recovered by membrane

filtration and dried in a vacuum overnight. The solid was dissolved in anhydrous MeOH

and treated with 1 equivalent of 1 M HNO3, stirred at room temperature for 30 minutes

and precipitated by 300 mL anhydrous diethyl ether. The product was further purified by

recrystallization from hot ethanol to give 618 mg of hybrid P4-A3 a microcrystalline

orange material (yield: 87%). 1H NMR (300 MHz, MeOH-d4) δ 8.40 (d, J = 8.8 Hz, 2H),

7.87 (t, J = 6.8, 2H), 7.72 (dd, J = 8.7, 1.2 Hz, 1H), 7.49 (d, J = 8.4 Hz, 2H), 4.31 (t, J =

7.0 Hz, 2H), 4.10 (brs, 2H), 3.91 (t, J = 6.8 Hz, 1H), 3.75 (s, 2H), 3.65 (t, J = 4.9 Hz, 2H),

3.49 (t, J = 4.9 Hz, 2H), 2.58 (s, 3H). 13C NMR (75 MHz, MeOH-d4) δ 167.49, 160.02,

141.40, 136.55, 126.49, 125.30, 119.77, 114.14, 60.73, 49.86, 48.16, 47.39, 23.02. MS

(ESI, positive mode): for C19H28ClN6OPt ([M]+), 587.00; found: 585.2.

Synthesis of conjugate [PtCl(NH3)2C45H47N5O4](NO3)2 (P4-A12). Platinum complex

A4 (75 mg, 0.22 mmol) was converted to its nitrate salt by reaction with AgNO3 (35.6 mg,

0.21 mmol) in 7 mL of anhydrous DMF. AgCl was removed by syringe filtration, and the

filtrate was cooled to -10 °C. Acridine precursor A12 (150 mg, 0.22 mmol) was added to

132

the solution, and the mixture was stirred at 4 °C for 5 days. After treatment with

activated carbon, the reaction mixture was added into 300 mL of diethyl ether. The

yellow slurry was stirred for ~30 min, and the precipitate was recovered by membrane

filtration, washed by 10 mL of anhydrous DCM and dried in a vacuum overnight. The

solid was dissolved in anhydrous MeOH containing 1 equivalent of 1 M HNO3, stirred at

room temperature for 30 minutes and precipitated with 300 mL of anhydrous diethyl

ether. The product was further purified by recrystallization from hot ethanol to give 116

mg of P4-A12 (yield 47 %). 1H NMR (500 MHz, DMF-d7) δ 14.28 (s, 0.5H), 10.61 (s,

0.5H), 9.64 (s, 0.5H), 9.40 (s, 0.5H), 8.81 (s, 2H), 7.10 (s, 2H), 7.28 - 6.00 (m, 15H), 4.74

(s, 3H), 4.50 (s, 3H), 4.59 - 3.87 (m, 10H), 3.65 (t, J = 5.6 Hz, 1H), 3.55 (t, J = 6.3 Hz,

1H), 3.02 (s, 2H), 2.96 (s, 3H), 2.78 (s, 3H), 2.54 - 2.37 (m, 2H), 0.97 - 0.82 (m, 3H). 13C

NMR (126 MHz, DMF-d7) δ 165.54, 161.88, 158.26, 157.48, 156.61, 156.01, 155.67,

142.56, 140.32, 138.34, 136.31, 135.06, 131.60, 130.41, 129.75, 127.93, 125.71, 123.69,

118.93, 115.13, 113.87,113.36, 65.84, 62.39, 48.38, 21.99, 12.96. MS (ESI, positive

mode): for C45H54ClN7O4 Pt([M - 2H]+), 986.37; found: 985.6.

Synthesis of conjugate [PtCl(NH3)2C49H55N5O6](NO3)2 (P4-A13). P4-A13 was

prepared according to the procedure described for P4-A12, by using A13 as the precursor

with the yield of 84%. 1H NMR (500 MHz, DMF-d7) δ 9.65 (s, 0.5H), 9.39 (s, 0.5H),

8.64 (s, 2H), 8.06 - 7.92 (m, 3H), 7.78 - 6.35 (m, 16H), 4.67 - 3.87 (m, 14H), 3.87 - 3.35

(m, 22H), 3.01- 2.98 (m, 1H), 2.59 - 2.36 (m, 2H), 1.79 (m, 6H), 0.90 (s, 3H). 13C NMR

(126 MHz, DMF-d7) δ 165.75, 161.87, 156.85, 156.00, 142.79, 140.58, 140.30, 138.43,

136.29, 134.49, 131.81, 130.48, 129.74, 127.92, 126.02, 115.11, 114.37, 114.15, 113.33,

133

97.85, 70.12, 69.22, 67.32, 65.53, 64.56, 34.29, 34.12, 29.17, 29.00, 25.36, 21.97, 13.15.

MS (ESI, positive mode): for C49H61ClN7O6Pt ([M+H]+), 1074.40; found: 1074.8.

4.4.2. Stability Study of Compound P4-A12

10 µM solutions of compound P4-A12 were prepared in the following buffers: 10 mM

phosphate-buffered saline (PBS), pH 7.4; 10 mM phosphate (PB), pH 7.4; acetate, pH 5.0.

Each sample was heated at 37 °C in the dark for 48 h and then analyzed by LC-ESMS.

The chromatographic separations were performed with a 4.6 mm × 150 mm reverse-

phase Agilent ZORBAX SB-C18 (5 µm) analytical column with the column temperature

maintained at 25 °C and a binary mobile phase system consisting of: solvent A, optima

water, and solvent B, methanol/0.1% formic acid, a gradient of 95% A to 5% A over 20

minutes and a flow rate of 0.5 mL/min.

4.4.3 Cytotoxicity studies and cell viability assays

4.4.3.1. Sample preparation and cell culture

Stock solutions of compound P4-A3, compound P4-A12 and compound P4-A13 were

prepared in DMF and concentrations were determined spectrophotometrically by ε413 =

10000 M-1 cm-1. Cisplatin, endoxifen and estradiol were purchased from Sigma. Stock

solution of cisplatin was prepared in phosphate-buffered saline (PBS) and the

134

concentration was determined by ε300 = 132 M-1 cm-1. Stock solutions of endoxifen and

tamoxifen were prepared in DMSO.

The human breast adenocarcinoma cells MCF-7 (ER-positive), and MDA-MB-231

(ER-negative) were obtained from American Type Culture Collection (Rockville, MD,

USA). MCF-7 cells were maintained in phenol red DMEM/F-12 media (Gibco)

containing 2.4 g/L sodium bicarbonate and 2.5 mM L-glutamine supplemented with 10%

fetal bovine serum (FBS), 10% penstrep (P&S), and 10 μg/mL insulin. MDA-MB-231

cells were cultured in high glucose DMEM media (Gibco) containing 4.5 g/L D-glucose,

110 mg/L sodium pyruvate supplemented with 10% fetal bovine serum (FBS), 10%

penstrep (P&S), and 10% L-glutamine. All cultures were grown at a constant temperature

at 37 °C in a humidified atmosphere containing 5% CO2, and cells were subcultured once

a week in order to maintain cells in logarithmic growth.

4.4.3.2. Cytotoxicity studies

The cytotoxicity studies of selected compounds were carried out by the CellTiter 96

Aqueous Non-Radioactive Cell Proliferation Assay (Promega, Madison, WI, USA) as

described previously.[277] Briefly, cells suspensions were harvested and seeded into 96-

well microplates at a certain density (MCF-7, 10000 cells/well; MDA-MB-231, 12000

cells/well). After overnight incubation, cells cells were treated with varying

concentrations (starting concentrations were 80 or 100 μM, and DMF or DMSO amount

was less than 0.5%) of test compound for 72 h at 37 ºC in an atmosphere of 5% CO2.

After incubation periods of 72 h, 20 μL of MTS/PMS solution was added to each well

and incubated at 37 °C for 4 h. The absorbance of tetrazolium dye was measured at 490

135

nm using an enzyme-linked immunesorbent assay (ELISA) reader. The reported IC50 data

were calculated from non-linear curve fits using a sigmoidal dose-response equation in

GraphPad Prism (GraphPad Software, La Jolla, CA) and are averages of two individual

experiments performed in triplicate.

136

CHAPTER 5

DESIGN OF ENZYMATICALLY CLEAVABLE PRODRUGS OF A POTENT

PLATINUM-CONTAINING ANTICANCER AGENT

The work contained in this chapter was initially published in Chem. Eur. J in 2014. The

manuscript, including figures and schemes, was drafted by Song Ding and Dr. U.

Bierbach and edited by Dr. U. Bierbach before submission to the journal. Since the

publication, changes in both format and content have been made to maintain a consistent

format throughout this work. The experiments described were performed by Song Ding,

except for the the confocal microscopy and cytotoxicity assays which were contributed

by Dr. Amanda J. Pickard

137

5.1. Introduction and Design Rationale

Traditional chemotherapies often suffer from high systemic toxicity and a narrow

therapeutic window. To improve the pharmacokinetics (PK) and toxicity profiles of

anticancer drugs, various avenues are being pursued, such as nano-sized delivery

platforms, receptor-targeted conjugates, and prodrug designs [332, 333]. The rationale

behind the latter approach is to generate a precursor molecule, which is converted post

administration to the bioactive form of the drug either enzymatically or in response to a

chemical stimulus. Bioactivation may occur during absorption, circulation, or at the

tumor site [334]. Potential benefits of lipophilic prodrugs include efficient retention in,

and absorption from, circulation, as well as improved penetration of membranes and

accumulation in target tissues [332].

Figure 5.1. General structure of first- and second-generation platinum–acridine hybrid

agents.

PT-ACRAMTU, (Figure 5.1; ACRAMTU = 1-[2-(acridin-9-ylamino)ethyl]-1,3-

dimethylthiourea) represents the prototype of a class of DNA-targeted platinum–acridine

138

hybrid agents, which have shown exquisite potency in several solid tumor models [295].

Non-small cell lung cancer (NSCLC) cells prove to be particularly sensitive to this

pharmacophore, with the newer derivatives showing IC50 values in NSCLC cell lines in

the low-nanomolar range and activity in tumor xenografts [279, 295]. Unlike cisplatin

[cis-diamminedichloroplatinum(II)] and its analogues, platinum–acridines derived from

PT-ACRAMTU do not cross-link DNA bases but produce structurally unique hybrid

adducts that are an intrinsically more severe form of DNA damage than the former

bifunctional adducts [275, 296, 305]. The classical structure–activity relationship (SAR)

approach based on modular library screening has previously been used to tune the

chemical stability and reduce the off-target reactivity of the pharmacophore [76]. The

desired improvements were achieved essentially by modifying the ligand and donor sets

around the electrophilic metal [295]. These efforts have led to the development of a

derivative (P8-A1, Figure 5.1) that shows three orders of magnitude higher potency than

cisplatin (the IC50 values for P8-A1 in NCI-H460 and A549 lung cancer cells were 1.3

and 3.9 nM, respectively; see the Figure E.3. in APPENDIX E for dose–response curves).

Despite their promising cell kill in chemoresistant, intractable cancers, platinum–

acridines show unfavorable ADME [332] (absorption, distribution, metabolism, and

excretion) properties, which slow their preclinical development. P3-A1, for instance,

while inhibiting the growth of xenografted NCI-H460 tumors in mice, showed signs of

severe toxicity in the test animals, resulting in a low maximum tolerated dose (MTD)

[279]. Mice necropsied after treatment with P3-A1 showed high levels of platinum in

normal tissues, but insufficient accumulation in tumors, as well as discoloration of the

kidneys, a possible sign of hepatotoxicity or nephrotoxicity [279]. To improve the

139

pharmacological properties of platinum–acridines and to potentially open the therapeutic

window for systemic treatment with these agents, we have now designed lipophilic ester-

based derivatives that can be specifically tuned to undergo enzymatic hydrolysis. This

concept has resulted in the first case of a platinum-containing agent that is recognized as

a substrate by human carboxylesterase-2 (hCES-2), a key enzyme involved in the

activation of several anticancer prodrugs [332].

Scheme 1. Synthesis of target compounds. Reagents and conditions: (i) AgNO3, DMF, rt;

(ii) appropriate nitrile–ester (5-7 ─ 5-10), DMF, 60 °C, 4 h, (iii) (1) A1, DMF, 4 °C, (2) 1

M HNO3. Abbreviations: en = ethane-1,2-diamine, pn = propane-1,3-diamine.

5.2. Results and Discussion

5.2.1. Design and chemistry

Platinum–acridines show excellent cytotoxicity but unfavorable drug-like properties.

The most cytotoxic derivatives exist in their fully protonated [pKa (9-aminoacridine) ≈

9.5–10], dicationic form [295]. Because these hybrids are too hydrophilic, they show

140

poor tissue distribution and are most likely removed too rapidly from circulation via renal

clearance [335]. Thus, the goal of the structural modifications introduced here was to

increase the lipophilicity of the agents while maintaining good water solubility. The

alkyl residue of the amidine donor group (residue R, for Y = NH, see Figure 5.1) was

chosen as the site of attachment for the carboxylic acid ester groups. The design involved

installation of a hydroxymethyl (MeOH) or an extended 3-hydroxypropyl (Pr3-OH) group as

R in place of simple Me and Et in prototypes and masking of the terminal OH function as

lipophilic esters (see Scheme 5.1). A primary carboxylic acid, butanoic (butyric) acid,

and a bulkier, branched derivative, 2-propylpentanoic (valproic) acid, were introduced as

acyl components (see Scheme 5.1). Use of the latter residue was inspired by an

analogous valproic amide-based prodrug of the anticancer drug gemcitabine, which is

activated by hCES-2 [336].

Two distinct mechanisms of ester cleavage have to be considered: chemical and

enzymatic hydrolysis. The former mechanism has previously been observed in

chemically related carboxylic acid-modified platinum–acridines containing a reversed

ester linkage (Pt-linker-C(O)OR´, instead of Pt-linker-OC(O)R´ used in this study) [68].

It is dominated by platinum-assisted ester cleavage in a chloride-ion concentration

dependent manner. Low intracellular chloride favors aquation of platinum, which serves

as a Lewis-acidic metallohydrolase and accelerates ester cleavage. The second mode of

activation involves carboxylesterase isoenzymes, in particular hCES-2, which is not only

expressed at high levels in the gastrointestinal tract and the liver, but also in tumor tissue

[332]. The choice of hCES-2 rather than hCES-1 as the target enzyme was based on the

well-documented substrate selectivity of the two forms, according to which hCES-2

141

preferentially recognizes esters containing bulky alcohols (here, the hydroxyl-modified

hybrid agent itself) in combination with relatively smaller acyl moieties [337].

A total of seven ester-protected hybrids were synthesized. Structural diversity in this

set of compounds was achieved by varying the chain length, n, and the nature of the acyl

residue, R´ (Scheme 1). In addition, different non-leaving amines (L) were introduced to

tune the reactivity of the platinum moiety (see discussion below). All Compounds P9-A1

─P15-A1 were generated from the platinum precursors (P9─P15) containing the

appropriate ester-modified nitrile ligands (5-7 ─ 5-10, see the Experimental Section).

The ligand substitution reactions leading to the precursor complexes containing short-

chain nitrile–esters (n = 1, A9 ─ A12) produced significantly lower yields than reactions

performed with the long-chain derivatives (n = 3, A13─A15) (20–30 % vs. 80–90 %).

This previously observed effect can be attributed to the high CH acidity and reactivity of

the nitrile ligands in the former set of derivatives [68]. (Attempts to generate analogous

derivatives with n = 2 failed due to unexpected β-elimination of the hydroxyl group; data

not shown). The final step affording hybrid agents P9-A1─P15-A1 (yield > 75 %,

analytical purity > 95 %) involved addition of the NHMe group in N-(acridin-9-yl)-N´-

methylethane-1,2-diamine (A1) across the metal-activated nitrile triple bond [338],

producing the desired amidine linkage, and subsequent protonation to generate the

dinitrate salts.

An anomaly in the stereochemistry of the addition reactions leading to hybrids P9-

A1─P12-A1 was observed. The extended-chain ester derivatives (P13-A1, P14-A1, and

P15-A1) and the prototype (P3-A1) almost exclusively (> 90 %) exist as E-isomers in

which the platinum moiety and amidine-NMe group adopt a trans configuration with

142

respect to the N(imino)=C double bond, as typically observed in amidine ligands formed

from secondary amines.[339] By contrast, hybrids P9-A1─P12-A1 form a relatively high

amount of the Z-isomer (> 25 %, based on 1-D and 2-D NMR analysis, see Figure

A.47.─ Figure A.49. in APPENDIX A). We attribute this outcome to the increased steric

hindrance produced by the short-chain (n = 1) acyl groups around the nitrile group.

Intramolecular hydrogen bonding between the imino proton and the ester group

(NH•••O=C-O) may also contribute to this effect (see Figure A.47 ─ Figure A.49 in

APPENDIX A) [340].

All newly synthesized hybrids maintain excellent solubility of > 10 mg/mL in relevant

aqueous media. To demonstrate the effect of the pendant ester groups on the

hydrophilicity/lipophilicity balance of the compounds, we studied the partitioning of

selected derivatives between octanol and phosphate-buffered saline (PBS) (expressed as

log[Coctanol/CPBS] = logD, the distribution coefficient for protonable pharmacophores

[334]). The experiment was performed in PBS at pH 7.4 rather than water to suppress

complex aquation and platinum-mediated ester hydrolysis as described in the following

section. This setup also takes into consideration the pH dependence of logD to faithfully

mimic conditions in plasma. For the unmodified, hydrophilic agent, P3-A1, a logD of –

0.98 (±0.19) was determined. By contrast, compound P15-A1, the presumably most

lipophilic derivative (n = 3, valproic ester, L = pn), preferentially partitions into the

octanol phase with a logD of 0.73 (±0.06), which reflects an increase in lipophilicity by

50-fold relative to compound P3-A1. An intermediate logD of –0.31 (±0.06) was

determined for compound P14-A1 (n = 3, L = en, butyric ester). The logD value

generated for this compound has to be interpreted with caution because of minor ester

143

hydrolysis, which was unavoidable under the conditions of the experiment (< 10 %, see

the following section).

5.2.2. Metal-assisted ester hydrolysis.

One of the proposed mechanisms of activation of compounds P9-A1─P15-A1 as

prodrugs involves platinum-promoted ester cleavage. To mimic the chloride ion

concentration differential that exists between serum during circulation and after uptake

into target cells, compounds were incubated at 37 ºC in PBS (~150 mM NaCl, pH 7.4)

and in phosphate buffer (PB, pH 7.4), respectively. The reaction mixtures were analyzed

at appropriate time points by in-line high-performance liquid chromatography–

electrospray mass spectrometry (LC-ESMS). Reaction products were identified as 1+ or

2+ charged molecular ions in mass spectra recorded in positive-ion mode and quantified

by integrating UV-visible HPLC traces at an acridine-specific wavelength (see the

APPENDIX B.4 for a complete set of data).

The time course of ester hydrolysis yielding hydroxyl-modified platinum–acridine and

butyric/valproic acid is summarized for both media in Figure 5.2. Generally, in sets of

analogues characterized by common spacers linking the platinum and ester moieties,

(CH2)n, the valproic esters show significantly slower conversion than the butyric esters,

or no conversion at all (P10-A1─P12-A1 vs. P9-A1, and P13-A1, P14-A1 vs. P15-A1).

In phosphate-buffered solution in the absence of chloride (Figure 5.2A), the most

extensive hydrolysis is observed for butyric ester-based compounds P9-A1 (n = 1) and

P14-A1 (n = 3), with the former producing approximately two-fold higher levels of

cleaved product after 36 h of continuous incubation. Hydrolytic activity is also observed

144

for the valproic ester derivatives P10-A1, P11-A1, and P12-A1 (all n = 1), but at a much

slower rate. Most strikingly, hybrids P13-A1 and P15-A1, which contain the same

secondary acyl moiety but on an extended linker (n = 3), are completely resistant to

cleavage under these conditions. When incubations were performed in buffer

supplemented with physiological chloride, a major reduction in ester hydrolysis of up to

75 % was observed (Figure 5.2B) compared to reactions in chloride-free media,

consistent with the notion that (reversible) aquation of the platinum moiety plays a role in

the cleavage mechanism.

145

Figure 5.2. Cleavage of ester moieties in compounds P9-A1─P15-A1 monitored by

quantitative HPLC for hydrolysis reactions in phosphate buffer, PB, pH 7.4 (panel A),

phosphate-buffered saline, PBS, pH 7.4 (panel B), and in PBS in the presence of hCES-2

(panel C). Plotted data are the mean of three incubations ± standard deviations. Yields of

conversion for compound P14-A1 in panel C represent the sum of chemical (minor) and

enzymatic (major) cleavage, which produced indistinguishable products. Reactions were

performed at 37 ºC.

146

The LC-ESMS profiles of compounds P10-A1─P12-A1 share common features and

support the proposed mechanism of platinum-mediated ester cleavage. We have chosen

compound P9-A1, which was also suitable for a kinetic study by 1H NMR spectroscopy,

for a detailed discussion (for complete sets of LC-ESMS data for all other analogues, see

the Supporting Information). The only hydrolysis product formed in incubations of

compound P9-A1 in PB was identified as a chelate in which the chloro leaving group has

been replaced with the unprotected hydroxyl oxygen of the cleaved butyric ester (see

peak labeled ‘a’ in Figure 5.3A and the corresponding mass spectrum/structure in Figure

5.3D/E, respectively). A minor amount of chemically unchanged P9-A1 and a product

resulting from substitution of chloride by phosphate containing intact ester are also

observed after 36 h of incubation (peaks b and c). The same array of products was

observed when compound P9-A1 was incubated in the presence of sodium chloride in

PBS, but ester cleavage and chelate formation were significantly suppressed based on the

relative abundances of each species (Figure 5.3B). The absence of an opened-chelate

form in this high-chloride environment attests to the inertness of the five-membered

amidine-N/hydroxyl-O chelate. This was also confirmed in incubations with biological

nitrogen and sulfur model nucleophiles, in which this product was completely unreactive

(data not shown). Unlike compounds P10-A1─P12-A1 (n = 1), compound P14-A1 (n =

3) exclusively forms hydrolysis products containing a dangling hydroxyl group,

confirming that a seven- membered, presumably less stable, chelate does not form (see

the APPENDIX B.4).

147

Fig

ure

5.3

. R

ever

se-p

has

e H

PL

C t

race

s fo

r th

e se

par

atio

n o

f re

acti

on p

roduct

s re

sult

ing f

rom

est

er c

leav

age

in c

om

pound P

9-A

1

in P

B (

A),

in P

BS

(B

), a

nd b

y h

CE

S-2

(C

).

Mas

s sp

ectr

a re

cord

ed i

n p

osi

tive-i

on m

ode

for

frac

tions

lab

eled

a–

d a

nd t

he

corr

espondin

g s

tru

cture

s an

d m

/z v

alues

of

the

mole

cula

r io

ns

are

giv

en i

n p

anel

s D

and E

, re

spec

tivel

y.

148

The dramatic effect of chloride ion on the kinetics of ester hydrolysis was confirmed

for compound P9-A1 in arrayed 1H NMR experiments (for an analysis of 1H NMR

spectra, see the Appendix A). Cleavage of the butyric ester follows a (pseudo-) first-

order rate law with rate constants of k = 3.9 ± 10-5 s-1 in PB and k = 1.2 ± 10-5 s-1 in PBS,

which corresponds to half-lives of 5 h and 16 h, respectively (Figure 5.4). Thus, chloride

slows cleavage of the pendant ester in complex P9-A1 by approximately 70 %.

Figure 5.4. Kinetics of ester hydrolysis in compound P9-A1 monitored by 1H NMR

spectroscopy [D2O, pH* 7.0, 2 mM Pt in 10 mM PB (black trace) or PBS (red trace)] at

37 ºC for 24 h. The data sets were fitted to the first-order exponential decay function y =

y0 + A e-x/t, where t-1 is the rate constant, k [h-1], and x is the reaction time, t [h]. Data

points represent means of at least three integrated signal intensities ± standard deviations.

The experiment was performed in duplicate with similar results.

149

On the basis of the above product analysis, a mechanism of platinum-mediated,

intramolecular ester cleavage is proposed (Scheme 5.2). Cleavage is triggered by

(reversible) aquation of the platinum complex. The Lewis-acidic metal assists in the

deprotonation of the aqua to a hydroxo ligand, which undergoes a nucleophilic attack on

the acyl carbon to promote cleavage of the ester linkage. Cleavage results in the loss of

the acyl protecting group as carboxylic acid and in a dangling or chelated

alcohol/alkoxide moiety. High chloride concentrations shift the aquation reaction toward

the chloro-substituted hybrid, which quenches ester cleavage. An intramolecular attack

by platinum-associated hydroxide is also supported by the following observations: (i)

hydrolysis of ester is approximately twice as efficient for complex P11-A1 bearing

ammine (NH3) non-leaving groups than for the en-substituted analogue P10-A1. This

effect is consistent with the lower pKa value of the aqua ligands in Pt–ammine than in Pt–

en complexes [20], which produces higher concentrations of the more reactive hydroxo

ligand. (ii) For derivatives containing the same acyl moiety, ester cleavage is

dramatically reduced for n = 3 vs. n = 1. This effect can be rationalized with the fact that

internal nucleophilic attack by the hydroxo ligand in the former compound requires

formation of a 9-membered, thermodynamically less favorable, macrocyclic intermediate.

5.2.3. Ester Cleavage by human carboxylesterase-2 (hCES-2)

To determine if the compounds, especially those not undergoing chemical ester

cleavage (P13-A1 and P15-A1), are susceptible to deesterification by a

pharmacologically relevant prodrug-converting enzyme, incubations we also performed

150

with recombinant human carboxylesterse-2 (hCES-2). The particular engineered form of

the protein used was sufficiently robust to provide enzymatic activity over long reaction

times (> 30 min), which allows for identification of cleavage products for non-classical,

slowly metabolized substrates [341]. Reactions were performed in a buffer supplemented

with chloride to suppress non-enzymatic hydrolysis and monitored for a relatively short

period of time to minimize the effects of loss of enzyme activity over time and chemical

hydrolysis. Unavoidable contributions from the latter pathway have been subtracted,

where possible, from the data presented in Figure 2C. LC-ESMS analysis of the reaction

mixtures shows minimal or no enzyme-mediated cleavage of ester for compounds P9-

A1─P12-A1, respectively, which contain short spacers (n = 1). By contrast, derivatives

P13-A1─P15-A1(n = 3) show deesterification under these conditions with yields of ~20

% at the 10-h time point. Importantly, the bulky valproic esters in compounds P13-A1

and P15-A1, which are completely resistant to chemical hydrolysis, are efficiently

cleaved by the enzyme. The HPLC trace recorded for compound P9-A1 after 10 h of

incubation with enzyme (Figure 5.3C) shows the same reaction products formed in PB

and PBS, as well as an additional peak (d, ~5 %) not observed for chemical hydrolysis.

Positive-ion mass spectra unequivocally identified this product as the platinum–chloro

complex containing a dangling hydroxyl group (Figure 5.3D,E), consistent with a non-

metal-mediated cleavage mechanism.

151

Figure 5.5. Reverse-phase HPLC traces for the separation of reaction products resulting

from ester cleavage in compound P13-A1 by hCES-2 (A). Mass spectra recorded in

positive-ion mode for fractions labeled a and b and the corresponding structures and m/z

values of the molecular ions are given in panels B and C, respectively.

152

The fact that no opened chelate was detected for chemical hydrolysis of derivatives

with n = 1, but only five-membered N,O chelate, can be taken as evidence that the former

must be an enzyme-specific product. The synthesis of cisplatin derivatives containing

chelated aminoalcoholato ligands requires basic conditions to help deprotonate the

alcohol, while opening of such chelates only occurs under acidic conditions [342]. This is

in agreement with the observation that the two forms (peaks a and d, Figure 5.3C) do not

interconvert under the conditions and on the time scale of the assay. Unlike the butyric

ester derivative P14-A1, which shows dual cleavage reactivity resulting in the

danglinghydroxyl form of the hybrid agent, the valproic ester derivatives P13-A1 and

P15-A1 only undergo enzyme-mediated cleavage to produce the corresponding

deesterified form (shown for compound P13-A1 in Figure 5.5). It can be concluded that

only esters installed on an extended linker are viable substrates for hCES-2 and that only

the chemically robust valproic ester confers true selectivity for enzymatic cleavage.

5.2.4. Cytotoxicity Studies

The cell kill properties of the newly generated compounds P9-A1─P15–A1 were

assessed using a colorimetric cell proliferation assay and compared to that of prototype

P3-A1 (chloride salt). (Note: attempts to synthesize the deesterified, proposed active

hydroxyl forms of P9-A1─P15–A1 in pure form for screening in this assay have been

unsuccessful due to the chemical instability of the platinum–nitrile precursor complexes.)

Two NSCLC adenocarcinoma cell lines were screened for chemosensitivity: A549,

which responds well to treatment with platinum–acridines, and NCI-H1435, a DNA

repair-proficient and highly chemoresistant form of this cancer [343]. The latter cell line

153

has also been shown to express significantly higher levels (3–4-fold) of hCES-2 enzyme

than A549 and other lung adenocarcinomas [344]. Thus, cytotoxicity levels observed in

the two cell lines may also provide hints about potential intracellular prodrug activation.

Both cell lines were dosed with hybrid agents for 72 h at three selected concentrations,

which were based on relative chemosensitivities established for compound P3-A1. A549

was dosed at 5.0, 50, and 500 nM, whereas 20-fold higher concentrations, 0.1, 1.0, and 10

µM, were chosen for NCI-H1435 to account for the relatively higher resistance observed

in this cell line. The dose–response data resulting from the chemosensitivity screen are

presented in Figure 5.6.

As a general trend, the ester-modified derivatives show reduced cytotoxicity levels

compared to the unmodified hybrid, P3-A1, except for compound P14-A1, which shows

a response similar to that of the prototype in both cell lines. Because the butyrate-

protected derivative is efficiently cleaved both chemically and enzymatically on a time

scale relevant to the cell culture assay (see previous section), it is proposed that the

hydroxyl form of this agent is the major contributor to the cell kill. It also suggests that

the n-Pr3-OH residue in deesterified P14-A1 in place of the Et residue in P3-A1 does not at

the DNA level. By contrast, the derivatives containing bulky valproic ester show

compromise the potency of the pharmacophore, possibly pointing to a similar mechanism

greatly reduced activity. Compound P13-A1, for instance, which would generate the

same active hydroxyl form as compound P14-A1, and compound P15-A1 show no cell

kill in A549 at the highest dose and significantly reduced activity in NCI-H1435. This

observation is in agreement with the less efficient intracellular cleavage of the bulkier

valproic esters P13-A1 and P15-A1, which, in their intact form, are less cytotoxic,

154

possibly due to their reduced reactivity with cellular DNA. The same trend is observed

for compounds P10-A1─P12–A1, which proved to be relatively resistant to chemical and

enzymatic hydrolysis. For compounds sharing the same linkers and ester moieties, the

nature of the non-leaving group had only minor or no effects on the activity. Finally,

compound P9-A1, although cleaved very efficiently in chloride-free media (see previous

section), showed activity inferior to the prototype. This outcome was expected because

of the inertness of the 5-membered chelate generated in the process.

Compounds P13-A1 and P15-A1, which are cleaved by hCES-2, show a more

pronounced enhancement in activity relative to compound P3-A1 in NCI-H1435 (high

hCES-2) than in A549 (low hCES-2). The opposite is the case for compound P14-A1.

As a preliminary measure of chemoselectivity, we defined the selectivity index, S: S =

[CVi,NCI-H1435/CV1´,NCI-H1435]/[CVi,A549/CV1´,A549], where CVi and CV1´ are the % viabilities

of the ester-based compounds and compound P3-A1 for their highest doses, respectively.

CV1´ was introduced to normalize for differences in chemosensitivities between the two

cell lines. Assuming that CV1´ is the highest cell kill that can be achieved with any of the

cleaved esters, S values smaller than ‘1’ would indicate a relative advantage of a given

derivative in NCI-H1435, and vice versa. For compounds P13-A1, P14-A1, and P15-A1,

S values of 0.6, 1.4, and 0.5 were calculated, respectively. It can be concluded that the

esters that are not cleaved chemically (P13-A1, P15-A1) but show hCES-2-mediated

cleavage perform relatively better in NCI-H1435. Conversely, compound P14-A1, which

efficiently converts to its active form via an enzyme-independent pathway, appears to

have a relative advantage in A549. These findings may indicate that compounds P13-A1

and P15-A1 act as prodrugs at high levels of hCES-2, which confers sensitivity to NCI-

155

H1435 cells. It should be noted, however, that a direct comparison of cytotoxic

responses between two different cell lines has to be interpreted with caution. Because

cellular uptake and efflux, among other factors, may also contribute to the observed

differences, additional experiments in hCES-2 knockdown or hCES-2 transfected cells11

of the same type are necessary.

Figure 5.6. Results of the chemosensitivity screen of compounds P3-A1─P15-A1 in

A549 (A) and NCI-H1435 (B) cancer cells. Relative cell viabilities are means of two

independent experiments performed in triplicate ± the standard deviation.

156

5.3. Conclusions

Lipophilic, ester-modified derivatives of platinum–acridines have been synthesized and

evaluated with the goal of improving the drug-like properties of these highly cytotoxic

anticancer agents. On the basis of their physicochemical properties, chemical reactivities,

their ability to serve as hCES-2 substrates, and their performance in cell lines,

compounds P13-A1 and P15-A1 can be considered candidates for slow cleavage by

hCES-2. Compound P14-A1 holds promise of providing a dual mode of cleavage, in

which chemical activation may be necessary for applications requiring extended

circulation of precursor in plasma and self-immolative activation in cells lacking hCES-2.

By contrast, the more lipophilic and chemically less reactive analogues P13-A1 and P15-

A1, which should achieve even longer half-lives in circulation, would require activation

by target enzyme expressed in the liver (similar to the anticancer prodrug irinotecan) or in

tumor tissue. Whether the butyric and valproic esters are prone to hydrolysis by other

ubiquitous serum esterases remains to be determined. To this end, incubations of P9-

A1─P15–A1 with fetal bovine serum (FBS, Thermo Scientific HyClone), which is

commonly used as a model for human serum [345], have shown no evidence for this

unwanted reactivity (data not reported).

Another potential advantage of bulky prodrugs of platinum–acridines may be their poor

DNA binding properties. If ester cleavage occurs primarily intracellularly in tumor tissue,

differential DNA recognition by the ester-protected (inactive) and hydroxo (active) forms

of the hybrid agent might confer prodrug selectivity to cancer cells while sparing normal

cells. In addition to an improved ADME profile, lipophilic prodrugs also have the

potential benefit of improving drug safety [332]. To test if this supposition holds for our

157

prodrug design, dose escalation studies were performed with the chemically robust

derivative P15-A1 in Swiss Webster mice. Mice tolerated this analogue without showing

signs of toxicity and weight loss when injected intraperitoneally (i.p.) once a day for five

consecutive days (qd × 5) at a dose of 1.6 mg/kg (see the Supporting Information). For

comparison, compound P3-A1 was significantly more toxic and showed an MTD of 0.1

mg/kg when the same dosing schedule was applied [279].

In conclusion, we have developed a versatile platform for tuning the pharmacological

parameters of potent platinum–acridines and have demonstrated for the first time that a

metal-based pharmacophore is compatible with the classical concept of carboxylesterase-

mediated prodrug activation. This feature may provide a strategy of improving the

ADME and safety of platinum–acridines and other metallopharmacophores.

5.4. Experimental Section

5.4.1. Materials, general procedures, and instrumentation.

All reagents were used as obtained from commercial sources without further

purification unless indicated otherwise. Compound P3-A1[279] (chloride salt) and N-

(acridin-9-yl)-N´-methylethane-1,2-diamine[346] (A1) were prepared according to

published procedures. The platinum-nitrile precursors (P9-P15) were synthesized from

the corresponding nitriles and silver-ion activated diam(m)inedichloroplatinum(II)

complexes. 1H NMR spectra of the target compounds and intermediates were recorded on

Bruker Advance DRX-500 and 300 MHz instruments. Proton-decoupled 13C NMR

158

spectra were recorded on a Bruker DRX-500 instrument operating at 125.8 MHz. 2-D

1H-13C gradient-selected Heteronuclear Multiple Bond Coherence (gsHMBC)

experiments and temperature-dependent spectra were acquired on a Bruker DRX-500

instrument equipped with a TBI probe and a variable-temperature unit. 2-D HMBC

spectra were collected with 2048 pts in t2 (sw = 6510 Hz), 256 pts in t1 (sw = 27670 Hz),

128 scans per t1 increment, and a recycle delay (d1) of 1.5 s. Chemical shifts (δ) are

given in parts per million (ppm) relative to internal standard tetramethylsilane (TMS). 1H

NMR data is reported in the conventional form including chemical shift (δ, ppm),

multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br = broad),

coupling constants (Hz), and signal integrations. 13C NMR data are reported as chemical

shift listings (δ, ppm). The NMR spectra were processed and analyzed using the

MestReNova software package. HPLC-grade solvents were used for all HPLC and mass

spectrometry experiments. LC-ESMS analysis was performed on an Agilent

1100LC/MSD ion trap mass spectrometer equipped with an atmospheric pressure

electrospray ionization source. Eluent nebulization was achieved with a N2 pressure of 50

psi and solvent evaporation was assisted by a flow of N2 drying gas (350 °C). Positive-

ion mass spectra were recorded with a capillary voltage of 2800 V and a mass-to-charge

scan range of 150 to 2200 m/z. To establish the purity of target compounds, samples

were diluted in methanol containing 0.1 % formic acid and separated using a 4.6 mm ×

150 mm reverse-phase Agilent ZORBAX SB-C18 (5 mm) analytical column at 25 °C, by

using the following solvent system: solvent A, optima water, and solvent B,

methanol/0.1 % formic acid, with a flow rate of 0.5 mL/min and a gradient of 95 % A to

5 % A over 30 minutes. HPLC traces were recorded with a monitoring wavelength range

159

of 363-463 nm. Peak integration was done using the LC/MSD Trap Control 4.0 data

analysis software. Analytical purity of greater than 95 % was confirmed this way for all

target compounds prior to analytical and biological experiments (see the Supporting

Information for spectroscopic and analytical data).

Scheme 5.2. The synthesis of precursors 5-7 ─ 5-10.

5.4.2. Synthesis and characterization of target prodrugs

Synthesis of cyanomethyl butyrate (5-7). To the mixture of butyric acid (10.6 g, 0.12

mol), anhydrous potassium carbonate (20.7 g, 0.15 mol) and sodium iodide (22.5 g, 0.15

mol) in 80 ml of acetonitrile was added 2-chloroacetonitrile (7.5 g, 0.1 mol) in 5 ml

acetonitrile at room temperature. When the addition was complete, the mixture was

refluxed for 16 h. The acetonitrile was then removed by rotary evaporation and the

residue was redissolved in 100 ml CH2Cl2, washed with 10% aqueous K2CO3 and brine,

respectively. The organic layers were collected, dried over MgSO4, and concentrated.

This crude material was connected to the vacumm at 45°C overnight to remove unreacted

2-chloroacetonitrile, affording 11.2 g colorless oil (yield: 88%). 1H NMR (300 MHz,

160

CDCl3) δ 4.73 (s), 2.40(t, J = 7.3 Hz, 2H), 1.68(h, J = 7.6 Hz, 3H), 0.98 (t, J = 7.4 Hz,

2H). 13C NMR (75 MHz, Chloroform-d) δ 171.94, 114.58, 48.12, 35.22, 18.09, 13.46.

Synthesis of cyanomethyl 2-propylpentanoate (5-8). Compound 5-8 was prepared

according to the procedure described for 5-7, by using 4-chlorobutanenitrile as the

precursor with the yield of 74%. 1H NMR (300 MHz, CDCl3) δ 4.19 (t, J = 6.0 Hz, 2H),

2.47 (t, J = 7.2 Hz, 2H), 2.31 (t, J = 7.4 Hz, 2H), 2.01 (h, J = 7.0, 5.8 Hz, 2H), 1.66 (h, J

= 7.4 Hz, 2H), 0.96 (t, J = 7.4 Hz, 3H). 13C NMR (75 MHz, CDCl3) δ 173.27, 118.90,

61.86, 35.89, 24.80, 18.28, 14.21, 13.58.

Synthesis of 3-cyanopropyl butyrate (5-9). Compound 5-9 was prepared according to

the procedure described for 5-8, by using valproic acid as the precursor with the yield of

84%. 1H NMR (300 MHz, CDCl3) δ 4.67 (s, 2H), 2.43 (m, 1H), 1.48 (m, 4H), 1.24 (m,

4H), 0.85 (t, J = 7.3 Hz, 6H). 13C NMR (75 MHz, CDCl3) δ 174.84, 114.55, 47.96, 44.69,

34.32, 20.49, 13.84, 0.96.

Synthesis of 3-cyanopropyl 2-propylpentanoate (5-10). Compound 5-10 was

prepared according to the procedure described for 5-9, by using 4-chlorobutanenitrile and

valproic acid as the precursors with the yield of 69%. 1H NMR (300 MHz, CDCl3) δ 4.09

(t, J = 6.0 Hz, 2H), 2.37 (t, J = 6.1 Hz, 2H), 2.30 (m, 1H), 1.92 (m, 2H), 1.40 (m, 4H),

1.18 (m, 4H), 0.81 (m, 6H). 13C NMR (75 MHz, CDCl3) δ 176.10, 118.77, 61.66, 45.09,

34.50, 24.84, 20.56, 14.13, 13.88.

Synthesis of precursor [PtCl(en)C6H9NO2]Cl (P9). A mixture of 0.652 g (2.00 mmol)

of [PtCl2(en)] and 0.338 g (2.00 mmol) of AgNO3 in 10 mL of anhydrous DMF was

stirred at room temperature in the dark for 16 h. Precipitated AgCl was filtered off

through a Celite pad, 1.78 g (14 mmol) of 5-7 was added to the filtrate, and the

161

suspension was stirred at 55 °C for 3 h in the dark. The solution was evaporated to

dryness in a vacuum at 30 °C yielding a yellow residue, which was dissolved in 20 ml of

dry methanol. Activated carbon was added, and the solution was stirred for 15 min.

Carbon was filtered off, and the solution was concentrated to a final volume of 5 mL,

which is further precipitated in 200 mL anhydrous ethyl ether to afford 297 mg (yield:

32%) off-white microcrystalline precipitate. 1H NMR (300 MHz, D2O) δ 5.98 – 5.58 (m,

2H), 5.25 (s, 2H), 2.86 – 2.36 (m, 5H), 1.64 (h, J = 7.4 Hz, 2H), 0.92 (t, J = 7.6 Hz, 3H).

The 13C NMR was not taken due to decomposition.

Synthesis of precursor [PtCl(en)C10H17NO2]Cl (P10). Precursor P10 was prepared

according to the procedure described for P9, by using [PtCl2(en)] and 5-8 as the

precursors with the yield of 37%. 1H NMR (300 MHz, D2O) δ 5.95 – 5.43 (m, 2H), 5.22

(s, 1H), 2.79 – 2.42 (m, 5H), 1.65 – 1.39 (m, 4H), 1.23 (h, J = 7.5 Hz, 4H), 0.83 (t, J =

7.3 Hz, 6H).

Synthesis of precursor [PtCl(NH3)2C10H17NO2]Cl (P11). Precursor P11 was prepared

according to the procedure described for P9, by using [(NH3)2PtCl2] and 5-8 as the

precursors with the yield of 29%. 1H NMR (300 MHz, MeOH-d4) δ 4.77 (s, 2H), 2.46 -

2.35 (m, 1H), 1.62 – 1.33 (m, 4H), 1.32 – 1.13 (m, 4H), 0.82 (t, J = 7.3 Hz, 6H).

Synthesis of precursor [PtCl(pn)C10H17NO2]Cl (P12). Precursor P12 was prepared

according to the procedure described for P9, by using [PtCl2(Pn)] and 5-8 as the

precursors with the yield of 39%. 1H NMR (300 MHz, MeOH-d4) δ 5.95 – 5.43 (m, 2H),

5.22 (s, 1H), 2.83 – 2.38 (m, 5H), 1.78 –1.66 (m, 2H), 1.65 – 1.39 (m, 4H), 1.23 (h, J =

7.5 Hz, 4H), 0.83 (t, J = 7.3 Hz, 6H).

162

Synthesis of precursor [PtCl(en)C12H21NO2]Cl (P13). Precursor P13 was prepared

according to the procedure described for P9, by using [PtCl2(en)] and 5-10 as the

precursors with the yield of 84%. 1H NMR (300 MHz, MeOH-d4) δ 4.21 (t, J = 6.1 Hz,

2H), 3.08 (t, J = 7.1 Hz, 1H), 2.69 – 2.41 (m, 5H), 2.19 – 2.03 (m, 2H), 1.72 – 1.39 (m,

4H), 1.42 – 1.24 (m, 4H), 0.94 (t, J = 7.2 Hz, 6H).

Synthesis of precursor [PtCl(en)C8H13NO2]Cl (P14). Precursor P14 was prepared

according to the procedure described for P9, by using [PtCl2(en)] and 5-9 as the

precursors with the yield of 86%. 1H NMR (300 MHz, MeOH-d4) δ 4.21 (t, J = 6.0 Hz,

2H), 3.08 (t, J = 7.0 Hz, 2H), 2.82 – 2.48 (m, 2H), 2.42 (t, J = 7.4 Hz, 2H), 2.19 – 1.99

(m, 2H), 1.68 (h, J = 7.4 Hz, 2H), 0.98 (t, J = 7.4 Hz, 3H).

Synthesis of precursor [PtCl(pn)C12H21NO2]Cl (P15). Precursor P15 was prepared

according to the procedure described for P9, by using [PtCl2(pn)] and 5-10 as the

precursors with the yield of 84%. 1H-NMR (300 MHz, MeOH-d4) δ 4.76 – 3.94 (m, 2H),

3.00 – 2.87 (m, 2H), 2.81 – 2.27 (m, 5H), 2.00 (p, J = 6.9 Hz, 2H), 1.73 (p, J = 5.4 Hz,

3H), 1.62 – 1.28 (m, 4H), 1.29 – 1.16 (m, 4H), 0.83 (t, J = 7.2 Hz, 8H).

Synthesis of compound [PtCl(pn)C18H21N4](NO3)2 (P8-A1). Compound P8-A1 was

prepared according to the procedure reported previously[289]. 1H NMR (300 MHz,

DMF-d7) δ 13.89 (s, 1H), 9.89 (s, 1H), 8.77-8.67 (m, 2H), 8.19-7.91 (m, 4H),7.66-7.59

(m, 2H), 6.19 (s, 1H), 5.35-4.98 (m, 4H), 4.51 (t, J = 5.7 Hz, 2H), 4.12 (t, J = 6.5 Hz, 2H),

3.48 – 3.37 (m, 4H), 3.19 (s, 3H), 2.69 (s, 3H), 1.83 (brs, 2H). 13C NMR (75 MHz, DMF-

d7) δ 166.57 , 159.26 , 140.81 , 135.92 , 126.48 , 124.52 , 119.60 , 113.67, 65.87 , 47.62,

43.95, 43.32 , 28.59, 22.33 , 15.50.

163

Synthesis of [PtCl(en)C22H27N4O2](NO3)2 (P9-A1). Precursor complex P9 (240 mg,

0.5 mmol) was dissolved in 10 mL of anhydrous DMF and the solution was cooled to -

20 °C. Acridine–amine A1 (138 mg, 0.55 mmol) was added to the solution, and the

suspension was stirred at 4 °C for 24 h. The reaction mixture was added dropwise into

200 mL of anhydrous ethyl ether, and the resulting yellow slurry was vigorously stirred

for 30 min. The precipitate was recovered by membrane filtration, dried in a vacuum

overnight, and dissolved in 20 mL of methanol containing 1 equivalent of HNO3. After

removal of the solvent by rotary evaporation, the crude product was recrystallized from

hot ethanol, affording 2 as a 3:1 mixture of E and Z isomers. Yield: 238 mg (71 %). 1H

NMR (300 MHz, MeOH-d4) δ 8.61 – 8.51 (m, 2H, Acridine-1/8), 8.13 – 7.75 (m, 4H,

Acridine-3/4/5/6), 7.75 – 7.45 (m, 2H, Acridine-2/7), 6.87 – 6.57 (m, 1H, Pt-

NH=CCH2O- ), 5.50 (s, 1.5H, Pt-NH=CCH2O- ,E-isomer), 5.49 – 5.39 (m, 4H, -

NH2CH2CH2NH2-), 4.76 (s, 0.5H, Pt-NH=CCH2O-, Z-isomer), 4.59 (m, 0.5H, -

NHCH2CH2N(CH3)-,Z-isomer), 4.45 (t, J = 6.3 Hz, 1.5H, -NHCH2CH2N(CH3)-,E-

isomer), 4.02 (t, J = 6.3 Hz, 1.5H, -NHCH2CH2N(CH3)-,E-isomer), 3.56 – 3.39 (m, 0.5H,

-NHCH2CH2N(CH3), Z-isomer), 3.14 (s, 2.25H, -NHCH2CH2N(CH3)-,E-isomer), 2.61 –

2.57(m, 4H, -NH2CH2CH2NH2-), 2.38 – 2.06 (m, 2H, -COCH2CH2CH3), 1.76 – 1.40 (m,

2H, -COCH2CH2CH3), 1.02 – 0.75 (m, 3H, -COCH2CH2CH3). 13C NMR (75 MHz,

MeOH-d4) δ 172.50 (-COCH2CH2CH3, E-isomer), 171.76 (-COCH2CH2CH3, Z-isomer),

165.14 (Pt-NH=CCH2O-, Z-isomer), 163.38 (Pt-NH=CCH2O-, E-isomer), 158.48

(Acridine-9C), 139.87 (Acridine-11C/13C), 135.21(Acridine-3C/6C), 124.99 (Acridine-

1C/8C), 124.02(Acridine-2C/7C), 118.28 (Acridine-4C/5C), 112.63 (Acridine-10C/12C),

63.55 (Pt-NH=CCH2O-, E-isomer), 61.87 (Pt-NH=CCH2O- ,Z-isomer), 35.03 (-

164

COCH2CH2CH3), 17.69 (-COCH2CH2CH3), 12.43(-COCH2CH2CH3). MS (ESI, positive-

ion mode): m/z for C24H35ClN6O2Pt ([M-H]+), 669.22; found, 669.3.

Synthesis of [PtCl(en)C26H35N4O2](NO3)2 (P10-A1). Compound P10-A1 was

prepared according to the procedure described for P9-A1 from precursor P10 with a yield

of 69 %. 1H NMR (300 MHz, MeOH-d4) δ 8.82 – 8.40 (m, 2H, Acridine-1/8), 8.11 – 7.78

(m, 4H,Acridine-3/4/5/6), 7.77 – 7.50 (m, 2H,Acridine-2/7), 5.52 (s, 1.5H, Pt-

NH=CCH2O-, E-isomer), 5.42 – 5.13 (m, 4H, -NH2CH2CH2NH2-), 4.64 – 4.54 (m, 0.5H,

NHCH2CH2N(CH3)-, Z-isomer), 4.48 (t, J = 6.5 Hz, 1.5H, NHCH2CH2N(CH3)-, E-

isomer), 4.03 (t, J = 6.4 Hz, 1.5H, -NHCH2CH2N(CH3)-, E-isomer), 3.15 (s, 2.25H, -

NHCH2CH2N(CH3)-, E-isomer), 2.75 – 2.47 (m, 4H, -NH2CH2CH2CH2NH2-), 2.44 –

2.11 (m, 1H, -COCH(CH2CH2CH3)2), 1.54 – 1.13 (m, 8H, -COCH(CH2CH2CH3)2), 0.97

– 0.71 (m, 6H, -COCH(CH2CH2CH3)2). 13C NMR (75 MHz, MeOH-d4) δ 176.65 (-

CO(CH2CH2CH3)2, E-isomer), 176.01 (-COCH(CH2CH2CH3)2, Z-isomer), 167.30 (Pt-

NH=CCH2O-, Z-isomer), 164.85 (Pt-NH=CCH2O-, E-isomer), 159.90 (Acridine-9C),

141.38 (Acridine-11C/13C), 136.77 (Acridine-3C/6C), 126.49 (Acridine-1C/8C), 125.61

(Acridine-2C/7C), 119.88 (Acridine-4C/5C), 114.12 (Acridine-10C/12C), 66.90 (Pt-

NH=CCH2O- ,E-isomer), 65.27 (Pt-NH=CCH2O- ,Z-isomer), 49.87 (-NH2CH2CH2NH2-),

48.16 (-NHCH2CH2N(CH3)-), 46.35 (-COCH(CH2CH2CH3)2), 35.48 (-

COCH(CH2CH2CH3)2), 21.58 (-COCH(CH2CH2CH3)2), 14.31 (-COCH(CH2CH2CH3)2).

MS (ESI, positive-ion mode): m/z for C28H43ClN6O2Pt ([M-H]+), 724.28; found, 724.4.

Synthesis of [PtCl(NH3)2C26H35N4O2](NO3)2 (P11-A1). Compound P11-A1 was

prepared according to the procedure described for P9-A1 from precursor P11 with a yield

of 74 %. 1H NMR (300 MHz, MeOH-d4) δ 8.56 (d, J = 8.7, 2H, Acridine-1/8), 8.03 (ddd,

165

J = 9.2, 5.8, 1.8 Hz, 2H, Acridine-3/6), 7.95 – 7.79 (m, 2H, Acridine-4/5), 7.75 – 7.51 (m,

2H, Acridine-2/7), 5.56 (s, 1.5H, Pt-NH=CCH2O-,E-isomer), 4.67 – 4.54 (s, 0.5H, -

NHCH2CH2N(CH3)-,Z-isomer), 4.48 (t, J = 6.5 Hz, 1.5H, -NHCH2CH2N(CH3)-, E-

isomer), 4.18 (bs, 3H,NH3), 4.02 (t, J = 6.5 Hz, 1.5H, NHCH2CH2N(CH3)-, E-isomer),

3.93 (bs, 3H, NH3), 3.14 (s, 2.25, H-NHCH2CH2N(CH3)-,E-isomer), 2.36 – 2.16 (m, 1H,

-COCH(CH2CH2CH3)2), 1.56 – 1.03 (m, 8H, -COCH(CH2CH2CH3)2), 0.90 – 0.68 (m,

6H, -COCH(CH2CH2CH3)2). 13C NMR (75 MHz, MeOH-d4) δ 176.67 (-

CO(CH2CH2CH3)2, E-isomer), 176.07 (-COCH(CH2CH2CH3)2, Z-isomer), 164.74 (Pt-

NH=CCH2O-, E-isomer), 159.89 (Acridine-9C), 141.39 (Acridine-11C/13C), 136.76

(Acridine-3C/6C), 126.77 (Acridine-1C/8C), 125.60 (Acridine-2C/7C), 119.88 (Acridine-

4C/5C), 114.20 (Acridine-10C/12C), 65.32 (Pt-NH=CCH2O- ,E-isomer), 46.14 (-

COCH(CH2CH2CH3)2), 35.49 (-COCH(CH2CH2CH3)2) 21.58 (COCH(CH2CH2CH3)2),

14.29 (-COCH(CH2CH2CH3)2). MS (ESI, positive-ion mode): m/z for C26H41ClN6O2Pt

([M-H]+), 699.26; found, 699.3.

Synthesis of [PtCl(pn)C26H35N4O2](NO3)2 (P12-A1). Compound P12-A1 was

prepared according to the procedure described for P9-A1 from precursor P12 with the

yield of 77 %. 1H NMR (300 MHz, MeOH-d4) δ 8.61 (dd, J = 8.8, 4.5 Hz, 2H, Acridine-

1/8), 8.14 – 7.95 (m, 2H, Acridine-3/6), 7.87 (dd, J = 9.0, 4.2 Hz, 2H, Acridine-4/5), 7.75

– 7.53 (m, 2H, Acridine-2/7), 6.83 – 6.73 (m, 1H, Pt-NH=CCH2O-), 5.53 (s, 1.5H, Pt-

NH=CCH2O-, E-isomer), 5.29 – 4.88 (m, 4H, -NH2CH2CH2CH2NH2-), 4.68 – 4.58 (m,

0.5H, -NHCH2CH2N(CH3), Z-isomer), 4.55 (t, J = 6.6 Hz, 1.5H, -NHCH2CH2N(CH3)-,E-

isomer), 4.03 (t, J = 6.4 Hz, 1.5H, -NHCH2CH2N(CH3)-,E-isomer), 3.68 – 3.44 (m, 0.5H,

m, 0.5H, - NHCH2CH2N(CH3)-,Z-isomer), 3.15 (s, 2.25H, -NHCH2CH2N(CH3)-, E-

166

isomer), 2.95 – 2.52 (m, 4H, -NH2CH2CH2CH2NH2-), 2.41 – 2.07 (m, 1H, -

COCH(CH2CH2CH3)2), 1.92 – 1.69 (m, 2H, -NH2CH2CH2CH2NH2-), 1.59 – 0.98 (m, 8H,

-COCH(CH2CH2CH3)2), 0.98 – 0.64 (m, 6H, -COCH(CH2CH2CH3)2). 13C NMR (75

MHz, MeOH-d4) δ 176.68 (-CO(CH2CH2CH3)2, E-isomer), 176.22 (-

COCH(CH2CH2CH3)2, Z-isomer), 164.89 (Pt-NH=CCH2O-, E-isomer), 159.90

(Acridine-9C),, 141.41 (Acridine-11C/13C), 136.76 (Acridine-3C/6C), 126.56 (Acridine-

1C/8C), 125.60 (Acridine-2C/7C), 119.88 (Acridine-4C/5C), 113.48 (Acridine-10C/12C),

46.29 (-COCH(CH2CH2CH3)2), 44.42 (-NH2CH2CH2CH2NH2-), 43.63 (-

NH2CH2CH2CH2NH2-), 35.54 (-COCH(CH2CH2CH3)2) , 29.39 (-NH2CH2CH2CH2NH2-),

21.60 (-COCH(CH2CH2CH3)2), 14.31 (-COCH(CH2CH2CH3)2). MS (ESI, positive-ion

mode): m/z for C29H45ClN6O2Pt ([M-H]+), 739.29; found, 739.4.

Synthesis of [PtCl(en)C28H39N4O2](NO3)2 (P13-A1). Compound P13-A1 was

prepared according to the procedure described for P9-A1 from precursor P13 with the

yield of 83 %. 1H NMR (300 MHz, MeOH-d4) δ 8.46 (dd, J = 8.5, 4.3 Hz, 2H, Acridine-

1/8), 7.91 (m, 2H, Acridine-3/6), 7.77 (dd, J = 9.0, 4.2 Hz, 2H, Acridine-4/5), 7. 54 (m,

2H, Acridine-2/7), 6.08 (s, 1H, Pt-NH=CCH2CH2CH2O-), 5.36 – 5.11 (m, 4H, -

NH2CH2CH2NH2-), 4.33 (t, J = 6.6 Hz, 2H, -NHCH2CH2N(CH3)-), 4.03 (t, J = 6.4 Hz,

2H, -NHCH2CH2N(CH3)-), 3.88 (m, 2H, Pt-NH=CCH2CH2CH2O-), 3.03 (s, 3H, -

NHCH2CH2N(CH3)-), 2.97 (m, 2H, Pt-NH=CCH2CH2CH2O-), 2.55 – 2.37 (m, 4H, -

NH2CH2CH2NH2-), 2.10 – 2.03 (m, 3H, Pt-NH=CCH2CH2CH2O-,-

COCH(CH2CH2CH3)2), 1.59 – 0.98 (m, 8H, -COCH(CH2CH2CH3)2), 0.98 – 0.64 (m, 6H,

-COCH(CH2CH2CH3)2). 13C NMR (75 MHz, MeOH-d4) δ 178.08 (-CO(CH2CH2CH3)2),

170.02 (Pt-NH=CCH2O-), 159.98 (Acridine-9C), 141.37 (Acridine-11C/13C), 136.71

167

(Acridine-3C/6C), 126.52 (Acridine-1C/8C), 125.53 (Acridine-2C/7C), 119.85 (Acridine-

4C/5C), 114.10 (Acridine-10C/12C), 64.83 (Pt-NH=CCH2CH2CH2O-), 46.53 (-

COCH(CH2CH2CH3)2), 35.81 (-COCH(CH2CH2CH3)2), 29.39 (-NH2CH2CH2CH2NH2-),

27.15 (Pt-NH=CCH2CH2CH2O-), 21.67 (-COCH(CH2CH2CH3)2), 14.34 (-

COCH(CH2CH2CH3)2). MS (ESI, positive-ion mode): m/z for C30H47ClN6O2Pt ([M-H]+),

752.31; found, 752.4.

Synthesis of [PtCl(en)C24H31N4O2](NO3)2 (P14-A1). Compound P14-A1 was

prepared according to the procedure described for P9-A1 from precursor P14 with the

yield of 89 %. 1H NMR (300 MHz, MeOH-d4) δ 8.72 – 8.46 (m, 2H, Acridine-1/8), 8.01

(ddd, J = 8.1, 6.9, 1.0 Hz, 2H, Acridine-3/6), 7.92 – 7.80 (m, 2H, Acridine-4/5), 7.64

(ddd, J = 8.3, 6.9, 1.2 Hz, 2H, Acridine-2/7), 6.16 (s, 1H, Pt-NH=CCH2CH2CH2O-), 5.44

– 5.22 (m, 4H,-NH2CH2CH2NH2-), 4.42 (t, J = 6.5 Hz, 2H, -NHCH2CH2N(CH3)-), 4.18 (t,

J = 6.5 Hz, 2H, -NHCH2CH2N(CH3)-), 3.98 (t, J = 6.5 Hz, 2H, Pt-NH=CCH2CH2CH2O-),

3.12 – 3.02 (m, 5H, Pt-NH=CCH2CH2CH2O-,-NHCH2CH2N(CH3)-), 2.74 – 2.46 (m, 4H,

-NH2CH2CH2NH2-), 2.34 – 1.99 (m, 4H, Pt-NH=CCH2CH2CH2O-, -COCH2CH2CH3),

1.58 (d, J = 7.3 Hz, 2H, -COCH2CH2CH3), 0.90 (t, J = 7.4 Hz, 3H, -COCH2CH2CH3).

13C NMR (75 MHz, MeOH-d4) δ 175.33 (-COCH2CH2CH3), 170.11 (Pt-

NH=CCH2CH2CH2O-), 160.04 (Acridine-9C), 141.38 (Acridine-11C/13C), 136.69

(Acridine-3C/6C), 126.43 (Acridine-1C/8C), 125.50 (Acridine-2C/7C), 119.87 (Acridine-

4C/5C), 114.12 (Acridine-10C/12C), 64.7 (Pt-NH=CCH2CH2CH2O-), 36.84 (-

COCH2CH2CH3), 32.07 (Pt-NH=CCH2CH2CH2O-), 27.10 (Pt-NH=CCH2CH2CH2O-),

19.37 (-COCH2CH2CH3), 13.94 (-COCH2CH2CH3). MS (ESI, positive-ion mode): m/z

for C26H39ClN6O2Pt ([M-H]+), 697.25; found, 697.3.

168

Synthesis of [PtCl(pn)C28H39N4O2](NO3)2 (P15-A1). Complex P15-A1 was prepared

according to the procedure described for P9-A1 from precursor P15 with the yield of

76 %. 1H NMR (300 MHz, MeOH-d4) δ 8.54 (d, J = 8.6 Hz, 2H, Acridine-1/8), 8.10 –

7.76 (m, 4H, Acridine-3/4/5/6), 7.59 (t, J = 7.4 Hz, 2H, Acridine-2/7), 4.40 (t, J = 6.1 Hz,

2H, -NHCH2CH2N(CH3)-), 4.17 (t, J = 6.2 Hz, 2H, -NHCH2CH2N(CH3)-), 3.97 (t, J =

6.0 Hz, 2H, Pt-NH=CCH2CH2CH2O-), 3.21 – 2.98 (m, 5H, Pt-NH=CCH2CH2CH2O-,-

NHCH2CH2N(CH3)-), 2.94 – 2.54 (m, 4H, -NH2CH2CH2CH2NH2-), 2.40 – 1.95 (m, 3H,

Pt-NH=CCH2CH2CH2O-,-COCH(CH2CH2CH3)2), 1.79 (m, 2H, -NH2CH2CH2CH2NH2-),

1.56 – 0.99 (m, 8H, -COCH(CH2CH2CH3)2), 0.92 – 0.68 (m, 6H, -

COCH(CH2CH2CH3)2). 13C NMR (75 MHz, MeOH-d4) δ 178.02 (-CO(CH2CH2CH3)2),

169.87 (Pt-NH=CCH2CH2CH2O-), 159.85 (Acridine-9C), 141.33 (Acridine-11C/13C),

136.67 (Acridine-3C/6C), 126.61 (Acridine-1C/8C), 125.50 (Acridine-2C/7C), 119.87

(Acridine-4C/5C), 114.13 (Acridine-10C/12C), 64.86 (Pt-NH=CCH2CH2CH2O-), 46.56

(-COCH(CH2CH2CH3)2), 44.31 (-NH2CH2CH2CH2NH2-), 43.63 (-NH2CH2CH2CH2NH2-

), 35.83 (-COCH(CH2CH2CH3)2), 29.29 (-NH2CH2CH2CH2NH2-), 26.76 (Pt-

NH=CCH2CH2CH2O-), 21.69 (-COCH(CH2CH2CH3)2), 14.37 (-COCH(CH2CH2CH3)2).

MS (ESI, positive-ion mode): m/z for C31H49ClN6O2Pt ([M-H]+), 767.33; found, 767.4.

5.4.3. Time-dependent NMR spectroscopy.

NMR spectra in arrayed experiments were collected at 37 °C on a Bruker 500 DRX

spectrometer equipped with a triple-resonance broadband inverse probe and a variable

temperature unit. Reactions were performed with 2 mM platinum complex dissolved in

either 600 µL of 10 mM phosphate buffer (PB, D2O, pH* = 7.0) or in 600 mL of

169

phosphate-buffered saline (1× PBS, D2O, pH* = 7.0). The 1-D NMR kinetics

experiments were carried out as a standard Bruker arrayed 2-D experiment using a

variable-delay list. Each incremented spectrum was processed using the same procedure,

and suitable signals of the ester moiety were integrated. Data were processed with

MestReNova NMR software. The concentrations of platinum complex at each time point

were deduced from relative peak intensities, averaged over multiple signals to account for

differences in proton relaxation, and fitted to a first-order exponential decay function in

Origin 8.0 (OriginLab, Northampton, MA).

5.4.4. Chemical hydrolysis assay.

The ester hydrolysis study of compounds P9-A1─P15–A1 was carried out by

incubating 1 mM of each test compound in 10 mM phosphate buffer (pH 7.4) or 1× PBS

containing ~150 mM NaCl at 37 °C. At various time points samples were withdrawn

from the reaction mixture and analyzed by in-line LC-ESMS. Chromatographic

separations were performed with a 4.6 mm × 150 mm reverse-phase Agilent ZORBAX

SB-C18 (5 μm) analytical column with the column temperature maintained at 25 °C. The

following solvent system was used: solvent A, optima water, and solvent B,

methanol/0.1 % formic acid, at a flow rate of 0.5 mL/min and a gradient of 95% A to 5%

A over 15 min.

170

5.4.5. Enzymatic cleavage assay.

To study ester cleavage in compounds P9-A1─P15-A1 by recombinant human

carboxylesterase-2 (rhCES-2), 30 μM of each compound was incubated with 400 μg/mL

hCES-2 (BD Biosciences, San Jose, CA, USA) at 37 °C in 1× PBS. Aliquots were

withdrawn at various time points, quenched in an equal volume of methanol, and

centrifuged for 5 min at 10000 g to denature and precipitate protein. The supernatant was

collected and subjected to product separation and analysis using in-line LC-ESMS.

Chromatography was performed on a 4.6 mm × 150 mm reverse-phase Agilent ZORBAX

SB-C18 (5 μm) analytical column with the column temperature maintained at 25 °C. The

following solvent system was used: solvent A, optima water, and solvent B, methanol/

0.1% formic acid, at a flow rate of 0.5 mL/min and a gradient of 95% A to 5% A over 15

min.

5.4.6. Determination of partition coefficients (logD).

To obtain octanol-saturated water and water-saturated octanol, 100 mL of PBS was

stirred with 100 mL of octanol for 24 h, followed by centrifugation for 5 min. The

platinum complexes were dissolved in 1.0 mL of octanol-saturated PBS to a typical

concentration of 0.1 mM and then mixed with 1.0 mL water-saturated octanol.

Triplicates of each experiment were mixed in a multi-tube vortexer incubator for 16 h at

room temperature and then centrifuged for 5 min. The layers were separated carefully,

and the content of platinum–acridines was determined spectrophotometrically at 413 nm

(with λ413 = 10,000 M-1 cm-1 in octanol-saturated PBS and λ413 = 8,600 M-1 cm-1 in PBS-

saturated octanol). The partition coefficients (D) of the samples were then determined as

171

the quotient of the concentration of compound in octanol and the concentration in the

aqueous layer. Reported logD values are the mean ± standard deviations of three

determinations.

5.4.7. Cell based assays

5.4.7.1. Cell culture maintenance

The human non-small cell lung cancer cell lines, NCI-H1435 and A549

(adenocarcinomas) were obtained from the American Type Culture Collection (Rockville,

MD, USA). A549 cells were cultured in HAM’s F12K media (Gibco) supplemented with

10 % fetal bovine serum (FBS), 10 % penstrep (P&S), 10 % L-glutamine, and 1.5 g/L

NaHCO3. NCI-H1435 cells were cultured in serum-free 1:1 DMEM/F12 media (Gibco)

containing 2.436 g/L NaHCO3, 0.02 mg/mL insulin, 0.01 mg/mL transferrin, 25 nM

sodium selenite, 50 nM hydrocortisone, 1 ng/mL epidermal growth factor, 0.01 mM

ethanolamine, 0.01 mM phosphorylethanolamine, 100 pM triiodothyronine, 0.5 % (w/v)

bovine serum albumin (BSA), 10 mM HEPES, 0.5 mM sodium pyruvate, and an extra 2

mM L-glutamine (final concentration 4.5 mM). Cells were incubated at a constant

temperature at 37 °C in a humidified atmosphere containing 5% CO2 and were

subcultured every 2–3 days in order to maintain cells in logarithmic growth, except for

slowly proliferating NCI-H1435, which was subcultured every 7 days.

172

5.3.7.2. Cytotoxicity assay.

The cytotoxicity studies were carried out according to a standard protocol using the

Celltiter 96 aqueous nonradioactive cell proliferation assay kit (Promega, Madison, WI).

Stock solutions (5–10 mM) of P3-A1─P15–A1 were made in DMF and serially diluted

with media prior to incubation with cancer cells. All drugs and controls were tested at

the indicated concentrations in triplicate wells on duplicate plates. Incubations were

carried out for 72 h and cell viabilities were determined by comparing drug-treated wells

with control cells.

173

CHAPTER 6

DEVELOPMENT OF A CHEMICAL TOOLBOX FOR THE DISCOVERY OF

MULTIFUNCTIONAL PLATINUM-ACRIDINE AGENTS

174

6.1. Introduction

Platinum-based cytotoxics continue to be important component in cancer chemotherapy

regimens [3, 213] despite the progress that has been made in the field of molecularly

targeted drugs [347] in recent years. However, there is still considerable interest in safer

and more effective platinum complexes with a wider anticancer spectrum and improved

safety profile. Among them, platinum-acridine (Figure 6.1) hybrids represent a novel

class of non-classical platinum based anticancer agents [348], which, unlike cisplatin, do

not generate cross-links in DNA, but intercalate into DNA base pairs and subsequently

produce monofunctional hybrid adducts [348]. This unique mode of DNA binding has

been demonstrated to cause more severe DNA damage than classic bifunctional platinum

complexes [275, 349], as corroborated by their extraordinary cytotoxicity across a broad

range of cancer cell lines. In particular, the IC50 of P8-A1, one of the most active

analogues identified in non-small cell lung cancer cell lines, was found to be 1.3 nM in

NCI-H460 cells, which is about 1000-fold more potent than cisplatin under the same

conditions [68, 76]. Unfortunately, the narrow therapeutic window and chemoresistance

in some cancers have hampered the development of platinum-acridines as therapeutics

[76, 338, 350].

To overcome these drawbacks, we are pursuing multifunctional conjugates, in which

one or more pharmacophores would be chemically hybridized with the platinum-acridine

agents through a stable or cleavable linker [338, 351]. The rationale behind this approach

is to create molecules capable of interacting with multiple molecular targets that are

clinically relevant in a potentially synergistic way [352]. This concept may lead to more

efficacious treatments for heterogeneous and resistant cancers. Several multifunctional

175

platinum conjugates containing Pt(II) [231, 315, 351] or Pt(IV) analogues [6, 7, 258, 353-

355] have been generated, which show promising activity in vitro and in xenograft

models. In principle, the same goal can also be accomplished with drug cocktails or

formulations containing multiple components [356]. However, the drawbacks of

complicated dose regimens are poor patient compliance [357] and potential severe side

effect as a result of drug-drug interactions [358]. The latter is caused by complicated

pharmacodynamics and pharmacokinetic profiles [359].

Figure 6.1. General structure of selected platinum-acridine hybrid agents.

It should be noted that chemical hybridization would not always lead to an

enhancement in biological action. Mutual interference between the pharmacophores, both

mechanistic and structural, mechanistically or structurally are the major concerns in the

design of multifunctional agents, which may, compared with the parent drugs, lead to

176

reduced efficacy and off-target effects [360]. The goal of this study was to devise an

efficient method that allows us to rapidly screen for viable combinations of

pharmacophores and develop the chemistry by which they can be linked. We have

previously described a breast cancer-targeted platinum-acridine-endoxifen conjugate in

an effort to combine a DNA-damaging platinum agent with an antiestrogen moiety [351].

To build the conjugate, an acridine-endoxifen ligand was assembled prior to attachment

of the platinum center [351]. This may not be the most effective way to assemble a

chemically diverse library of conjugates. A better synthetic strategy would be the direct

attachment of functionalized platinum-acridine analogues to another biologically active

or carrier group. This is not a simple task due to the poor solubility of platinum

complexes in most organic solvents and the chemical reactivity of platinum that may

interfere with the subsequent coupling reaction. In this study, carboxylic acid-modified

platinum-(benz)acridine precursors were synthesized (Figure 6.2), which can be attached

to molecules containing amino groups using coupling reagents routinely exploited in

peptide chemistry [361]. Since amino groups are often found in commercially available

clinically relevant drugs, this approach would not require any additional modifications.

By contrast, Cu(I)-catalyzed click chemistry [310, 362], for instance, has the

disadvantage of depending on starting materials that are either expensive or require to

multi-step synthesis.

177

6.2. Results and discussion

6.2.1. Optimization of coupling reaction

In our previous studies, P3-A7 (Figure 6.1), a carboxylic acid modified platinum-

acridine analogue that shows promising potency in NSCLC cell line NCI-H460 [338],

was selected as a precursor for the development of multifunctional agents. To initiate the

optimizations of coupling conditions, 2-azidoethan-1-amine (Figure 6.3, 6-1) was chosen

as a model amine to produce an amide bond in platinum complexes. To our surprise, no

desired product was detected by high-performance liquid chromatography-electrospray

mass spectrometry (LC-ESMS) under any of the conditions tested. We reasoned that the

poor reactivity of P3-A7 may be caused by the bulky platinum moiety, which could

present a huge steric hindrance as other reactants approached. The possibility of installing

an extended linker in place of –(CH2)2COOH in P3-A7, which might reduce steric

hindrance, was dismissed since this modification has been demonstrateded to

compromise the biological activity of the conjugate [338]. In addition, it is unclear if the

amide linkage in conjugates derived from P3-A7 can be cleaved by protease

intracellularly to release the platinum-acridine agents from the conjugate. For the

conjugate design in this study, we selected platinum precursors that allow ester-based

conjugation, and the point of attachment of the linker was moved from the acridine

nitrogen (acidine fragment, see Figure 1.15 in Chapter 1) in P3-A7 to the alkylene chain

attached to the amidine carbon (nitrile fragment, see Figure 1.15 in Chapter 1).

178

Figure 6.2. (A) Unsuccessful coupling reactions using P3-A7 as the precursor. (B)

Structures of ester based platinum-(benz)acridine analogues.

Recently, we have reported a butyric acid-masked lipophilic prodrug (P14-A1) of

platinum-acridines. P14-A1 is activated by a platinum-mediated, self-immolative ester

cleavage mechanism [350], which generates a hydroxyl-modified, highly cytotoxic

platinum-acridine. In continuation of this work, butyric acid was replaced with succinic

acid as an alternative acyl component, leading to the new ester analogues P16-A1 and

P16-B1, which resemble the structure of P14-A1 but contain an extended carboxylic acid

179

for further chemical ligation (Figure 6.2). Benz(c)acridine was also introduced as the

intercalating moiety that has been shown to be a more selective DNA binder and produce

less severe toxicity in animals than acridine itself [363].

The synthesis of carboxylic acid precursors P16-A1 and P16-B1 was accomplished as

shown in Scheme 6.1. Since the nitrile-modified platinum intermediate is highly

hydroscopic and readily decomposes during work-up, we developed a simple three-step,

one-pot synthetic method starting from [Pt(pn)Cl2] (pn = 1,3-diaminopropane). In the

first step, this platinum precursor was reacted overnight with AgNO3 in DMF at room

temperature, which activated the platinum complexes for the subsequent ligand

substitution reaction. The resulting white precipitate in the reaction mixture (AgCl) was

filtered off, and the filtrate was incubated with the ester-modified nitrile precursor at

60 °C for 4 hours. Without any purification at this stage, the reaction mixture, containing

the key intermediate carboxylic acid-modified “platinum-nitrile” complex, was then

cooled to 4 °C and immediately mixed with N1-(acridin-9-yl)-N2-methylethane-1,2-

diamine or N1-(benzo[c]acridin-7-yl)-N2-methylethane-1,2-diamine, which was kept at

4 °C for another 48 hours to allow for formation of amidine linkage as a result of an

efficient amine addition to the platinum-activated nitrile (Scheme 6.1). Finally, the target

compounds were purified by recrystallization from hot ethanol and recovered an overall

yield of 40 - 50% after three steps.

We next explored the coupling conditions for the assembly of the amide based

multifunctional platinum-acridine agents. As a preliminary experiment, P16-A1 was

reacted with 2-azidoethan-1-amine, which is a simple, primary amine. We investigated

various solution-phase conditions with a number of coupling reagents (for the structures

180

of coupling reagents used in this study, see APPENDIX I), and the results are listed in

Table 6.1. (Note: solid-phase synthesis [364] was not attempted since the chemistry

required for removal of the coupled conjugate would be incompatible with the platinum-

acridine moiety.) The carbodiimide-based reagents [361, 365], DCC and EDC, were

unable to mediate the formation of the desired amide bond. By contrast, the target

conjugate was produced with a high conversion yield (> 90%) when uronium-based

HBTU and phosphonium-based PyBOP were used. The very different outcomes can be

attributed to the rapid coupling kinetics of the latter reagents [365]. COMU, another

uronium-based coupling reagent, although a better reagent than HBTU in peptide

synthesis [366], did not perform as effectively as anticipated in the presence of platinum.

Good conversion was also achieved by using excess (10 equivalents) CDI, which

required the removal of unreacted CDI before reaction with the amine. This additional

step may become a major disadvantage when using this method for high-throughput

library assembly.

181

Sch

eme

6.1

. S

ynth

esis

of

targ

et c

om

po

un

ds.

i)

AgN

O3,

DM

F,

rt,

ii)

HO

OC

(CH

2) 2

C(O

)O(C

H2) 3

CN

, D

MF

, 60 ⁰

C,

4 h

, ii

i) N

1-

(acr

idin

-9-y

l)-N

2-m

eth

yle

than

e-1,2

-dia

min

e, D

MF

, 4 ⁰

C, iv

) N

1-(

ben

zo[c

]acr

idin

-7-y

l)-N

2-m

eth

yle

than

e-1,2

-dia

min

e, D

MF

, 4 ⁰

C.

182

To test the influence of alkyl substituents attached to the amines on the coupling

efficiency, the secondary amino group in endoxifen was studied in conjugation reaction

with P16-A1. HBTU and PyBOP were found to mediate amide coupling with high yields

similar to the primary amine (Table 6.1). Since uronium-based reagents may react with

amines, yielding undesired guanidinium byproducts [361, 367], PyBOP was selected as

the coupling reagent for the library assembly. The optimized conditions for amide

formation with this reagent use 1 equivalent of carboxylic acid, 1.1 equivalents of amine,

1 equivalent of DIPEA and 1.5 equivalents of PyBOP. High conversion yields can be

achieved when reactions were performed at room temperature for 16 hours in DMF at a

Pt concentration of 5 mM or higher.

6.2.2 Assembly of a fragment library for multifunctional platinum-acridine

conjugates

To further evaluate the potential of the PyBOP coupling reaction for producing a

structurally diverse library of platinum-acridine conjugates for biological screening, we

designed a fragment library consisting of 2 carboxylic acid-modified platinum and 10

biological relevant amines, and performed 20 conjugation reactions in a parallel manner.

Specifically, we selected P16-A1 and P16-B1 as the carboxylic acid fragments because

of their potent but different biological behaviors found in cancer cells [363]. P16-A1

contains the prototypical platinum-acridine derivative, which preferentially binds to

duplex DNA and effectively kill NSCLC cells at low-nanomolar concentrations. The

183

benz[c]acridinepharmacophore in P16-B1, although generally not as potent as the

acridine-based agents in terms of cytotoxicity, have shown extraordinary sensitivity in

several cancer cell lines harboring p53 mutation [363], a defect known to cause platinum

drug resistance [363].

To demonstrate the scope and utility of the amide/ester-based conjugation chemistry, a

set of 10 amines with diverse structural and biological features were included in the

fragments library. They include (1) the azide linker (6-1), which allows further chemical

ligation [310], (2) fatty amines and phospholipids producing lipophilic conjugates that

can potentially self-assemble into micelles [368] (6-2) or liposomes [369] (6-3) in

aqueous solution, (3) the arginine analogue (6-4) that may enhance cell uptake due to the

interactions of the guanidinium group with phospholipids in the cell membrane [370], (4)

clinically approved drugs such as Rucaparib [371] (6-5, PARP inhibitor), endoxifen [351]

(6-6, ER antagonist), and a gefitinib analogue [372] (6-7, EGFR inhibitor), (5) the

mitochondria-targeted phosphonium moieties [373] (6-8 and 6-9), and (6) the dendrimer

PAMAMAG3.5 [374].

184

Tab

le 6

.1.

Condit

ions

for

Co

upli

ng R

eact

ions

wit

h M

odel

Pri

mar

y (

Top)

and S

econdar

y (

Bott

om

) A

min

es

Entr

ya

Coupli

ng

reag

ents

Bas

e M

ola

r ra

tio o

f re

acta

nts

(Pt:

Am

ine:

coupli

ng r

eagen

t: b

ase)

Pt

conce

ntr

atio

n i

n

reac

tion (

mM

)

Conver

sion

b (

%)

1

DC

C

--

1:1

.2:1

.2:0

5

0

2

DC

C

DIP

EA

1:1

.2:1

.2:1

5

0

3

DC

C+

NH

S

DIP

EA

1:1

.2:1

.2(D

CC

):1.2

(NH

S):

1

5

0

4

ED

C

--

1:1

.2:1

.2:0

5

0

5

ED

C

DIP

EA

1:1

.2:1

.2:1

5

0

6

ED

C+

NH

S

DIP

EA

1:1

.2:1

.2(E

DC

):1.2

(NH

S):

1

5

0

7

CD

I --

1:1

.2:1

.2:0

5

76

± 2

8c

CD

I --

1:1

0:1

.2:0

5

98

± 2

9

HB

TU

D

IPE

A

1:1

.1:1

.1:1

5

94

± 3

10

PyB

OP

D

IPE

A

1:1

.1:1

.1:1

5

99 ±

0.4

11

CO

MU

D

IPE

A

1:1

.1:1

.1:1

5

83

± 1

12

HB

TU

D

IPE

A

1:1

.1:1

.1:1

0.5

22

± 3

13

PyB

OP

D

IPE

A

1:1

.1:1

.1:1

0.5

17

± 2

14

CO

MU

D

IPE

A

1:1

.1:1

.1:1

0.5

18

± 1

185

Tab

le 6

.1 (

conti

nued

)

En

try

a

Coupli

ng

reag

ents

Bas

e M

ola

r ra

tio o

f re

acta

nts

(Pt:

Am

ine:

coupli

ng r

eagen

t: b

ase)

Pt

conce

ntr

atio

n i

n

reac

tion (

mM

)

Conver

sion

b (

%)

15

c C

DI

--

1:1

0:1

.2:0

5

67 ±

3

1

6

HB

TU

D

IPE

A

1:1

.1:1

.1:1

5

92 ±

1

17

P

yB

OP

D

IPE

A

1:1

.1:1

.1:1

5

90 ±

2

18

C

OM

U

DIP

EA

1:1

.1:1

.1:1

5

73 ±

1

19

H

BT

U

DIP

EA

1:1

.1:1

.5:1

5

88 ±

1

20

PyB

OP

D

IPE

A

1:1

.1:1

.5:1

5

99 ±

0.7

aE

ach r

eact

ion w

as p

erfo

rmed

in t

ripli

cate

and c

har

acte

rize

d b

y L

C-E

SM

S w

ithout

puri

fica

tion.

bT

he

conver

sion y

ield

was

calc

ula

ted

bas

ed o

n t

he

inte

gra

tions

of

HP

LC

tra

ces

at a

n a

crid

ine-

spec

ific

wav

elen

gth

. c C

DI

was

pre

-in

cubat

ed w

ith P

t

com

ple

xes

for

5 h

, th

en t

he

reac

tion m

ixtu

re w

as p

reci

pit

ated

in d

ieth

yl

ether

to r

emove

unre

acte

d C

DI

bef

ore

mix

ing w

ith

amin

e.

186

Fig

ure

6.3

. (A

) S

chem

atic

rep

rese

nta

tion o

f th

e co

upli

ng r

eact

ion.

(B)

Str

uct

ure

s of

carb

ox

yli

c ac

id p

recu

rsors

. (C

) S

truct

ure

s of

amin

e pre

curs

ors

.

187

To perform the parallel synthesis of 20 platinum-acridine conjugates, stock solutions

(25 mM) of each reactant were prepared in DMF. A total of 20 microscale reactions were

assembled in 1 mL Eppendorf tubes with a final Pt concentration of 5 mM and incubated

for 16 hours at room temperature. Species generated in each reaction were characterized

by automated LC-ESMS based on their molecular-ion peaks and fragments. The reactions

resulted in clean conversion to the desired conjugates without the formation of major side

products (for conversion yields and MS data, see Table 6.2). A representative HPLC

chromatogram and MS spectra for the characterization of conjugate 6-16 are shown in

Figure 6.4. Two species were found in the HPLC trace, which were identified as the

desired conjugate (peak 2) and the ethyl ester derivative of P16-A1 (peak 1), respectively.

The latter product was produced due to the reaction of unreacted P16-A1 with ethanol,

the solvent chosen for the preparation of LC-ESMS samples, in the presence of excess

coupling reagent in the mixture. To confirm the connectivity produced in the conjugates,

6-16 was synthesized as a representative example on a large scale (100-200 mg) using the

same method, purified by recrystallization and characterized by 2-D NMR spectroscopy

(see Figure A.60-Figure A.62 in APPENDIX A).

188

Figure 6.4. (A) Reverse-phase HPLC trace for the reaction of P16-A1 and 6-7. (B)

ESMS spectrum of 6-25 and target conjugate 6-16 recorded in positive-ion mode.

Characteristic molecular and fragment ions of 6-16 are [M]+, m/z 1113.6; [M+H]2+, m/z

557.3, [M+2H]3+ , m/z 371.6.

The amount of platinum-acridine precursors that were converted to the desired products

in each reaction can be calculated based on their integrations in the HPLC traces at a

chromophore-specific wavelength. In most cases, baseline separations in HPLC traces

were found under identical analytical conditions, which facilitates the determination of

189

conversion yields. Of the 20 reactions, 14 went to completion (> 95%) (see APPENDIX

B6). The coupling efficiency greatly depends on the steric hindrance of the amine

components. This was evidenced by the reactions of carboxylic acid precursors with the

bulky, positively charged triphenylphosphonium (TPP) analogues 6-8 and 6-9. Notably,

while 6-9 containing an extended 11-aminoundecanoic acid linker gave a conversion

yield >98%, derivative 6-8 produced only 70-80% conjugate. Unreacted carboxylic acid

precursors was recovered in the reactions with PAMAM dendrimer G3.5 (6-10), the

macromolecular amine that is much bulkier than any other small molecules. Furthermore,

two reactions with 1,2-distearoyl-sn-glycero-3-phosphoethanolamine (DSPE, 6-3), a

flexible and primary amine, resulted in reaction mixtures lacking any detectable product,

which was attributed to the limited solubility of the lipid in DMF at the concentration

required for the coupling reactions.

Purification of platinum-acridines is generally challenging due to their reactivity in

aqueous media and poor solubility in most organic solvents. Pure derivatives can be

isolated by repeated recrystallization from ethanol or methanol, which is not suitable for

generating libraries. A typical purification method employed in conventional chemical

libraries is preparative HPLC [375, 376], however, when applied to platinum-acridine

analogues, partial decomposition of the platinum complexes is observed. Since the

optimized conditions we have developed generally led to coupling reactions that virtually

went to completion, to accelerate the drug screening process, only reactions with

conversion yields higher than 95% were selected in the subsequent studies. Only these

products were finally confirmed by HR-ESMS (Table 6.2).

190

In summary, we have developed an efficient coupling reaction to produce amide

linkages in the presence of platinum with high conversion yields, which allowed facile

and rapid generation of a library of multifunctional platinum-acridine agents. This

approach can be readily translated to a 96-well plate-based high-throughput screening

platform (Figure 6.5). All the reactions would be carried out in 96 well plates,

subsequently serial diluted with physiological relevant buffers and then, along with all

the precursors included as controls, directly pre-screened in cell-based assays without

purification. Finally, conjugates indicating promising activity in pre-screen assays will be

synthesized and purified, and inhibition concentrations (IC50 values) determined using the

same assay.

191

Fig

ure

6.5

. S

chem

atic

ill

ust

rati

on o

f a

pla

te-b

ased

hig

h t

hro

ughput

dru

g d

isco

ver

y s

cree

n.

192

Tab

le 6

.2. T

he

Conver

sion Y

ield

and H

R-E

SM

S R

esult

s fo

r P

lati

num

-Acr

idin

e B

ased

Conju

gat

es

Com

pound

Code

Conju

gat

es

Conver

sion

yie

ld (

%)a

Cal

cula

ted

[M]+

Obse

rved

[M]+

To

lera

nce

(ppm

)

6-1

1

99 ±

1

808.2

778

808.2

786

0.9

89

6-1

2

97 ±

1

963.4

955

963.4

954

-0.1

04

6-1

3

100 ±

2

910.3

458

910.3

466

0.8

79

6-1

4

96 ±

3

1045.3

619

1045.3

645

2.4

87

6-1

5

97 ±

2

1095.4

227

1095.4

242

1.3

69

193

6-1

6

98 ±

2

1111.3

604

1111.3

615

0.9

89

6-1

7

100 ±

0.2

612.7

688

612.7

696

1.3

06

6-1

8

98 ±

2

858.2

934

858.2

940

0.6

99

6-1

9

97 ±

2

1013.5

111

1013.5

125

1.3

81

6-2

0

99 ±

1

960.3

615

960.3

610

-0.5

21

194

6-2

1

99 ±

1

1095.3

776

1095.3

784

0.7

30

6-2

2

97 ±

1

1145.4

383

1145.4

391

0.6

98

6-2

3

96 ±

3

1162.3

760

1162.3

744

-1.3

76

6-2

4

100 ±

0.5

637.7

766

(2+

)

637.7

769

(2+

) 0.4

70

aE

ach r

eact

ion w

as p

erfo

rmed

in t

ripli

cate

. T

he

conver

sion y

ield

was

cal

cula

ted b

ased

on t

he

inte

gra

tions

of

HP

LC

tra

ce a

t an

acri

din

e-sp

ecif

ic w

avel

ength

. R

eact

ions

wit

h c

onver

sion y

ield

low

er t

han

90%

are

not

report

ed.

195

6.2.3. Aqueous conditions for the synthesis platinum-acridine based conjugates

We are also pursuing to develop the coupling conditions for amide formation in

aqueous media. This would become particularly useful when the molecules, which are

only soluble in water, such as proteins, peptides, and polysaccharides, are chosen to be

conjugated as sensitizers or targeting moieties. Toward this goal, 8-aminooctanoic acid,

an amino acid that is insoluble in most organic solvents but fairly soluble in water, was

selected as a model compound for conjugation with P16-B1 through an amide bond. This

experiment will help establish coupling conditions for tethering short peptides, such as

tumor targeted RGD peptide [377]. It should be noted that the protection and deprotection

steps, which are commonly employed in peptide synthesis, may have to be avoided, as

these harsh conditions may lead to decomposition of the platinum-acridine complexes.

Another application of the extended 8-aminooctanoic acid chain would be as a linker

module for conjugation of sterically hindered molecules, such as proteins and polymers,

which, did not readily react with P16-A1 and P16-B1 (see section 6.2.2).

Scheme 6.2. “Two-step” synthesis of platinum-acridine conjugates in aqueous media.

196

Figure 6.6. (A) Reverse-phase HPLC trace of the reaction mixture for preparation of 6-

27. (B) ESMS spectrum of P16-B1 and target conjugate 6-27 recorded in positive-ion

mode. (c) Structures and characteristic molecular and fragment ions of P16-B1 and 6-27.

197

To avoid cross-reactions of unprotected amino acids in the presence of coupling

reagents, we proposed a simple two-step synthesis method (Scheme 6.2). The crucial

steps in this approach involved the activation of carboxylic acid precursor to an N-

hydroxysuccinimide (NHS) based ester, an intermediate that is known to be stable and

reactive in aqueous solution, followed by complete removal of coupling reagent before

adding the amino acid. Inspired by our previous experience that uronium-based coupling

reagents, such as HBTU, are highly efficient at producing amide bond in platinum-

acridines, TSTU [378], the NHS analogue that works in both organic and aqueous

solution, was selected in this study. To generate the active NHS ester intermediate, 1

equivalent of P16-B1, was incubated with 1.5 equivalents of TSTU in DMF for 16 h.

The product was then precipitated with diethyl ether, which subsequently underwent

centrifugation for two times. Since the platinum complexe is not soluble in diethyl ether,

but TSTU remains dissolved, unreacted TSTU can be separated from the desired product.

The orange precipitate, which is the NHS modified P16-B1 (6-26), was then reacted with

1 equivalent of 8-aminooctanoic acid in PBS (pH = 7.4). The high chloride concentration

in this buffer helps suppress aquation of the platinum complex. This step was performed

at room temperature for another 16 h, and the reaction mixture was characterized by LC-

ESMS as described previously. About 74% of P16-B1 was converted to the desired

product (Figure 6.6A, peak 2) after two steps without the detection of major side products.

Specifically, the formation of possible side products due to cross-reactivities of

unprotected amino acid was not observed. The same procedures have also been used to

conjugate 8-aminooctanoic acids with P16-A1 with similar results (data not shown). In

conclusion, we have demonstrated a two-step conjugation approach for platinum-acridine

198

derivatives that can be performed in aqueous solution with conversion yields >70%.

Future research will address the purification of the reaction mixture, which can

potentially accomplished by recrystallization or size-exclusion chromatography.

6.2.4. “One-tube” assembly of multifunctional platinum-acridine conjugates

Fragment-based drug discovery represents a novel strategy to identify promising drug

leads [378, 379]. However, this approach has not been used for the discovery of metal-

based drugs. In another set of experiments, we explored the feasibility of producing

multifunctional platinum-acridine conjugates by assembling three components without

laborious purifications. Such a method, in combination with automated synthesis, would

lead to a combinatory assay that rapidly generates a large number of platinum-acridine

analogues with diverse structural features and different sensitivities in cancer cells

derived from various types of tumors [21, 380, 381]. As a preliminary experiment, we

attempted to prepare a known platinum-acridine conjugate (6-14) by a “one-tube”

synthesis (Figure 6.7). [Pt(pn)Cl2] was activated with 0.95 equivalents of AgNO3, and the

AgCl was removed by syringe filtration. The filtrate was then combined with 8

equivalents of carboxylic acid modified nitrile precursor 4-(3-cyanopropoxy)-4-

oxobutanoic acid and the mixture reacted for another 4 h at 65 ºC. The mixture was then

cooled to 4 ºC, combined with 1 equivalent of acridine-amine (A1), and incubated at that

temperature for another 48 h. Unreacted nitrile and acridine precursors were removed

using ether precipitation as described in section 6.2.3. These procedures afforded P16-A1

199

with 85 % conversion yield, as confirmed by LC-ESMS (Figure 6.8A). The major side

product detected by LC-ESMS was assigned to complex 6-28, which formed due to the

reaction of unreacted platinum precursor with A1. To generate the target conjugate (6-14),

a solution containing 0.95 equivalents of rucaparib, 1 equivalent of Hünig's base and 1.2

equivalent of PyBOP in DMF was mixed with the solution containing P16-A1 and

allowed to react at room temperature for 16 h. The mixture was then examined

characterized by LC-ESMS, showing that the desired conjugate had formed with a yield

of 72% (Figure 6.8B). The molecular ion and fragments (Figure 6.8C, peak 3, 20.4 min)

of the conjugate were consistent with the data obtained for this compound in section 6.2.2.

In conclusion, we have demonstrated, using multifunctional platinum-acridine agent as an

example, that the synthesis of inorganic complexes can be accomplished through the

fragment-based, one-tube approach. However, the success of this method significantly

relies on the selection of reactions producing high yields at each step. Poor conversion

yield in only one step would lead to a complicated mixture of products.

200

Figure 6.7. Schematic illustration of “one-tube” assembly of multifunctional platinum-

acridine conjugates.

201

Figure 6.8. (A) Reverse-phase HPLC trace of the reaction mixture before the coupling

reaction (Reaction mixture IV). (B) Reverse-phase HPLC trace of the final reaction

mixture (Reaction mixture V). (C) ESMS spectrum of species in the reaction mixture

recorded in positive-ion mode. Characteristic molecular and fragment ions of 6-28 are

[M]+, m/z 555.4, characteristic molecular and fragment ions of P16-A1 are [M]+, m/z

741.5; [M+H]2+, m/z 371.3; characteristic molecular and fragment ions of 6-14 are [M]+,

m/z 1046.6, [M+H]2+, m/z 523.9.

202

6.2.5. Platinum mediated ester hydrolysis

In chapter 5, we demonstrated that at low chloride concentration, ester-based platinum-

acridine prodrugs undergo hydrolysis. Based on our results for P14-A1, the butyrate

analogue, we anticipated that the multifunctional platinum-acridine conjugates would

behave similarly. To evaluate their ester cleavage, seven representative conjugates were

selected and incubated in phosphate buffer (PB, pH 7.4) and phosphate buffered saline

(PBS, pH 7.4, ≈ 150 mM NaCl), which simulate the chloride ion concentration in the

cytosol of cancer cells and in serum during circulation, respectively. The selected

conjugates were generated by the microscale reactions as described before with

conversion yields of > 95%, and the byproducts were removed by precipitating in ethyl

ether followed by size-exclusion chromatography (Sephedex LH-20, 10 × 1 cm column,

eluted with methanol-acetonitrile of 1:1, v/v). All incubations were performed at 37 °C

and analyzed at various time points by LC-ESMS. A representative LC-ESMS profile is

shown for conjugate 6-11 was shown in Figure 6.9. Cleavage of the amide bond was not

detected, suggesting the ester hydrolysis predominated the drug release.The hydrolyzed

products were identified as a mixture of hydroxyl-containing platinum-acridine that had

undergone chloride-exchange reactions with water and formic acid during LC-MS

analysis. The ligand-exchange reactions were significantly suppressed in chloride-

containing PBS buffers. Other species found in the incubation mixture included the intact

conjugates and their ligand-exchange products, particularly in the absence of chloride.

Only minor decomposition was observed in PB after incubation for 48 hours for

conjugates showing slow hydrolytic rates (6-14 and 6-16).

203

Fig

ure

6.9

. (A

) R

ever

se-p

has

e H

PL

C t

race

of

the

hydro

lyti

c pro

duct

s of

6-1

1 in

PB

S (

top)

and i

n P

B (

bott

om

). (

B)

ES

MS

spec

trum

of

hyd

roly

tic

pro

duct

s re

cord

ed i

n p

osi

tive-i

on m

ode.

(C

) S

truct

ure

s an

d c

har

acte

rist

ic m

ole

cula

r io

ns

for

hydro

lysi

s

pro

duct

.

204

The time course of ester hydrolysis in PB and PBS is summarized in Figure 6.10. In

agreement with previous results (see chapter 5), with the exception of 6-11 and 6-22 that

underwent cleavage extremely rapidly, a 2-3 fold increase in ester hydrolysis occurred in

other conjugates when incubated in phosphate buffer without chloride compared with the

reactions performed in buffer supplemented with chloride. This suggests that the selected

conjugates underwent metal-assisted, chloride concentration dependent ester hydrolysis

(Figure 6.11). To our surprise, although the conjugates share common succinyl ester

linkage, -C(CH2)3OC(O)(CH2)2C(O)NH-, their hydrolysis rates vary significantly,

indicating that the amine components have an effect on the stability of the ester. In

general, rapid ester cleavage was found in the conjugates with flexible and lipophilic

primary amines (6-11, 6-12, 6-17, and 6-18) yielding more than 60% hydrolytic products

after incubation for 24 hours in PB, while only approximately 40% of P14-A1 was

cleaved in the same buffer after 36 hours [350]. The type of chromophore does not seem

to affect hydrolysis (6-11 vs 6-22, showing similar hydrolytic rates). A dramatic

reduction in ester cleavage of up to 40% was found in the conjugates with fused aromatic

heterocyclic ring systems connected through secondary amides (6-14, 6-16 and 6-22)

after incubation in PB for 48 hours. An explanation for this effect can not be given.

205

Figure 6.10. Cleavage if ester moieties in selected conjugates in phosphate buffer, PB,

pH 7.4 (A) and phosphate buffered saline, PBS, pH 7.4 (B). Each incubation was

performed in triplicate.

206

Fig

ure

6.1

1. P

ropose

d m

echan

ism

of

chem

ical

hydro

lysi

s.

207

6.3. Conclusions

To overcome the drawbacks of current platinum based chemotherapy, thousands of

platinum complexes with complicated molecular architectures have been produced for the

biological screening. Many endeavors in these platinum based anticancer agents were

made to confer them multifunctionality and tumor selectivity for better anticancer

effacacy. However, the rapid generation of multifunctional platinum complexes for drug

screening was usually impeded by many synthetic difficulties associated with

platinum(II). Due to the presence of platinum (II), even those seemly straightforward

reactions that afford quantitative yields might result in extremely complicated products.

To remove these synthetic bottlenecks for platinum complexes, our solution is to develop

highly reliable and effective platinum(II) compatible reactions as a highly generalized

approach to sequentially assemble common modular fragments. In this chapter, we

developed a highly efficient coupling reaction for the formation amide linkage in

platinum complexes. The synthetic strategy enables the construction of a library of

multifunctional platinum-acridine conjugates in a fast and economical manner, which

would be helpful for the screening of synergistic pharmacophores particularly for

platinum-acridine derivatives in various cancers. The scope and diversity of this approach

can be further expanded by the variations in modular fragments, and the same concept

can even apply to the discovery of other metal-based drugs such as ruthenium and radium,

which were known to activate nitrile for addition reactions. Furthermore, the ultimate

goal of this technology is to assemble an oncologist toolbox composed of various

multifunctional platinum-acridine conjugates that would be selectively sensitive to a

certain type of cancer cells. Since tumor in each patient is unique, this toolbox has the

potential to evolve to a personalized therapeutic option for individual patient in future.

208

6.4. Experimental Section

6.4.1. Materials, general procedures, and instrumentation.

All reagents were used as obtained from commercial sources without further

purification unless indicated otherwise. 1H NMR spectra of the target compounds and

intermediates were recorded on Bruker Advance DRX-500 and 300 MHz instruments.

Proton-decoupled 13C NMR spectra were recorded on a Bruker DRX-500 instrument

operating at 125.8 MHz. 2-D 1H-13C gradient-selected Heteronuclear Multiple Bond

Coherence (gsHMBC) experiments and temperature-dependent spectra were acquired on

a Bruker DRX-500 instrument equipped with a TBI probe and a variable-temperature unit.

2-D HMBC spectra were collected with 2048 pts in t2(sw = 6510 Hz), 256 pts in t1 (sw =

27670 Hz), 128 scans per t1 increment, and a recycle delay (d1) of 1.5 s. Chemical shifts

(δ) are given in parts per million (ppm) relative to internal standard tetramethylsilane

(TMS). 1H NMR data is reported in the conventional form including chemical shift (δ,

ppm), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br =

broad), coupling constants (Hz), and signal integrations. 13C NMR data are reported as

chemical shift listings (δ, ppm). The NMR spectra were processed and analyzed using

the MestReNova software package. HPLC-grade solvents were used for all HPLC and

mass spectrometry experiments. LC-ESMS analysis was performed on an Agilent

1100LC/MSD ion trap mass spectrometer equipped with an atmospheric pressure

electrospray ionization source. Eluent nebulization was achieved with a N2 pressure of 50

psi and solvent evaporation was assisted by a flow of N2 drying gas (350 °C). Positive-

ion mass spectra were recorded with a capillary voltage of 2800 V and a mass-to-charge

209

scan range of 150 to 2200 m/z. HR-ESMS was performed on a Thermo Scientific LTQ

Orbitrap XL equipped with an electrospray source, and the data was processed with

Xcalibur 2.1 (Thermo Scientific). To establish the purity of target compounds, samples

were diluted in methanol containing 0.1 % formic acid and separated using a 4.6 mm ×

150 mm reverse-phase Agilent ZORBAX SB-C18 (5 mm) analytical column at 25 °C, by

using the following solvent system: solvent A, optima water, and solvent B,

methanol/0.1 % formic acid, with a flow rate of 0.5 mL/min and a gradient of 95 % A to

5 % A over 30 minutes. HPLC traces were recorded with a monitoring wavelength range

of 363-463 nm. Peak integration was done using the LC/MSD Trap Control 4.0 data

analysis software. Analytical purity of greater than 95 % was confirmed this way for all

target compounds prior to analytical and biological experiments.

6.4.2. Synthetic procedures and product characterization

Synthesis of [PtClC27H38N6O4](NO3) (P16-A1). P16-A1 was synthesized using a

two-step, one-pot approach. A mixture of 0.678 g (2.00 mmol) of [PtCl2(pn)] and 0.338 g

(2.00 mmol) of AgNO3 in 10 mL of anhydrous DMF was stirred at room temperature in

the dark for 16 h. Precipitated AgCl was filtered off through a syringe filter. 2.96 g (16

mmol) of nitrile precursor 4-(3-cyanopropoxy)-4-oxobutanoic acid was added to the

filtrate, and the suspension was stirred at 65 °C for 4 h in the dark. Then the reaction

mixture was cooled to 4 °C with ice bath. Acridine–amine A1 (0.51 g, 2.02 mmol) was

added to the solution, and the suspension was stirred at 4 °C for another 48 h. To quench

the reaction, the reaction mixture was added dropwise into 200 mL of anhydrous ethyl

ether, and the resulting yellow slurry was vigorously stirred for 30 min. The precipitate

210

was recovered by membrane filtration and recrystallized from hot ethanol, affording P16-

A1 as a yellow solid. Yield: 704 mg (43 %). 1H NMR (300 MHz, MeOH-d4) δ 8.56 (d, J

= 8.7 Hz, 2H), 7.95 (t, J = 7.5 Hz, 2H), 7.81 (d, J = 8.4 Hz, 2H), 7.57 (t, J = 7.6 Hz, 2H),

5.08 (brs, 2H), 4.80 (brs, 2H), 4.38 (t, J = 6.7 Hz, 2H), 4.32 – 4.22 (m, 3H), 3.94 (t, J =

6.6 Hz, 3H), 3.13 – 2.96 (m, 5H), 2.82 – 2.46 (m, 8H), 2.15 (s, 2H), 1.81 (s, 2H). 13C

NMR (75 MHz, MeOH-d4) δ 174.33, 172.68, 168.67, 161.75, 161.75, 154.16, 142.51,

131.29, 126.77, 121.35, 119.47, 117.88, 65.29, 63.64, 44.43, 43.58, 33.34, 32.57, 31.97,

28.99, 14.91. HR-ESMS (positive-ion mode): m/z for C27H38ClN6O4Pt ([M]+), Calculated

740.2291; found, 740.2297 (tolerance: 0.81 ppm).

Synthesis of [PtClC31H40N6O4](NO3) (P16-B1). P16-B1 was synthesized by the same

procedure as P16-A1 with 0.678 g (2.00 mmol) of [PtCl2(pn)] and benzo[c]acridine–

amine B1 (0.61 g, 2.02 mmol) as precursors. Yield: 782 mg (46 %). 1H NMR (300 MHz,

MeOH-d4) δ 9.06 (d, J = 7.9 Hz, 1H), 8.55 (d, J = 8.6 Hz, 1H), 8.27 (t, J = 9.3 Hz, 2H),

8.14 – 7.76 (m, 5H), 7.69 (t, J = 7.7 Hz, 1H), 5.08 (brs, 2H), 4.80 (brs, 2H), 4.35 (t, J =

6.5 Hz, 2H), 4.20 (d, J = 5.9 Hz, 2H), 3.93 – 3.89 (m, 2H), 3.19 – 2.90 (m, 5H), 2.88 –

2.45 (m, 8H), 2.08 (s, 2H), 1.79 (s, 2H). 13C NMR (75 MHz, MeOH-d4) δ 175.59, 173.30,

168.88, 154.31, 150.58, 143.04, 135.22, 132.17, 131.75, 130.20, 128.42, 127.56, 126.32,

125.37, 123.09, 122.94, 122.07, 116.70, 113.97, 64.04, 44.13, 43.48, 32.77, 31.75, 28. 93,

27.50, 15.44. HR-ESMS (positive-ion mode): m/z for C27H38ClN6O4Pt ([M]+),

Calculated 790.2447; found, 790.2463 (tolerance: 2.02 ppm).

Synthesis of [PtCl(pn)C43H47ClFN9O4](NO3)2 (6-14). 100 mg of P16-A1 (0.122

mmol), 71 mg of 6-7 (0.183 mmol), 127 mg of PyBOP (0.244 mmol) and 31 mg of

DIPEA were mixed in 3 ml anhydrous DMF. The reaction mixture was incubated at room

211

temperature for 16 h. To quench the reaction, the reaction mixture was added dropwise

into 200 mL of anhydrous ethyl ether, and the resulting yellow slurry was vigorously

stirred for 30 min. The precipitate was recovered by membrane filtration, which was

recrystallized from the mixture of ethanol and ethyl ethanol (3:1) to give121 mg (yield:

83.4%) yellow solid. 1H NMR (500 MHz, DMF-d7) δ 13.91 (brs, 1H), 10.05 (brs, 1H),

8.91-8.87 (m, 1H), 8.74-8.68 (m, 2H), 8.17-8.12 (m, 1H), 8.03-7.96 (m, 5H), 7.92-7.84

(m, 1H), 7.68-7.51 (m, 4H), 6.72-6.27 (m, 2H), 5.49 (brs, 2H), 5.05 ((brs, 2H), 4.58-4.50

(m, 2H), 4.37-3.99 (m, 6H), 3.86-3.74 (m, 2H), 3.50-3.47(m, 2H), 3.28-3.13 (m, 8H),

2.83-2.53 (m, 8H), 2.30-2.17 (m, 2H), 1.90-1.80 (m, 2H), 1.56-1.43 (m, 3H). 13C NMR

(126 MHz, DMF-d7) δ 172.94, 172.91, 172.74, 171.31, 168.33, 158.56, 158.31, 157.43,

157.22, 156.23, 154.28, 154.08, 146.39, 140.75, 139.82, 135.30, 135.16, 126.27, 125.99,

124.58, 123.82, 119.75, 119.60, 118.95, 116.93, 116.76, 109.17, 109.03, 99.03, 98.59,

97.20, 96.77, 65.45, 63.53, 56.69, 47.53, 45.99, 43.45, 42.82, 41.57, 41.08, 33.05, 28.15,

27.87, 14.92, 13.91. HR-ESMS (positive-ion mode): m/z for C46H57Cl2FN11O4Pt ([M]+),

Calculated 1111.3604; found, 1111.3615 (tolerance: 0.989 ppm).

6.4.3. Assembly of multifunctional platinum-acridine agents through a library

approach

Stock solutions of the carboxylic acid-modified platinum–(benz)acridines (25 mM),

coupling reagent PyBOP (25 mM), amines to be linked via amide bond formation (25

mM), and Hünig's base (25 mM) are prepared in anhydrous DMF. Microscale coupling

reactions are carried out in 1.5-mL Eppendorf tubes by mixing all the reaction

components using the following volumes: carboxylic acid platinum precursor (40 L),

212

PyBOP (60 L), Hünig's base (40 L), and selected amines (44 L). The reaction

mixtures are then replenished with 16 μL fresh DMF to adjust the final platinum

concentration to 5 mM, followed by incubation at room temperature for 16 hours. To

characterize conjugates and to determine conversion yields, 5-L samples are removed

from the PBS solutions and serial diluted with 995 L ethanol containing 0.1% formic

acid prior to LC-ESMS analysis. Chromatographic separations are performed with a 4.6

mm × 150 mm reverse-phase Agilent ZORBAX SB-C18 36 (5 m) analytical column

with the column temperature maintained at 25 °C. The binary mobile phase consisted of:

solvent A, optima water, and solvent B, methanol/0.1% formic acid delivered at a

gradient of 95% A to 5% A over 30 minutes and a flow rate of 0.5 mL/min. The

formation and the extent of conversion of the precursors was 5 monitored in the

corresponding chromatograms using the LC/MSD Trap Control 4.0 data analysis

software. To quench the reactions, 1 mL of diethyl ether is added to precipitate platinum

complexes, which are collected by centrifugation, washed with 0.5 mL of

dichloromethane, and redisolved in 50 µL of phosphate-buffered saline (PBS) (pH 7.4).

Traces of unreacted precursors and byproducts are soluble in diethyl ether, or

dichloromethane, and can be removed by washing the solid and discarding the

supernatant.

6.4.4. Chemical hydrolysis assay

All selected conjugates were produced by the microscale reactions as described before.

To purify these conjugates, the reaction mixtures were precipitated in ethyl ether,

followed by size exclusion chromatography (Sephedex LH-20, 10 × 1 cm column) eluted

213

with methanol-acetonitrile (1:1, v/v). The ester hydrolysis study of compounds was

carried out by incubating 0.1 mM of each test compound in 10 mM phosphate buffer (pH

7.4) or 1× PBS containing ~150 mM NaCl at 37 °C. At various time points samples were

withdrawn from the reaction mixture and analyzed by in-line LC-ESMS.

Chromatographic separations were performed with a 4.6 mm × 150 mm reverse-phase

Agilent ZORBAX SB-C18 (5 μm) analytical column with the column temperature

maintained at 25 °C. The following solvent system was used: solvent A, optima water,

and solvent B, methanol/0.1 % formic acid, at a flow rate of 0.5 mL/min and a gradient of

95 % A to 5 % A over 30 min.

214

CHAPTER 7

LIPOSOMAL DELIVERY OF PLATINUM-ACRIDINE ANTICANCER FOR

NON-SMALL CELL LUNG CANCER

The experiments described in this chapter were performed by Song Ding, except for the

the confocal microscopy which were contributed by Dr. Fang Liu, and animal studies,

which were contributed by Washington Biotechnology Inc. ICP-MS was performed by

the Analytic Chemistry Service facility of Research Triangle Institute, Research Triangle

Park, NC.

215

7.1. Design Rationale

Previous studies have suggested that the severe side-effects of platinum-acridine agents

are likely caused by their indiscriminate accumulation in off-target tissues, such as the

liver and kidneys [279]. To improve the toxicity profile, we hypothesized that the

liposomal formulations might help deliver platinum-acridines to tumor sites passively as

a result of the EPR effect (see chapter 1, section 1.3.3) and slow tumor growth in mice

with higher efficacy than the free drug when injected intravenously [111]. Platinum-

acridine derivatives are generally cationic [279], highly hydrophilic [279], and, like other

charged molecules, not membrane permeable. This would make it difficult to load

platinum-acridines into liposomes. Therefore, a method that allows efficient

encapsulation of platinum-acridines must be developed prior to in vivo studies.

Our search for appropriate lipid components started with the lipids that are used in

FDA-approved Doxil, the liposomal doxorubicin made from hydrogenated soybean

phosphatidylcholine (HSPC), cholesterol and polyethylene glycol-2000-

distearoylphosphatidylethanolamine (mPEG2000-DSPE) (For structures of lipid

components used in this study, see APPENDIX J) [82]. This combination has been

widely used for the encapsulation of other hydrophilic drugs [382]. Specifically, HSPC is

a lipid with a high phase transition temperature (Tm = 55 ˚C), showing an ordered

crystalline phase at physiological temperature, which reduces drug leakage during

circulation. Incorporation of cholesterol into the lipid bilayer increases the stability of the

liposomal formulation [111]. Additionally, liposomes decorated with mPEG2000-DSPE, a

polyethylene glycol-modified lipid, can extend the circulation time of liposome by

evading the mononuclear phagocyte system (MPS, see chapter 1, section 1.3.3) [84, 111].

216

The mPEG with larger molecular weight, however, may slow down the cellular uptake

and compromise in vivo efficacy [383]. Moreover, lipids bearing negative charges, such

as 1,2-dihexadecanoyl-sn-glycero-3-phospho-(1'-rac-glycerol) (DPPG), which was

chosen in liposomal cisplatin (Lipoplatin) [133], will be also added to the lipid

composition. Their electrostatic interactions with the cationic platinum complexes will

not only increase the loading capacity, but also enhance drug retention in circulation.

Another benefit conferred by DPPG to the liposomes is the ability to fuse with the cell

membrane, as demonstrated for lipoplatin [133], which would release the payloads

directly into the cytosol of cancer cells.

7.2. Result and Disscussion

7.2.1. Aqueous stability of platinum-acridine analogues

Prior to liposomal encapsulation of platinum-acridines, we selected P2-A1 (Figure

7.1A) as a representative for their aqueous stability studies in water, 0.9 % saline and

phosphate buffered saline (pH = 7.4) at 60 ºC, the temperature that was required for the

preparation of liposomes. All the samples were incubated for 4 hours, which is

significantly longer than the time required for encapsulation, and the reaction mixtures

were characterized by LC-MS. P2-A1 was found to be quite stable in saline, but undergo

aquation in Milli Q water under testing conditions. Aquation typically occurs when the

environmental chloride level is low and can be suppressed by high level chloride ions in

solution at 37 ºC (for example, 150 mM NaCl) [68]. This experiment suggested that 0.9 %

217

saline can completely inhibit this ligand exchange reaction at high temperature (60 ºC)

for at least 4h. Contrary to our expectations, the formation of a seven-membered, inactive

Pt-N chelate (7-1, Figure 7.1C) was detected in the sample incubated with PBS, even

though the chloride concentration was close to that of a saline solution. The reaction

mixture was then lyophilized and redissolved in D2O, followed by 1H-NMR analysis. The

appearance of the down-field shifted proton signal (10.5 ppm, Figure 7.1B), resulting

from the deshielding effect of platinum, further confirmed the formation of the chelate,

which had been observed previously [277]. The decomposition can be explained with the

coordination of phosphate to platinum, which is negligible at room temperature, but is

accelerated at higher temperature and competes with chloride. Thus, substitution of

chloride is accelerated by phosphate. The labile phosphate is then rapidly replaced by the

intramolecular nucleophilic 9-amino nitrogen of acridine to produce the Pt-N chelate. To

avoid the decomposition induced by the weakly coordinating buffer, we used water and

0.9% saline as the loading “buffers” for the preparation of liposomes.

218

Figure 7.1. (A) Structures of P2-A1 and P8-A1. (B) 1H-NMR characterization of the

same reaction mixture in D2O. (C) LC-ESMS analysis of P2-A1 incubated in water at 60

ºC for 4 hours. Peak 2 indicates the formation of side product (7-1) under the test

conditions.

7.2.2. Optimization of conditions for drug loading

7.2.2.1. Selection of lipids

We first screened the optimal lipid composition for the encapsulation of P8-A1 (Figure

7.1A), the most cytotoxic platinum-acridine analogue identified in NSCLC cells [350]. A

219

library of liposomes (Table 7.1) with various combination of lipid components was

prepared on a small scale (10 mg of total lipids for each) with a 1-mL mini-extruder using

the film-hydration method (see chapter 1). In the screen, all the preparations were

performed in Milli Q water with the same initial drug feeds, defined as the relative ratio

of “mg of Pt complex/mg of total lipids”. The content of mPEG was fixed at 5 mol% in

the library, which is based on the calculation for the estimated surface coverage of

mPEG-DSPE2000 for 100-nm sized liposomes [384]. Higher degrees of PEGylation will

not have a significant impact on the stabilities in circulation, but may introduce multiple

populations of nanoparticles, including mixed micelles and discs, if the content of mPEG

is more than 15 mol% [384, 385]. The liposomal encapsulations were evaluated based on

their encapsulation efficiency (EE%, Table 7.1). In initial experiments, we used

liposomes containing HSPC and cholesterol, such as liposome 1-3 (Table 7.1). No

encapsulation of P8-A1 was observed for these formulations, which can be attributed to

the poor lipid affinity of the cationic platinum complex. The addition of 30 mol% DPPG,

as expected, greatly increases EE% to 32 ± 1%, as evidenced in liposome 4, compared

with all the DPPG-free liposomes. This improvement can be explained with the

electrostatic interaction of DPPG with the cationic payload, which facilitates drug uptake

into the aqueous core. The encapsulation efficiency increased as more DPPG was

incorporated into the lipid bilayers, with the EE% reaching 53 ± 2% for a DPPG content

of 45 mol%.

No consequences to the drug loading were observed when the content of HSPC and

cholesterol was varied (Table 7.1). However, the inclusion of these components is

essential for maintaining the in vivo stability of the liposomes. For example, addition of

220

cholesterol is able to enhance the retention of payloads in the liposomes that are mainly

composed of lipids with a phase transition temperature (Tm) lower than 37 °C. They

increase the permeability of the gel phase bilayers (Tm > 37 °C) under physiological

conditions. In the latter case, incorporation of cholesterol may impede drug release [386],

leading to a reduction in anticancer efficacy. Therefore, two cholesterol free liposomes

were also included in the library. In particular, 87% of P8-A1, the highly hydrophilic

molecule, was successfully encapsulated in formulation 9 made of DPPG, HSPC and

mPEG-DSPE (80:15:5, molar ratio). In fact, high blood stability of cholesterol-free

liposomes comparable to conventional cholesterol-containing liposomes has been

demonstrated when phospholipids with a high Tm, such as DPPC, DPPG and HSPC, are

major components of the bilayers [387]. Decoration of cholesterol free liposomes with

mPEG further improves their in vivo stability and prolongs their circulation time in blood

[111].

221

Ta

ble

7.1

. C

ondit

ions

for

the

Lip

oso

mal

En

capsu

lati

on o

f P

8-A

1.

#

Lip

id C

om

po

nen

ts

Lip

ids

Rat

io

(mola

r)

Init

ial

Dru

g/L

ipid

(wt)

L

oad

ing m

edia

E

E%

1

HS

PC

/CH

OL

/mP

EG

-DS

PE

55/4

0/5

0.2

H2O

0

2

HS

PC

/CH

OL

/mP

EG

-DS

PE

73/2

4/5

0.2

H2O

0

3

HS

PC

/ m

PE

G-D

SP

E

95/5

0.2

H2O

0

4

DP

PG

/HS

PC

/CH

OL

/mP

EG

-DS

PE

30/2

5/4

0/5

0.2

H2O

32

± 1

%

5

DP

PG

/HS

PC

/CH

OL

/mP

EG

-DS

PE

45/1

5/4

0/5

0.2

H2O

53 ±

2%

6

DP

PG

/HS

PC

/mP

EG

-DS

PE

50/4

5/5

0.2

H2O

36

± 2

%

7

DP

PG

/HS

PC

/mP

EG

-DS

PE

8

0/1

5/5

0.2

H2O

87 ±

2%

8

DP

PG

/HS

PC

/mP

EG

-DS

PE

80/1

5/5

0.5

H2O

45

± 2

%

9

DP

PG

/HS

PC

/mP

EG

-DS

PE

71/2

4/5

0.2

H2O

46

± 1

%

10

D

PP

G/H

SP

C/m

PE

G-D

SP

E

47.5

/47.5

/5

0.2

H2O

39

± 1

%

11

D

PP

G/H

SP

C/m

PE

G-D

SP

E

80/1

5/5

0.2

0.9

% S

alin

e

49

± 1

%

12

D

PP

G/H

SP

C/m

PE

G-D

SP

E

80

/15/5

0.2

0.9

% S

alin

e

30 ±

0.3

%

13

D

PP

G/H

SP

C/m

PE

G-D

SP

E

80/1

5/5

0.2

5

% S

ucr

ose

, 0

.9%

Sal

ine

72

± 1

%

14

D

PP

G/H

SP

C/m

PE

G-D

SP

E

80/1

5/5

0.2

5

% S

ucr

ose

, 0

.9%

Sal

ine

40

± 2

%

222

7.2.2.2. The importance of freeze-thaw cycles for drug loading

All the liposomes in the library were prepared using 10 freeze-thaw cycles, during

which the liposomal suspensions were frozen in liquid nitrogen (−196 °C) and thawed at

a temperature above the Tm of the bulk lipids. To evaluate the impact of freeze-thaw

cycles on the encapsulation of P8-A1, a cholesterol free liposome formulation without

undergoing freeze-thaw process has also been generated with the bilayer compositions

identical to liposome 9 in the library. Results are shown in Figure 7.2A. Strikingly, about

54% more drug was encapsulated when applying the freeze-thaw technique, confirming

its importance in the drug loading process. The freeze-thaw technique has been widely

implemented to improve the encapsulation of hydrophilic drugs. It has been proposed that

the process causes a reequilibration of drug distribution inside and outside the liposomes,

most likely due to disruption of the lipid bilayers caused by the formation of ice crystals

[388, 389]. An increase in the trapping volume was also observed after freeze-thaw cycle,

which improves the encapsulation yield [389]. Moreover, since the liposomes lacking

cholesterol are very sensitive to changes in temperature and osmotic gradients, the freeze-

thaw technique is likely to be more effective for the cholesterol free liposomes than

conventional formulations [390].

223

Figure 7.2. (A) Effect of the freeze-thaw cycles on the encapsulation efficiency (EE%) of

DPPG/HSPC/mPEG2000-DSPE (80/15/5, molar ratio). The liposomes were prepared using

the film-hydration method with an initial drug/lipid ratio of 0.2 (wt/wt). (B) Effect of the

loading buffers on the EE% of DPPG/HSPC/mPEG2000-DSPE (80/15/5, molar ratio). The

liposomes were prepared using the film-hydration method with 10 freeze-thaw cycles.

The initial drug/lipid ratio was set as 0.2 (wt/wt). (C) Effect of the drug feeds on the EE%

of DPPG/HSPC/mPEG2000-DSPE (80/15/5, molar ratio). The liposomes were prepared

using the film-hydration method with 10 freeze-thaw cycles. (D) Effect of the

lyoprotectants on the EE% of DPPG/HSPC/mPEG2000-DSPE (80/15/5, molar ratio). The

liposomes were prepared using the film-hydration method with 10 freeze-thaw cycles.

The initial drug/lipid ratio was set as 0.2 (wt/wt).

7.2.2.3. The influence of loading buffers on drug loading

We next assessed the roles of the loading buffers in the liposomal encapsulation process.

Milli Q water, 0.9% saline, and 0.9% saline containing various cryoprotectants (glucose,

224

sucrose and mannitol) were used in this study to hydrate the lipid thin film composed of

DPPG, HSPC and mPEG-DSPE (80:15:5, molar ratio). The drug feeds were set as 0.2.

The EE%, obtained after 10 freeze-thaw cycles, is shown in the Figure 7.2D. The use of

water yielded the formulation with EE% of 88%, which dropped notably to about 49%

when the loading solution was replaced with 0.9% saline. This can be in part explained

by the increased ionic strength of the loading solution, which weakens the electrostatic

interactions of the lipid bilayer with the cationic platinum compounds, and therefore

reduces the amount of encapsulated drug. In addition, the high concentration of NaCl

(~150 mM) in saline inhibited the phase separation of ice and solute during freezing and

thus strongly inhibited liposomal encapsulation [391]. We found that the addition of

cryoprotectants can partly restore the EE%. Specifically, 72% of P8-A1 was encapsulated

using 0.9% saline containing 5% glucose as the loading solution. The cryoprotectants

may form multiple hydrogen bonds with the lipids. This interaction may stabilize the

bilayers, preventing them from being completely disrupted during freeze-thaw cycles,

which leads to the leakage of loaded drug. The increase in the amount of glucose from 5%

to 10% did not result in significant changes in EE. Since P8-A1 may undergo aquation in

water at high temperature, we decided to use 0.9% saline containing 5% as the loading

solution to suppress this process.

7.2.2.4. The influence of drug feeds on drug loading

To evaluate the effects of initial amounts of drug on encapsulation, drug feeds of 0.05,

0.1 and 0.2 were chosen in this study (Figure 7.2C). A formulation with the EE% of 92%

was produced with a drug feed of 0.05, which dropped to 72% as the drug feed was

225

increased to 0.2. This is as expected because with lower drug feed (higher DPPG: P8-A1

ratio), negatively charged lipids in the core can better wrap up the cationic platinum

complexes, which lead to higher EE%. In addition, about 60% of P8-A1 was not

encapsulated in the liposomes when the drug feed was increased to 0.5. This is most

likely caused by amount of dicationic P8-A1, which electrostatically neutralize the

negatively charged DPPG and lead to saturation with payload. Drug feeds higher than 0.5

were found to result in the formation of yellow precipitate regardless of the drug

concentration, possibly due to complete neutralization of lipids with platinum complexes.

7.2.2.5. The influence of loading methods on drug loading

We also compared different methods of formulating liposomes for P8-A1, including

film hydration, reverse-phase evaporation, and ethanol injection [119]. All the

formulations in this study were prepared with the lipid component DPPG, HSPC, and

mPEG-DSPE (molar ratio: 80:15:5) using 0.9% saline containing 5% sucrose as the

loading buffer, with an initial drug feed of 0.2. The film hydration method resulted in the

formulation with the highest EE% of about 80%, while a significant drop in EE% was

observed in the formulations prepared by reverse-phase evaporation (21%). This method

commonly leads to higher EE% than others for encapsulation of hydrophilic drugs

because a large internal aqueous volume is generated in the preparation. This observation

confirms our previous finding that the freeze-thaw procedure is critical for achieving the

highest EE%. Less than 5% of P8-A1 was entrapped in the liposomes using ethanol

injection under the same conditions. Because drug loading with this technique relied on

the self-assembly of the lipids with platinum complexes, poor partitioning of P8-A1

226

between lipids and aqueous solution may be responsible for the lower EE%. In addition

to EE%, the hydrodynamic size of these liposomes generated by different methods were

examined using dynamic light scattering (DLS) measurements. Liposomes prepared by

film hydration and reverse-phase evaporation method, which underwent serial extrusion

through 0.22 µM membranes, were homogeneous with a diameter of approximate 120

nm (polydispersity index, PDI < 0.2), whereas the ethanol injection method without

extrusion generated 133 nm liposomes with a broad size distribution (PDI > 0.2). As a

result, film hydration and subsequent extrusion is a better method for encapsulating

platinum-acridines and was chosen for all formulations prepared for further biological

evaluation.

7.2.3. Characterization of liposomes

Four PEGylated liposomal formulations of P8-A1, denoted as Lipo-1─Lipo-4, were

prepared for the following biological evaluations (compositions and properties are

summarized in Table 7.2). Lipo-1, which had the highest loading capacity, was a

cholesterol-free liposomal formulation with the lipid composition of DPPG, HSPC, and

mPEG-DSPE at a molar ratio of 80:15:5. Lipo-2 was inspired by lipoplatin [133],

liposomal cisplatin, which is currently in phase II/III clinical trials and is composed of

DPPG, HSPC, cholesterol, and mPEG-DSPE. Lipo-3 is composed of the identical lipid

components as Lipo-1 but with lower drug loading capacity. Lipo-4, containing the same

lipids as Lipo-1 yet with reduced level of DPPG in the bilayers, was produced to evaluate

the role of DPPG in the formulation. The average hydrodynamic diameter of all the

227

formulations, determined by dynamic light scattering (DLS), were approximately110-120

nm with small polydispersity indices (PDI) of less than 0.2, suggesting highly

homogeneous size distributions (see Appendix G). The liposomes with a size of about

100 nm are optimal for in vivo applications since they show favorable tumor retention

and long circulation time in blood [111]. We found that liposome size correlates with

drug loading capacity, with larger diameters in the formulations producing higher drug-

to-lipid ratios. In addition, due to the presence of anionic DPPG, all formulations showed

negative zeta potentials. This also confirms that the cationic platinum complexes did not

localize to the liposome surface, but remained trapped within the aqueous core.

Reduction of the zeta potential from ~ -40 mV in Lipo-1 to -29 mV in Lipo-4 was found

as the amount of DPPG decreased. Less negatively charged liposomes are thought to

have a longer circulating time, which may result in an enhanced EPR effect [111]. The

relative amounts of phospholipids and cholesterol were determined with Bartlett

assay[392] and reverse-phase HPLC, respectively. Lipo-1 containing a high level of

DPPG was found to be able to reach a drug-to-lipid ratio of 19.7% (wt/wt), while the

drug-to-lipid ratio of Lipo-2, which contained less DPPG and 9.3% (wt/wt) cholesterol,

was 6.8% (wt/wt). This value considerably dropped to 2.3% (wt/wt) in formulation Lipo-

4, which contained the lowest concentration of DPPG. This observation is consistent with

our earlier conclusion that the content of DPPG dominates the encapsulation efficiency of

P8-A1.

To further characterize the liposomes, transmission electronic microscopy (TEM) was

exploited to reveal structural details. All liposomes were negatively stained with uranyl

acetate prior to TEM image acquisition. Most of the liposomes were found to be spherical

228

with a size of less than 100 nm, which was smaller than the results obtained with DLS.

This difference in apparent size is typically observed in PEGylated liposomes, in which

the hydrophilic PEG residues form a hydration layer on the surface in solution. This

makes the particle size appear large when measured by DLS [111]. TEM also showed an

electron-dense platinum-filled core wrapped in a brighter layer in the cholesterol-free

formulations Lipo-1, Lipo-3, and Lipo-4. Careful measurement of the thickness of the

liposomal bilayer exhibited a range from 6 to 9 nm (Figure 7.4A). Because the thickness

of a double lipid bilayer should be at least 10.4 nm [391], we reasoned that our liposomes

consisted of a single bilayer of lipids.

By contrast, no fine structure was observed in Lipo-2. Ruptured liposomes and disc-

like structures (Figure 7.5B) were found in Lipo-3 and Lipo-4. This typically occurred in

the cholesterol-free liposomes containing DSPE-mPEG after undergoing freeze-thaw

cycles throughout their Tm [124]. This effect was not observed in Lipo-1, the cholesterol

free formulation with the highest loading capacity, suggesting that platinum-acridines

may help stabilize the lipid bilayer during the freeze-thaw process.

229

Figure 7.3. TEM images of Lipo-1(A), Lipo-2(B), Lipo-3(C) and Lipo-4(D). Liposomes

were negatively stained using uranyl acetate.

230

Tab

le 7

.2.

Char

acte

riza

tion o

f F

our

Lip

oso

mal

Fo

rmula

tions

Pre

par

ed f

or

P8-A

1

L

ipo

-1

Lip

o-2

L

ipo

-3

Lip

o-4

Lip

id c

om

ponen

ts

DP

PG

:HS

PC

:DS

PE

-

mP

EG

2k

(mola

r ra

tio:

80:1

5:5

)

DP

PG

:HS

PC

:Chol:

DS

PE

-

mP

EG

2k

(mola

r ra

tio:

40:3

2:2

0:8

)

DP

PG

:HS

PC

:DS

PE

-

mP

EG

2k

(mo

lar

rati

o:

80

:15:5

)

DP

PG

:HS

PC

:DS

PE

-

mP

EG

2k

(mola

r ra

tio:

20:7

5:5

)

Pay

load

a/l

ipid

b

(wt/

wt)

, (r

epo

rted

as %

paylo

ad)

19.7

± 1

.7

6.8

± 0

.4

1.7

± 0

.2

2.3

± 0

.3

Chole

ster

ol

(wt

%)c

--

9.3

± 0

.6

--

--

Siz

e (n

m)d

120.9

± 1

.1

112.3

± 1

.6

110.7

± 1

.2

102.4

± 0

.7

PD

Id

0.0

73 ±

0.0

09

0.0

37 ±

0.0

15

0.1

27 ±

0.0

27

0.1

22 ±

0.0

09

Zet

a pote

nti

al (

mV

)

d

-47.4

± 0

.8

-41.7

± 1

.0

-40.7

± 0

.7

-29.2

± 0

.1

TE

M i

mag

es

a S

pec

trophoto

met

ry. b

Bar

tlet

t as

say. c

Rev

erse

-phas

e H

PL

C.

d D

ynam

ic l

ight

scat

teri

ng (

DL

S).

231

Figure 7.4. (A) TEM image of a single liposome in Lipo-1 highlighting the width of the

lipid bilayer. (B) TEM image of Lipo-3 which was freshly prepared by cycling ten times

through Tm. White arrows indicate the formation of bilayer discs. The scale bar indicates

100 nm.

7.2.4. Stability of liposomes in PBS and during long-term storage

Pre-mature release of platinum complexes in the liposomes before reaching tumor sites

should be avoided to achieve a better therapeutic index. To test their stability under

physiologically relevant conditions, Lipo-1 ─ Lipo-4 were incubated in PBS (pH = 7.4)

at 37 ⁰C for 48 h. Samples were withdrawn at various time points and passed through a

size exclusion column to remove any released P8-A1 [393]. The purified liposomes were

then quantified by UV-Vis spectroscopy and their absorbance compared with that of

controls to determine the amount of drugs within the liposomes. To demonstrate that size

exclusion chromatography is able to effectively separate the free P8-A1 from the

232

liposomes, a mixture of the two forms was loaded onto a column (1 cm × 10 cm) filled

with Sephadex G-25. The chromatographic elution profiles of Lipo-1 and P8-A1 are

shown in Figure 7.5. As can be clearly seen in the elution profile, the liposome eluted

from the column using 5 mL of 0.9% saline, while P8-A1 at the same concentration

required about 9 mL of the same solution.

Figure 7.5. Sephadex G-25 elution profiles for the mixture of 100 µL Lipo-1(1 mg/mL,

based on P8-A1) and 100 µL P8-A1 (1 mg/mL). The sample was eluted with 0.9% saline

solution (pH 6.5 – 7.0) and collected every 0.5 mL. Fractions were quantified by UV-Vis

spectroscopy at a wavelength of 413 nm. The absorbance in each fraction was normalized

as the percentage of the maximal absorbance in the same group. Fractions were analyzed

for Lipo-1 (solid line) and P8-A1 (dash line). Note the difference in the retention times,

allowing the baseline separation of free P8-A1 from liposomes under these conditions.

We next assessed the stability of each formulation in PBS (pH 7.4) at 37 °C. All

liposomes lacking cholesterol (Lipo-1, lipo-3 and Lipo-4) showed little leakage after 48

hours (Figure 7.6), indicating excellent stability under the test conditions. This

observation is consistent with a phase transition temperature greater than 37 °C, which

233

renders the lipid bilayers impermeable to the platinum complexes at physiological

temperature. In addition, the electrostatic interactions of DPPG with P8-A1 may also

enhance the retention of platinum complex within the lipid bilayer. On the other hand, 40%

of P8-A1 was released by Lipo-2 under the same conditions, which can be explained

with the incorporation of cholesterol. Cholesterol is a modulator of membrane mobility

that permeabilizes the bilayer to release platinum complex at temperatures below the

phase transition temperature of the bulk lipid. Moreover, since the fluidity of cholesterol-

free lipid bilayers is highly sensitive to variations in temperature [124], cholesterol-free

liposomes in this study would become more permeable at higher temperature. Lipo-1,

mainly composed of DPPG with a phase transition temperature of 41 °C, was incubated

in PBS at 42 °C (Figure 7.7). 24% of P8-A1 was found to be released after 30 min, and

this amount increased after an incubation time of 1 hour (35%) or after elevating the

temperature to 44 °C (35%). A burst release of 46% P8-A1 was detected when Lipo-1

was incubated at 44 °C for 1 h. However, it should be noted that this stability study in

vitro does not faithfully mimic the in vivo behavior of liposomes. Other parameters, such

as serum proteins and components of the immune system, usually facilitate

decomposition of liposomes in blood, and thus release the payloads in off-target tissues.

The storage stability studies indicated that all liposomal formulations were highly stable

when stored in 0.9% saline at 4 ⁰C for at least 2 months without significant changes in

size and payload content.

234

Figure 7.6. In vitro leakage of P8-A1 from Lipo-1 ─ Lipo-4 during 48 h of incubation in

PBS (pH 7.4) at 37 °C.

Figure 7.7. In vitro release profile of P8-A1 from Lipo-1 in PBS (pH 7.4) incubated at

42 °C and 44 °C for 0.5 hour (red) and 1 hour (blue).

235

7.2.5. In vitro cellular uptake and localization study

Because of the intrinsic blue fluorescence of the 9-aminoacridine, cellular uptake and

intracellular accumulation of P8-A1 in liposomes can be directly observed using confocal

fluorescent microscopy [363]. In this study, P8-A1 (5 µM) and Lipo-1 (containing 5 µM

P8-A1) were incubated with non-small cell lung cancer NCI-H460 cells. Dosed cells

were then co-stained with Lyso Tracker (red fluorescence) to study the accumulation of

the P8-A1 in acidic lysosomes prior to the microscopiy sessions. As shown in Figure 7.8

(A) and (B), higher levels of intracellular blue fluorescence were detected in the Lipo-1

treated cells relative to cells dosed with P8-A1, indicating that the liposomes facilitate the

uptake of P8-A1 into cells. However, it is difficult to quantify the differential uptake,

since significant quenching of the fluorescence occurs when the 9-aminoacridine

chromophore is intercalated between DNA base pairs in the nucleus [310]. After

incubation for 4 hours, P8-A1, consistent with our previous findings [310], was found to

localize to the nucleoli in both groups. Importantly, nucleolar localization of free P8-A1

was observed after 2-hour incubation (Figure 7.8(A)), which was significantly faster than

liposomes, suggesting that they may have different mechanisms of cellular uptake.

A high degree of co-localization of P8-A1 with Lyso Tracker was found in the cells

incubated with free P8-A1 for only 15 min, confirming rapid accumulation of platinum

complexes in the lysosomes after entering into cancer cells. We reason that the

lysosomotropic properties of platinum-acridines can be attributed to their cationic nature,

which may favor interactions with the negatively charged lysosomal membranes to

facilitate sequestration [394]. The lysosomes seem to localize to the perinuclear region

and slowly release P8-A1 into the nucleus. Lack of effective lysosomal escape is the

236

primary reason for drug resistance observed for lysosomotropic agents if their major

biological targets are located in the nucleus [395]. This may be a shortcoming of P8-A1

in resistant cancer cells. To induce apoptosis in these cells, rapid accumulation in the

nucleus may be beneficial.

By contrast, Lipo-1 seemed to internalize into cells through a membrane-fusion

mechanism presumably as a result of the presence of fusogenic phospholipid DPPG [133].

Strikingly, a ring-shaped intense blue fluorescence is observed around the membranes of

cells treated with Lipo-1 after 15 minutes of incubation (Figure 7.8B). The membrane-

associated liposomes persist after 2 hours, indicating that the uptake of Lipo-1, unlike

that of free P8-A1, is a relatively slower process. After fusing with the cell membrane,

Lipo-1 releases loaded P8-A1 directly into the cytosol, as clearly evidenced by the blue

punctuations that does not co-localize with the lysosomes stained by red fluorescent Lyso

Tracker. Accumulation of P8-A1 to the lysosomes was also observed in these cells, but to

a lesser extent in cells treated with free drug at the same time points (Figure 7.8). We

reasoned that since liposomes are able to facilitate the delivery of a large amount of P8-

A1 directly into the cytosols, P8-A1 did not get sequestered by the lysosomes. Elevated

levels of blue fluorescence in vesicles distinct from the red-fluorescent lysosomes,

appeared in the perinuclear region of cells after incubation with Lipo-1 for 4 hours

(Figure 7.8B). This observation suggests that a high concentration of P8-A1 is able to

accumulate close to the nucleus, which may help the drug to reach its target and evade

cytosolic detoxification.

237

Figure 7.8. Confocal microscopy images of (A) P8-A1 (blue, 5 µM) and (B) Lipo-1(blue,

5 µM) incubated in real-time of NCI-H460. Cells were co-stained with LysoTracker Red.

Scale bars represent a distance of 5 μm. White arrow indicates that Lipo-1 internalize

into cells through a membrane-fusion mechanism.

238

Figure 7.9. Colocalization images of cells treated with Lipo-1 (A) and P8–A1 (B) for 4 h

and co-stained with LysoTracker Red. The arrows indicate blue-fluorescent intranuclear

vesicular structures that were separated from lysosome. Scale bars represent a distance of

5 μm.

7.2.6. Acute toxicity studies of liposomal P8-A1

To evaluate the acute toxicity of the formulations prepared in this study, Lipo-1─Lipo-

4 were administered intravenously to female Swiss-Webster mice. The maximum

tolerated dose (MTD) was determined based on clinical signs of toxicity, such as

piloerection, weight loss, and lethargy. Free P8-A1 was tested as a control under the

identical circumstances. P8-A1, in agreement with previous results [279], exhibited

severe systemic toxicity in mice models (Table 7.3) with MTD of 0.4 mg/kg for a single

dose or multiple doses (q4d×4). Typical strategies in the clinic to reverse the platinum-

induced toxicity include the use of hydration and chelating agents or agents which

otherwise react with platinum. Liposomal formulation of cisplatin also showed reduced

side effects. A two fold increase in MTD was observed with Lipo-2, which containing

cholesterol in the bilayer. However, the toxicity was not alleviated by encapsulation of

239

P8-A1 with cholesterol-free liposomes (Table 7.4). Lipo-1, Lipo-3 and Lipo-4 were

found to have the same MTD as the free drug. Since negatively charged liposomes are

rapidly recognized by the mononuclear phagocyte system (MPS), we reasoned that

macrophage uptake may be involved in the mechanism of toxicity of these formulations.

The data may indicate delivery of highly cytotoxic P8-A1 to liver and spleen, impairing

phagocytic activity of liver macrophages [396]. One common way to address this

problem, as suggested in other studies, is to co-inject a large amount of blank liposomes

or a similar formulation with extremely low drug/lipid ratio, so that it saturates the MPS

and protects the macrophages from the effects of cytotoxic agents [84, 392, 397].

7.2.7. In vivo antitumor efficacy

A549 (NSCLC, human adenocarcinoma, wild-type p53) was selected as the xenograft

model (athymic nude mice) to test the efficacy of P8-A1 and two formulations. Before a

decision was made in favor of this model, P8-A1 was tested in A549 cells using a

colorimetric cell proliferation assay, in which the derivative was highly active with an

IC50 value of 3.9 nM. The in vivo treatment schedule consisted of five i.v. injections at

doses of 0.4 mg/kg (days 0, 4, 8, 12, 16; q4d5; n = 8). Test animals were monitored for

weight loss and signs of toxicity three times a week. Data analysis for weight and tumor

volumes was done with the Student’s t-test. Only P values of less than 0.05 were

considered as statistically significant. One of the animals treated with Lipo-1 (group 3)

had to be euthanized prematurely because its tumor had grown approximately five times

the size (V = 2677 mm3, day 21) compared to the average of other tumors in that group

240

(passed Grubbs’ outlier test at P < 0.01). Few additional obvious outliers (P < 0.01 and P

< 0.05, confirmed by Grubbs’ outlier test) were identified, which caused major

differences in the variation of tumor sizes between the four groups and resulted in large

error bars (SEM). These artificially large error bars mask the true drug response of the

majority of tumors and were removed from the data sets.

The result of antitumor effects are presented in Figure 7.10A. While Treatment with

P8-A1 and Lipo-2 had no effect on tumor volumes compare to control, a modest, but

significant (P < 0.05), effect on tumor growth was found in the group treated with Lipo-

1, resulting in a ~40% reduction of final tumor volumes relative to untreated mice. To

further evaluate the systemic toxicity of liposomal formulations compared with free P8-

A1, the body weight change of mice were monitored and recorded in Figure 7.10A. Mice

in all treatment groups showed significant weight loss compared to control, however, the

mice dosed with liposomal formulations showed significantly less weight loss than mice

treated with unencapsulated P8-A1. In particular, weight loss was less than 20% and

appeared to be reversible after dosing was discontinued.

Tumors and organs were recovered from necropsied test animals on day 30, 14 days

after the last dose was administered. Residual platinum levels were determined by ICP-

MS to assess if differences in the tissue distribution exist between the free drug and the

two formulations. Lipo-1 (Group 3) produced significantly lower platinum levels in

liver, kidneys, and lungs than unencapsulated P8-A1 (Group 2). Nevertheless, between 1

and 6 ng Pt were detected in the tumors (similar to levels detected in the lungs; data not

corrected for tumor weight/volume; data not shown, n = 5) in all treatment groups.

Differences between the treatment groups were not significant. It also appeared that P8-

241

A1 was delivered more efficiently to large tumors, which faithfully mimic the poor

lymphatic drainage and defective blood vessels in the tumor tissue. This observation is

consistent with enhanced accumulation through EPR effect.

7.3. Conclusion

Liposomal formulations have been extensively investigated to improve the therapeutic

index of cytotoxic agents. In this study, we have developed a DPPG based liposome for

the most cytotoxic platinum-acridine analogue in non small lung cancer cells, P8-A1. A

method that ensures efficient drug loading has been optimized and established. Due to the

presence of DPPG, the liposome was found to be internalized into cells through a

membrane-fusion mechanism. To evaluate their in vivo anticancer efficacy, four

liposomal formulations were tested in A549 xenograft models in mice. Moderate

inhibition of tumor growth was found in mice dosed with Lipo-1, while the other

formulations and free P8-A1, had no effect on tumor size. Future studies will focus on

improving the anticancer efficacy of the liposomal formulations, including optimization

of lipid components, alteration of drug loading capacity and incorporation of tumor-

targeting ligands into the lipid bilayers. Alternatively, less cytotoxic platinum-acridine

analogues may show an advantage as payloads.

242

Ta

ble

7.3

. M

TD

Stu

dy f

or

P8

-A1

Co

mpound

D

osi

ng R

egim

en (

I.V

.)

Sig

ns

of

Tox

icit

y

P8

-A1

0.1

mg/k

g

No o

bvio

us

signs

of

tox

icit

y

P8

-A1

0.4

mg/k

g

Thre

e m

ice

hav

e m

ild p

iloer

ecti

on p

ost

day 6

P8

-A1

0.8

mg/k

g

All

mic

e hav

e pil

oer

ecti

on p

ost

day

11

P8

-A1

0.4

mg/k

g, qd×

5

All

mic

e hav

e pil

oer

ecti

on p

ost

day

4

P8

-A1

0.4

mg/k

g, q4d×

4

Tw

o m

ice

hav

e m

ild p

iloer

ecti

on p

ost

day 1

2

P8

-A1

0.6

mg/k

g, q4d×

4

Thre

e m

ice

hav

e m

ild p

iloer

ecti

on p

ost

day 1

2

qd×

5 :

dosi

ng o

nce

dai

ly (

i.v.)

for

5 d

ays.

q4d

×4 :

dosi

ng o

nce

eac

h (

i.v.)

on d

ays

0, 4, 8 a

nd 1

2 (

4 d

ose

s).

243

T

ab

le 7

.4.

MT

D S

tud

y o

f L

iposo

mal

Form

ula

tion

s o

f P

8-A

1.

Form

ula

tion

D

osi

ng R

egim

en (

I.V

.)

Sig

ns

of

Tox

icit

y

Lip

o-1

0.8

mg/k

g

All

mic

e hav

e pil

oer

ecti

on p

ost

day

11

Lip

o-1

0.4

mg/k

g

No o

bvio

us

signs

of

tox

icit

y

Lip

o-1

0.4

mg/k

g, qd×

5

All

mic

e hav

e pil

oer

ecti

on p

ost

day

4

Lip

o-1

0.4

mg/k

g, q4d×

4

Thre

e m

ice

hav

e m

ild p

iloer

ecti

on p

ost

day12

Lip

o-1

0.6

mg/k

g, q4d×

4

All

mic

e hav

e pil

oer

ecti

on p

ost

day

12

Lip

o-2

0.4

mg/k

g, q4d×

4

No o

bvio

us

signs

of

tox

icit

y

Lip

o-2

0.8

mg/k

g, q4d×

4

Tw

o m

ice

had

mil

d p

iloer

ecti

on p

ost

day 1

2

Lip

o-2

1.6

mg/k

g, q4d×

4

Tw

o m

ice

hav

e pil

oer

ecti

on an

d hunch

ed post

ure

, an

d th

ree

mic

e hav

e pil

oer

ecti

on p

ost

day

0.

Lip

o-3

0.4

mg/k

g, q4d×

4

One

mouse

has

mil

d p

iloer

ecti

on p

ost

day 1

2

Lip

o-3

0.8

mg/k

g,

q4d×

4

All

mic

e hav

e pil

oer

ecti

on a

nd

hu

nch

ed p

ost

ure

po

st d

ay 2

4 a

nd

two m

ice

wer

e fo

und d

ead.

Lip

o-4

0.4

mg/k

g, q4d×

4

No o

bvio

us

signs

of

tox

icit

y

Lip

o-4

0.8

mg/k

g, q4d×

4

All

mic

e hav

e pil

oer

ecti

on a

nd h

unch

ed p

ost

ure

post

day

24

qd

×5

: d

osi

ng o

nce

dai

ly (

i.v.)

for

5 d

ays.

q4d

×4

: d

osi

ng o

nce

eac

h (

i.v.)

on d

ays

0, 4, 8 a

nd 1

2 (

4 d

ose

s).

244

Fig

ure

7.1

0.

Tum

or

volu

mes

± S

EM

(A

) an

d m

ou

se w

eights

(B

) in

A5

49

tum

or

bea

rin

g n

ude

mic

e re

cord

ed f

or

con

trol

(gro

up

1)

and t

reat

men

t gro

ups

(gro

ups

2–4).

M

ice

wer

e tr

eate

d o

n d

ays

0, 4, 8

, 12

, an

d 1

6 v

ia i

.v. in

ject

ions

into

the

tail

vei

n.

245

a Recovered from euthanized mice 14 days after administration of the last dose. Data are

ng Pt per organ ± SEM determined for 5 organs per group by ICP-MS. b Limit of

detection. * P < 0.05.

7.4. Experimental Section

7.4.1. Materials and instruments.

All reagents were used as obtained from commercial sources without further

purification unless indicated otherwise. HSPC and cholesterol were purchased from

Avanti Polar Lipids, Inc. DPPG and mPEG-DSPE were purchased from Lipoid.

Polycarbonate membranes with pore size of 100, 200 and 400 nm were purchased from

Millipore. Two extrusion devices were used in this study: 1-mL mini extruder (Avanti

Polar Lipids, Inc.) and 10-mL LIPEX™ Extruder (Northern Lipids Inc).

Table 7.5. Average ng Pt in Organsa

Liver Kidneys Lungs Spleen

control 0.4810.200 0.02780.0001b 0.04630.009 0.02780.0001b

P8-A1 1316226 17534 5.030.585 10.041.53

Lipo-1 617180* 55.919.8* 2.310.31* 28.48.2

Lipo-2 976183 15141 5.380.72 97.910.6

246

7.4.2. Optimization of the encapsulation conditions for P8-A1 using film hydration

method

Stock solutions (10 mg/mL) of each lipid were prepared by dissolving in a mixture of

chloroform and methanol with the proportion of 3:1 (V/V). To produce the formulations

with varying lipid compositions, various combinations of stock solutions, the volumes of

which were determined according to the desired molar ratio of lipids, were mixed in a 5-

mL round bottom flask to result in solutions containing 10 mg total lipids. The organic

solvent was then removed by rotary evaporator and subsequently dried in a oil-pump to

allow the formation of a thin lipid film. The lipid film was hydrated at 65 °C in 1 mL of

either Milli Q water or 0.9% saline containing 10 mg/mL P8-A1 for 30 min, followed by

ten freeze-thaw cycles using liquid nitrogen and a water bath (65 °C). The liposomal

suspension was repeatedly extruded at 65 °C through polycarbonate membranes with

pore sizes of 200 and 100 nm (10 times each) using a 1-mL mini extruder (Avanti Polar

Lipids). The non-encapsulated platinum complex was removed from the liposomal

suspension by gel permeation chromatography using a PD-10 Sephadex G-25 column

(GE Healthcare, Buckinghamshire, UK), eluted with 0.9% saline solution containing 5%

glucose.

7.4.3. Preparation of liposomes using ethanol injection method

The lipids DPPG, HSPC and mPEG-DSPE (80:15:5, molar ratio) were dissolved in 0.4

mL ethanol to produce the alcoholic solution of lipids with a concentration of 25 mg/mL.

Next, the ethanol solution was added slowly into 20-times volume of a 0.9% saline

solution containing 25 mg/mL P8-A1. The solution was stirred at room temperature for 1

247

h, followed by dialysis against 0.9% saline solution containing 5% glucose at 4 °C for 24

h. The liposomes formed spontaneously as the ethanol was completely removed after

dialysis. The liposomal suspension in water was then filtered through the 220 nm filter

membrane for sterilization, and stored at 4 °C.

7.4.4. Preparation of liposomes using reverse-phase evaporation method

DPPG, HSPC and mPEG2000-DSPE at a molar ratio of 80:15:5 (the total weight of

lipids is 10 mg) were dissolved in a mixture of chloroform and diethyl ether (1:2, v/v),

which was added into an aqueous solution containing 2 mg of P8-A1 dissolved in 0.9%

saline/5% glucose solution. The ratio of aqueous to organic phase was 1:3 (V/V). The

mixture was then sonicated at room temperature for 10 min, and the organic solvents

were slowly removed by rotary evaporation. The resultant aqueous suspension was

extruded with a mini extruder following the method described in section 7.4.2.

7.4.5. Preparation of liposomes for biological evaluation

Liposomes used for biological testing were prepared using the film hydration method as

described in section 7.4.2. The formulations were homogenized using a 10-mL LIPEX™

Extruder (Northern Lipids Inc.). The unbound platinum complex was removed from

liposomal suspension by gel permeation chromatography using a PD-10 Sephadex G-25

column (GE Healthcare, Buckinghamshire, UK), eluted with 0.9% saline/5% glucose

solution, followed by dialysis against 0.9% saline/5% glucose solution at 4 °C for 24

hour. The liposomal formulation was store at 4 °C when not in use.

248

7.4.6. Characterization of liposomes

7.4.6.1. Size and zeta-potential of liposomes

Liposome suspensions (~10 mg/mL lipids, 25 μL) were diluted in 1 mL of Milli-Q

water and filtered through a membrane filter with a pore size of 220 nm. The

hydrodynamic diameter was assessed by dynamic light scattering (DLS) using a Zetasizer

ZS90 (Malvern Instruments) in automatic mode (Software version 6.34) using the

intrument’s default settings for refractive index and viscosity of the dispersant. All

measurements were run at 25 °C. The zeta-potential of was assessed in Milli-Q water at

pH 6.8–7 using the Zetasizer ZS90 instrument. Triplicate measurements were performed.

The results are presented as paticle sizes and charges ± SD.

7.4.6.2. Transmission Electron Microscopy

The liposome samples were diluted before analysis with 0.9% saline and then pipetted

on carbon-stabilized Formvar copper grids. Excess sample was removed with a filter

paper. A drop of 2% (w/v) aqueous solution of uranyl acetate was added and the sample

was allowed to stain for 1 min. The samples were dried at room temperature and then

imaged using FEI Tecnai BioTwin Transmission Electron Microscope operated at 80 kV.

The images were captured with an AMT 2vu camera capable of 12 megapixel images and

analyzed with the included AMT camera software.

249

7.4.6.3. Encapsulation efficiency of liposomes

To determine the content of P8-A1 entrapped in the liposomes, methanol was used to

lyse the liposomal lipid bilayer. Briefly, 100 µL of liposomes before and after

purification were mixed with 1900 µL of methanol. The mixtures were sonicated for 15

min and the amount of platinum-acridine was determined spectrophotometrically at 413

nm (ɛ413 = 10000 M-1cm-1). Encapsulation efficiencies (EE %) were determined using the

following equation: EE% = ([Drug]purified/[Drug]initial) × 100%.

7.4.6.4. Determination of lipid composition

Phospholipids concentration was determined using Bartlett’s assay [392]. The assay

relies on the interaction between the inorganic phosphates and ammonium molybdate,

which can be monitored by reaction with 4-amino-3-hydroxyl-1-naphtalene sulfonic acid

at 180 °C to yield a blue-colored product and quantified at 815 nm. The amount of

cholesterol was quantification by reversed phase HPLC using a UV detector with the

wavelength at 207 nm. The samples were separated using a 4.6 mm × 150 mm reverse-

phase Agilent ZORBAX SB-C18 (5 mm) analytical column at 25 °C, by using the

following solvent system: solvent A, optima water, and solvent B, methanol/0.1 % formic

acid, with a flow rate of 0.5 mL/min and a gradient of 95 % A to 5 % A over 20 minutes.

250

7.4.7. In vitro stability assay

7.4.7.1. Drug leakage under physiologically relevant conditions

Liposomal formulations (~1 mg/ml lipids) were incubated in 1 Ml PBS (pH 7.4) at

37 °C. At selected time-points, samples (in triplicate) were eluted with 0.9% NaCl

solution by size exclusion chromatography using PD-10 Sephadex G-25 columns to

remove released free P8-A1. The amount of drug retened was then measured

spectroscopically as described in 7.4.5.2.

7.3.6.2. In vitro drug release of P8-A1 at elevated temperature

Lipo-1 (~1 mg/mL lipids) was incubated in 1 mL PBS (pH 7.4) at 42 °C and 44 °C. At

selected time-points, samples (in triplicate) were eluted with 0.9% NaCl solution by size

exclusion chromatography using PD-10 Sephadex G-25 columns to remove released free

P8-A1. The amount of drug retention was then measured spectroscopically as described

in 7.4.2.

7.4.6.3. Storage stability

Purified liposomal formulations (10 mL) were sterilized by passing them through 0.22

µm membrane filters and stored at 4 °C for 1, 4, and 8 weeks. Samples were tested for

particle size and drug retention within liposomes. The particle size was evaluated by DLS

measurements. To determine the concentration of encapsulated P8-A1, samples were

passed through PD-10 Sephadex G-25 columns and eluted with 0.9% saline solution to

251

remove any free platinum complex. The amount of drug retented was then measured

spectroscopically as described above. The samples before the storage were used as

controls. The liposomes lysed by methanol were used as positive controls. Measurements

were performed in triplicate.

7.4.8. Cellular uptake of Lipo-1 in NCI-H460

NCI-H460 cells were seeded into poly-d-lysine coated glass-bottom Petri dishes

(MatTeck Corporation, Ashland, MD, USA) with 105 cells mL−1 suspended in 2 mL of

medium per dish. Cells were incubated overnight and then treated with Lipo-1 and P8–

A1 (5 μm), or medium for controls for designed time points. After treatment, medium

was removed and the cells were stained with 75 nM of LysoTracker Red DND-99

(Invitrogen) in 2 mL pre-warmed Hank’s Balanced Salt Solution (HBSS) at 37 °C for 30

min. After staining with LysoTracker, solutions were replaced with fresh medium.

Images were collected using a Zeiss LSM 710 confocal microscope (Carl Zeiss

MicroImaging, Thornwood, NY) using a 63× (PLAN APO, 1.2 NA) objective lens. All

images were acquired in multi-track configuration mode to minimize excitation cross talk

and emission bleed-through. The fluorescence of P8–A1 (blue), and Lyso–Tracker Red

DND-99, was excited/collected at 405/477 nm and 561/591 nm, respectively. ZEN

software (blue edition, Carl Zeiss, 2011) was used for image processing. Confocal image

planes for each channel were not contrast-adjusted (unless indicated) or otherwise

changed. Panels were assembled and annotated without any additional manipulation of

images in Adobe Photoshop CS2.

252

7.4.9. Animal studies

7.4.9.1. Evaluation of maximum tolerated dose (MTD)

MTD of P8-A1 and liposomeal formulations were evaluated in Swiss-Webster female

mice (5 to 6 week-old, 22-25 g). Mice (5 per group) were treated i.v. with the doses

ranging from 0.2-0.8 mg/kg every four days for four times. The acute toxicity was

assessed based on animal survival and visual signs of toxicity including body weight loss

(BWL) and clinical signs. Clinical signs observed were piloerection or hunched posture.

Animals displaying mild piloerection were considered to be indicating mild toxicity at

that dose.

7.4.9.2. A549 Xenograft study

A549 xenografts were established in nude athymic female mice via subcutaneous

injections. Treatment began when the average tumor volume was approximately 100 mm3.

The tumor-bearing mice were randomized depending on tumor volume into three groups

of eight test animals each: one control group receiving vehicle only, P8-A1 group treated

at 0.4 mg/kg q4d×4, Lipo-1 group treated at 0.4 mg/kg q4d×4, and Lipo-2 group treated

at 0.8 mg/kg q4d×4. Animal weights and tumor volumes were measured and recorded for

17 days after the first dose was administered. Tumor volumes were determined using the

formula :V (mm3) = d2 × D/2, where d and D are the shortest and longest dimension of

the tumor, respectively, and are reported as the sum of both tumors for each test animal.

At the end of the study, all animals were euthanized and disposed off according to

Standard Operating Procedures (SOPs). One tumor per test animal as well as lungs,

253

spleens, and kidneys were removed during necropsy. The concentration of Pt in these

tissues was determined by quadrupole inductively-coupled mass spectrometry (ICP-MS)

on a Thermo ICP-MS instrument at the Analytic Chemistry Service facility of Research

Triangle Institute, Research Triangle Park, NC. The xenograft study was performed by

Washington Biotechnology Inc. (Simpsonville, MD), PHS-approved Animal Welfare

Assurance, # A4912-01. Data analysis for weight and tumor volumes was done with the

Student’s t-test. Only P values of less than 0.05% were considered as statistically

significant. Few outliers were identified using Grubb’s outlier test at 90% and 95%

confidence and removed from the data set.

254

CHAPTER 8

SUMMARY AND OUTLOOK

Despite the success of cisplatin as a mainstay chemotherapeutic agent, tumor resistance

and severe side effects have inspired the development of new generations of platinum-

based agents. Platinum complexes with entirely different structural features have been

proposed to produce DNA-platinum adducts that would be recognized differently by the

proteins involved in cisplatin-induced cell death. Based on this reasoning, numerous

non-classical platinum agents have been generated. Among these are platinum-acridine

hybrid agents, which represent a unique class of monofunctional platinum complexes

capable of simultaneously intercalating and platinating DNA. While platinum-acridine

derivatives have demonstrated superior cytotoxicity in human lung cancer models, their

application as systemic cancer treatments is limited by severe systemic toxicity. The goal

of this dissertation was to develop the chemistry and devise new strategies for improving

the pharmacological properties of these agents. These include the design of lipophilic and

less reactive prodrugs, targeted and nontargeted vehicles for safer delivery of cytotoxic

payload to diseased tissue, and multifunctional agents to overcome the chemoresistance

of certain cancers. To achieve these challenging goals, new chemistry-based

methodologies had to be developed that allow platinum-acridines to be turned into

“smarter” anticancer agents. In particular, platinum-compatible chemistry was needed

that allows functionalization of platinum-acridines and their subsequent conjugation with

targeting and/or other bioactive molecules.

255

In Chapter 2, we have presented a library approach for the discovery of functionalized

platinum-acridine analogues. This method was based on a highly efficient addition

reaction between a platinum-activated nitrile precursor and a secondary amine, which

results in an amidine linkage that connects the intercalator to the metal. 60 microscale

reactions were performed in parallel, characterized and tested in cell models, which

helped us identify several conjugatable warheads that maintained potent cytotoxicity.

This assay also allowed us to study structure-activity relationships in functionalized

platinum-acridines.

Next, we explored various conjugation chemistries. To produce platinum-acridine

based conjugates, two strategies, “post-modification” and “pre-assembly”, have been

validated in this dissertation. In Chapter 3, a “post-labeling” method using a clickable

azide-modified platinum-acridine was developed. In a proof-of-concept study we

exploited this type of compound, in conjunction with copper-based and copper-free click

chemistries, to investigate the subcellular distribution of this type of platinum complex.

Cancer cells were first incubated with the azide derivative, followed by labeling of

intracellular platinum compound with an alkyne-containing fluorophore. Unlike the bulky

fluorophore-tagged derivative, the clickable azide derivative faithfully mimics the

subcellular localization of the parent platinum-acridines. The nucleolar localization of

platinum-acridine was discovered in this study, which is a previously unexplored target

for platinum-based drugs. Although very efficient, click chemistry may not be a widely

applicable method due to expensive precursors and additional steps needed to incorporate

alkyne groups into the non-platinum components.

256

Chapter 4 provides an example of the “pre-assembly” method, in which the platinum

moiety was not incorporated until the sensitizer/tumor-targeted ligand had been

assembled in its entirety. In this study, we linked a hydroxyl-modified platinum-acridine

to the estrogen receptor inhibitor endoxifen with the goal of producing a breast cancer

targeted agent. Coupling was achieved by a chemically robust carbamate bond. To

generate the conjugate, endoxifen was tethered to 9-amino acridine precursor to produce

the organic ligand that was in turn added to the platinum-nitrile precursor. Our “pre-

assembly” method holds promise of providing a universal approach for constructing

small molecule based conjugates, but may not be useful to conjugate functionalized

platinum-acridine warheads to macromolecules, such as monoclonal antibodies, proteins,

and synthetic polymers. The endoxifen-platinum-acridine conjugates were tested in breast

cancer cells where they showed the same level of cytotoxicity as that observed for the

parent “warhead”. The non-cleavable carbamate linker was identified as a potential

drawback of the design and prompted us to also pursue cleavable platinum-acridine

conjugates.

Chapter 5 described ester-based platinum-acridine prodrugs with improved drug-like

properties. Butyric acid and valproic acid were installed to confer lipohilicity to

platinum-acridines, which was expected to improve their systemic circulation and tumor

penetration. The valproate ester was found to undergo a carboxylesterase 2 (CES-2)-

mediated prodrug activation. Because CES-2 is overexpressed in many cancers, prodrug-

activation may occur tumor-selectively. However, activation by ubiquitous CES-2 in the

liver also has to be considered a possible mechanism of drug activation. In addition to

enzymatic cleavage, a platinum-mediated ester hydrolysis was also identified, in which

257

platinum(II) is able to accelerate ester hydrolysis in a low-chloride environment, such as

the cytosol.

As a continuation of the ester conjugation chemistry developed in Chapter 5, we

designed a fragment-based platform for the discovery of multifunctional platinum-

acridine conjugates (Chapter 6). Hydroxyl-modified platinum-acridines were extended

with a succinate linker, which was then conjugated with bioactive molecules containing a

primary or secondary amino group. Because of their steric hindrance, the ester groups in

the conjugates should not serve as substrates of CES-2 (and other esterases). We

demonstrated that these esters undergo metal-facilitated hydrolysis similar to the prodrug

models developed in Chapter 5. The major goal of this project was to develop

methodology that allows post-modification of the carboxylic acid modified platinum-

acridines via formation of an amide bond. We exhaustively screened various coupling

reagents and conditions that allow clean conversion of precursors to the desired products

with yields of higher than 90%. A total of 20 microscale reactions were assembled to

demonstrate the versatility and utility of this reaction. The high efficiency of this method

may allow the screening of high-throughput libraries in the future. Similar coupling

chemistries may also be an option for generating other cleavable linkers (e.g., pH-

sensitive hydrazones) for advanced applications such as conjugates containing antibodies

or targeted peptides.

Attempts to physically encapsulate platinum-acridine in nanocarriers are reported in

Chapter 7. In this chapter, we have developed a method to effectively load platinum-

acridines into negatively charged liposomes by taking advantage of the cationic nature of

the platinum complexes. One formulation shows improved anticancer efficacy relative to

258

the unencapsulated drug. Importantly, the liposomal formulations also appear to reduce

the accumulation of platinum in off-target tissues. Future efforts will focus on increasing

the MTD of the liposomal formulations by optimizing the lipid components and drug

loading and release properties, and by incorporating tumor-targeting ligands into the lipid

bilayers.

In conclusion, the research in this dissertation has established various modular

conjugation chemistries, new drug release mechanisms, and nanocarrier-based delivery,

which may be useful for the development of next-generation platinum-acridine agents

and of metallodrugs in general. In this respect, highly efficient transformations, such as

the platinum-mediated nitrile-amine addition and amide coupling chemistries developed

in Chapters 2 and 6 will prove indispensible tools.

259

REFERENCES

[1] N.J. Wheate, S. Walker, G.E. Craig, R. Oun, The status of platinum anticancer drugs

in the clinic and in clinical trials, Dalton Transactions, 39 (2010) 8113-8127.

[2] Z.H. Siddik, Cisplatin: mode of cytotoxic action and molecular basis of resistance,

Oncogene, 22 (2003) 7265-7279.

[3] L. Kelland, The resurgence of platinum-based cancer chemotherapy, Nature Reviews

Cancer, 7 (2007) 573-584.

[4] A. Tesniere, F. Schlemmer, V. Boige, O. Kepp, I. Martins, F. Ghiringhelli, L.

Aymeric, M. Michaud, L. Apetoh, L. Barault, J. Mendiboure, J.P. Pignon, V. Jooste,

P. van Endert, M. Ducreux, L. Zitvogel, F. Piard, G. Kroemer, Immunogenic death

of colon cancer cells treated with oxaliplatin, Oncogene, 29 (2010) 482-491.

[5] A.R. de Biasi, J. Villena-Vargas, P.S. Adusumilli, Cisplatin-Induced Antitumor

Immunomodulation: A Review of Preclinical and Clinical Evidence, Clinical

Cancer Research, 20 (2014) 5384-5391.

[6] D.Y.Q. Wong, W.W.F. Ong, W.H. Ang, Induction of Immunogenic Cell Death by

Chemotherapeutic Platinum Complexes, Angewandte Chemie-International Edition,

54 (2015) 6483-6487.

[7] D.Y.Q. Wong, C.H.F. Yeo, W.H. Ang, Immuno-Chemotherapeutic Platinum(IV)

Prodrugs of Cisplatin as Multimodal Anticancer Agents, Angewandte Chemie-

International Edition, 53 (2014) 6752-6756.

260

[8] H.S. Oberoi, N.V. Nukolova, A.V. Kabanov, T.K. Bronich, Nanocarriers for delivery

of platinum anticancer drugs, Advanced Drug Delivery Reviews, 65 (2013) 1667-

1685.

[9] S. Dasari, P.B. Tchounwou, Cisplatin in cancer therapy: Molecular mechanisms of

action, European Journal of Pharmacology, 740 (2014) 364-378.

[10] M.D. Hall, M. Okabe, D.W. Shen, X.J. Liang, M.M. Gottesman, The role of cellular

accumulation in determining sensitivity to platinum-based chemotherapy, Annual

Review of Pharmacology and Toxicology, 48 (2008) 495-535.

[11] C.A. Larson, B.G. Blair, R. Safaei, S.B. Howell, The role of the mammalian copper

transporter 1 in the cellular accumulation of platinum-based drugs, Molecular

Pharmacology, 75 (2009) 324-330.

[12] S. Zhang, K.S. Lovejoy, J.E. Shima, L.L. Lagpacan, Y. Shu, A. Lapuk, Y. Chen, T.

Komori, J.W. Gray, X. Chen, S.J. Lippard, K.M. Giacomini, Organic cation

transporters are determinants of oxaliplatin cytotoxicity, Cancer Research, 66 (2006)

8847-8857.

[13] P. Mistry, L.R. Kelland, G. Abel, S. Sidhar, K.R. Harrap, The relationships between

glutathione, glutathione-S-transferase and cytotoxicity of platinum drugs and

melphalan in eight human ovarian carcinoma cell lines, British Journal of Cancer,

64 (1991) 215-220.

[14] A.D. Lewis, J.D. Hayes, C.R. Wolf, Glutathione and glutathione-dependent enzymes

in ovarian adenocarcinoma cell lines derived from a patient before and after the

onset of drug resistance: intrinsic differences and cell cycle effects, Carcinogenesis,

9 (1988) 1283-1287.

261

[15] K.V. Ferry, T.C. Hamilton, S.W. Johnson, Increased nucleotide excision repair in

cisplatin-resistant ovarian cancer cells - Role of ERCC1-XPF, Biochemical

Pharmacology, 60 (2000) 1305-1313.

[16] E. Reed, ERCC1 and clinical resistance to platinum-based therapy, Clinical Cancer

Research, 11 (2005) 6100-6102.

[17] A. Gadducci, S. Cosio, S. Muraca, A.R. Genazzani, Molecular mechanisms of

apoptosis and chemosensitivity to platinum and paclitaxel in ovarian cancer:

biological data and clinical implications, European Journal of Gynaecological

Oncology, 23 (2002) 390-396.

[18] M. Galanski, B.K. Keppler, Searching for the Magic Bullet: Anticancer Platinum

Drugs Which Can Be Accumulated or Activated in the Tumor Tissue, Anti-Cancer

Agents in Medicinal Chemistry, 7 (2007) 55-73.

[19] J.J. Wilson, S.J. Lippard, In Vitro Anticancer Activity of cis-Diammineplatinum(II)

Complexes with beta-Diketonate Leaving Group Ligands, Journal of Medicinal

Chemistry, 55 (2012) 5326-5336.

[20] S. Hochreuther, R. Puchta, R. van Eldik, Thermodynamic and Kinetic Studies on

Novel Dinuclear Platinum(II) Complexes Containing Bidentate N,N-donor ligands,

Inorganic Chemistry, 50 (2011) 8984-8996.

[21] M.S. Robillard, M. Bacac, H. van den Elst, A. Flamigni, G.A. van der Marel, J.H.

van Boom, J. Reedijk, Automated parallel solid-phase synthesis and anticancer

screening of a library of peptide-tethered platinum(II) complexes, Journal of

Combinatorial Chemistry, 5 (2003) 821-825.

262

[22] K. Brasseur, V. Leblanc, F. Fabi, S. Parent, C. Descoteaux, G. Berube, E. Asselin,

ER alpha-Targeted Therapy in Ovarian Cancer Cells by a Novel Estradiol-

Platinum(II) Hybrid, Endocrinology, 154 (2013) 2281-2295.

[23] M. Kvasnica, L. Rarova, J. Oklestkova, M. Budesinsky, L. Kohout, Synthesis and

cytotoxic activities of estrone and estradiol cis-dichloroplatinum(II) complexes,

Bioorganic & Medicinal Chemistry, 20 (2012) 6969-6978.

[24] P. Saha, C. Descoteaux, K. Brasseur, S. Fortin, V. Leblanc, S. Parent, E. Asselin, G.

Berube, Synthesis, antiproliferative activity and estrogen receptor alpha affinity of

novel estradiol-linked platinum(II) complex analogs to carboplatin and oxaliplatin.

Potential vector complexes to target estrogen-dependent tissues, European Journal

of Medicinal Chemistry, 48 (2012) 385-390.

[25] O. Aronov, A.T. Horowitz, A. Gabizon, D. Gibson, Folate-targeted PEG as a

potential carrier for carboplatin analogs. Synthesis and in vitro studies,

Bioconjugate Chemistry, 14 (2003) 563-574.

[26] S. Top, E.B. Kaloun, A. Vessieres, G. Leclercq, I. Laios, M. Ourevitch, C. Deuschel,

M.J. McGlinchey, G. Jaouen, Tamoxifen derivatives for delivery of the antitumoral

(DACH)Pt group: Selective synthesis by McMurry coupling, and biochemical

behaviour, ChemBioChem, 4 (2003) 754-761.

[27] U. Kalinowska-Lis, J. Ochocki, K. Matlawska-Wasowska, Trans geometry in

platinum antitumor complexes, Coordination Chemistry Reviews, 252 (2008)

1328-1345.

263

[28] V. Brabec, J. Kasparkova, Modifications of DNA by platinum complexes - Relation

to resistance of tumors to platinum antitumor drugs, Drug Resistance Updates, 8

(2005) 131-146.

[29] E. BernalMendez, M. Boudvillain, F. GonzalezVilchez, M. Leng, Chemical

versatility of transplatin monofunctional adducts within multiple site-specifically

platinated DNA, Biochemistry, 36 (1997) 7281-7287.

[30] T.G. Appleton, A.J. Bailey, K.J. Barnham, J.R. Hall, Aspects of the solution

chemistry of trans-diammineplatinum(II) complexes, Inorganic Chemistry, 31

(1992) 3077-3082.

[31] N. Farrell, L.R. Kelland, J.D. Roberts, M. Vanbeusichem, Activation of the

transgeometry in platinum antitumor complexes-a survey of the cytotoxicity of

transcomplexes containing planar ligands in murine-L1210 and human tumr panels

and studies on their mechanism of action, Cancer Research, 52 (1992) 5065-5072.

[32] O. Novakova, J. Kasparkova, J. Malina, G. Natile, V. Brabec, DNA-protein cross-

linking by trans-[PtCl(2)(E-iminoether)(2)]. A concept for activation of the trans

geometry in platinum antitumor complexes, Nucleic Acids Research, 31 (2003)

6450-6460.

[33] M. Frybortova, O. Novakova, J. Stepankova, V. Novohradsky, D. Gibson, J.

Kasparkova, V. Brabec, Activation of trans geometry in bifunctional mononuclear

platinum complexes by a non-bulky methylamine ligand, Journal of Inorganic

Biochemistry, 126 (2013) 46-54.

[34] F.J. Ramos-Lima, O. Vrana, A.G. Quiroga, C. Navarro-Ranninger, A. Halamikova,

H. Rybnickova, L. Hejmalova, V. Brabec, Structural characterization, DNA

264

interactions, and cytotoxicity of new transplatin analogues containing one aliphatic

and one planar heterocyclic amine ligand, Journal of Medicinal Chemistry, 49

(2006) 2640-2651.

[35] S.M. Aris, N.P. Farrell, Towards Antitumor Active trans-Platinum Compounds,

European Journal of Inorganic Chemistry, (2009) 1293-1302.

[36] T.C. Johnstone, J.J. Wilson, S.J. Lippard, Monofunctional and Higher-Valent

Platinum Anticancer Agents, Inorganic Chemistry, 52 (2013) 12234-12249.

[37] T. Ren, D.P. Bancroft, W.I. Sundquist, A. Masschelein, M.V. Keck, S.J. Lippard,

Synthesis,characterization, DNA-binding properties, and solution thermochromism

of platinum(II) complexes of the ethidinium cation-regiospecificity in a DNA-

promoted reaction, Journal of the American Chemical Society, 115 (1993) 11341-

11352.

[38] S. Wu, X. Wang, C. Zhu, Y. Song, J. Wang, Y. Li, Z. Guo, Monofunctional

platinum complexes containing a 4-nitrobenzo-2-oxa-1,3-diazole fluorophore:

Distribution in tumour cells, Dalton Transactions, 40 (2011) 10376-10382.

[39] L.S. Hollis, A.R. Amundsen, E.W. Stern, Chemical and biological properties of a

new series of cis-diammineplatinum(II) antitumors agents containing 3 nitrogen

donors- Cis-Pt(NH3)2(N-dornor)Cl, Journal of Medicinal Chemistry, 32 (1989)

128-136.

[40] W.I. Sundquist, D.P. Bancroft, S.J. Lippard, Synthesis, characterization , and

biological-activity of cis-diammineplatinum (II) complexes of the DNA

intercalators 9-aminoacridine and chloroquine Journal of the American Chemical

Society, 112 (1990) 1590-1596.

265

[41] K.S. Lovejoy, R.C. Todd, S. Zhang, M.S. McCormick, J.A. D'Aquino, J.T. Reardon,

A. Sancar, K.M. Giacomini, S.J. Lippard, cis-

diammine(pyridine)chloroplatinum(II), a monofunctional platinum(II) antitumor

agent: Uptake, structure, function, and prospects, Proceedings of the National

Academy of Sciences of the U.S.A., 105 (2008) 8902-8907.

[42] K.S. Lovejoy, M. Serova, I. Bieche, S. Emami, M. D'Incalci, M. Broggini, E. Erba,

C. Gespach, E. Cvitkovic, S. Faivre, E. Raymond, S.J. Lippard, Spectrum of

Cellular Responses to Pyriplatin, a Monofunctional Cationic Antineoplastic

Platinum(II) Compound, in Human Cancer Cells, Molecular Cancer Therapeutics,

10 (2011) 1709-1719.

[43] G. Zhu, M. Myint, W.H. Ang, L. Song, S.J. Lippard, Monofunctional Platinum-

DNAAdducts Are Strong Inhibitors of Transcription and Substrates for Nucleotide

Excision Repair in Live Mammalian Cells, Cancer Research, 72 (2012) 790-800.

[44] G.Y. Park, J.J. Wilson, Y. Song, S.J. Lippard, Phenanthriplatin, a monofunctional

DNA-binding platinum anticancer drug candidate with unusual potency and

cellular activity profile, Proceedings of the National Academy of Sciences of the

U.S.A., 109 (2012) 11987-11992.

[45] N. Margiotta, S. Savino, V. Gandin, C. Marzano, G. Natile, Monofunctional

Platinum(II) Complexes with Potent Tumor Cell Growth Inhibitory Activity: The

Effect of a Hydrogen-Bond Donor/Acceptor N-Heterocyclic Ligand,

ChemMedChem, 9 (2014) 1161-1168.

[46] M.W. Kellinger, G.Y. Park, J. Chong, S.J. Lippard, D. Wang, Effect of a

Monofunctional Phenanthriplatin-DNA Adduct on RNA Polymerase II

266

Transcriptional Fidelity and Translesion Synthesis, Journal of the American

Chemical Society, 135 (2013) 13054-13061.

[47] B. Wang, Z. Wang, F. Ai, W.K. Tang, G. Zhu, A monofunctional platinum(II)-based

anticancer agent from a salicylanilide derivative: Synthesis, antiproliferative

activity, and transcription inhibition, Journal of Inorganic Biochemistry, 142 (2015)

118-125.

[48] N.P. Farrell, Progress in platinum-derived drug development, Drugs of the Future,

37 (2012) 795-806.

[49] N. Farrell, Polynuclear platinum drugs, Metal Ions in Biolgical Systems, Vol 42:

Metal Complexes in Tumor Diagnosis and as Anticancer Agents, 42 (2004) 251-

296.

[50] J.B. Mangrum, N.P. Farrell, Excursions in polynuclear platinum DNA binding,

Chemical Communications, 46 (2010) 6640-6650.

[51] C. Manzotti, G. Pratesi, E. Menta, R. Di Domenico, E. Cavalletti, H.H. Fiebig, L.R.

Kelland, N. Farrell, D. Polizzi, R. Supino, G. Pezzoni, F. Zunino, BBR 3464: A

novel triplatinum complex, exhibiting a preclinical profile of antitumor efficacy

different from cisplatin, Clinical Cancer Research, 6 (2000) 2626-2634.

[52] A. Hegmans, Y. Qu, L.R. Kelland, J.D. Roberts, N. Farrell, Novel approaches to

polynuclear platinum pro-drugs. Selective release of cytotoxic platinum spermidine

species through hydrolytic cleavage of carbamates, Inorganic Chemistry, 40 (2001)

6108-6114.

267

[53] Y. Zhao, W. He, P. Shi, J. Zhu, L. Qiu, L. Lin, Z. Guo, A positively charged

trinuclear 3N-chelated monofunctional platinum complex with high DNA affinity

and potent cytotoxicity, Dalton Transactions, (2006) 2617-2619.

[54] L. Zerzankova, H. Kostrhunova, M. Vojtiskova, O. Novakova, T. Suchankova, M.

Lin, Z. Guo, J. Kasparkova, V. Brabec, Mechanistic insights into antitumor effects

of new dinuclear cis Pt-II complexes containing aromatic linkers, Biochemical

Pharmacology, 80 (2010) 344-351.

[55] S. Komeda, Y.-L. Lin, M. Chikuma, A Tetrazolato-Bridged Dinuclear Platinum(II)

Complex Exhibits Markedly High in vivo Antitumor Activity against Pancreatic

Cancer, ChemMedChem, 6 (2011) 987-990.

[56] S. Komeda, T. Moulaei, K.K. Woods, M. Chikuma, N.P. Farrell, L.D. Williams, A

third mode of DNA binding: Phosphate clamps by a polynuclear platinum complex,

Journal of the American Chemical Society, 128 (2006) 16092-16103.

[57] J. Malina, N.P. Farrell, V. Brabec, Substitution-Inert Trinuclear Platinum Complexes

Efficiently Condense/Aggregate Nucleic Acids and Inhibit Enzymatic Activity,

Angewandte Chemie-International Edition, 53 (2014) 12812-12816.

[58] B.T. Benedetti, E.J. Peterson, P. Kabolizadeh, A. Martinez, R. Kipping, N.P. Farrell,

Effects of Noncovalent Platinum Drug-Protein Interactions on Drug Efficacy: Use

of Fluorescent Conjugates as Probes for Drug Metabolism, Molecular

Pharmaceutics, 8 (2011) 940-948.

[59] W.S. El-Deiry, The role of p53 in chemosensitivity and radiosensitivity, Oncogene,

22 (2003) 7486-7495.

268

[60] A.L. Harris, J.J. Ryan, N. Farrell, Biological consequences of trinuclear platinum

complexes: comparison of {trans-PtCl(NH3)2}2mu-(trans-Pt(NH3)2)H2N(CH2)6-

NH)2) (4+) (BBR 3464) with its noncovalent congeners, Molecular Pharmacology,

69 (2006) 666-672.

[61] J.B. Mangrum, B.J. Engelmann, E.J. Peterson, J.J. Ryan, S.J. Berners-Price, N.P.

Farrell, A new approach to glycan targeting: enzyme inhibition by oligosaccharide

metalloshielding, Chemical Communications, 50 (2014) 4056-4058.

[62] E.J. Peterson, V.R. Menon, L. Gatti, R. Kipping, D. Dewasinghe, P. Perego, L.F.

Povirk, N.P. Farrell, Nucleolar Targeting by Platinum: p53-Independent Apoptosis

Follows rRNA Inhibition, Cell-Cycle Arrest, and DNA Compaction, Molecular

Pharmaceutics, 12 (2015) 287-297.

[63] J. Malina, N.P. Farrell, V. Brabec, DNA Condensing Effects and Sequence

Selectivity of DNA Binding of Antitumor Noncovalent Polynuclear Platinum

Complexes, Inorganic Chemistry, 53 (2014) 1662-1671.

[64] H. Baruah, C.L. Rector, S.M. Monnier, U. Bierbach, Mechanism of action of non-

cisplatin type DNA-targeted platinum anticancer agents: DNA interactions of novel

acridinylthioureas and their platinum conjugates, Biochemical Pharmacology, 64

(2002) 191-200.

[65] E.T. Martins, H. Baruah, J. Kramarczyk, G. Saluta, C.S. Day, G.L. Kucera, U.

Bierbach, Design, synthesis, and biological activity of a novel non-cisplatin-type

platinum-acridine pharmacophore, Journal of Medicinal Chemistry, 44 (2001)

4492-4496.

269

[66] H. Baruah, C.G. Barry, U. Bierbach, Platinum-intercalator conjugates: From DNA-

targeted cisplatin derivatives to adenine binding complexes as potential modulators

of gene regulation, Current Topics in Medicinal Chemistry, 4 (2004) 1537-1549.

[67] M.E. Budiman, R.W. Alexander, U. Bierbach, Unique base-step recognition by a

platinum-acridinylthiourea conjugate leads to a DNA damage profile

complementary to that of the anticancer drug cisplatin, Biochemistry, 43 (2004)

8560-8567.

[68] L.A. Graham, J. Suryadi, T.K. West, G.L. Kucera, U. Bierbach, Synthesis, Aqueous

Reactivity, and Biological Evaluation of Carboxylic Acid Ester-Functionalized

Platinum-Acridine Hybrid Anticancer Agents, Journal of Medicinal Chemistry, 55

(2012) 7817-7827.

[69] S.M. Hess, J.G. Anderson, U. Bierbach, A non-crosslinking platinum-acridine

hybrid agent shows enhanced cytotoxicity compared to clinical BCNU and

cisplatin in glioblastoma cells, Bioorganic & Medicinal Chemistry Letters, 15

(2005) 443-446.

[70] R. Guddneppanavar, G. Saluta, G.L. Kucera, U. Bierbach, Synthesis, biological

activity, and DNA-damage profile of platinum-threading intercalator conjugates

designed to target adenine, Journal of Medicinal Chemistry, 49 (2006) 3204-3214.

[71] S.M. Hess, A.M. Mounce, R.C. Sequeira, T.M. Augustus, M.C. Ackley, U. Bierbach,

Platinum-acridinylthiourea conjugates show cell line-specific cytotoxic

enhancement in H460 lung carcinoma cells compared to cisplatin, Cancer

Chemotherapy and Pharmacology, 56 (2005) 337-343.

270

[72] R. Guddneppanavar, U. Bierbach, Adenine-N3 in the DNA Minor Groove - An

Emerging Target for Platinum Containing Anticancer Pharmacophores, Anti-

Cancer Agents in Medicinal Chemistry, 7 (2007) 125-138.

[73] Z. Ma, G. Saluta, G.L. Kucera, U. Bierbach, Effect of linkage geometry on

biological activity in thiourea- and guanidine-substituted acridines and platinum-

acridines, Bioorganic & Medicinal Chemistry Letters, 18 (2008) 3799-3801.

[74] R. Guddneppanavar, J.R. Choudhury, A.R. Kheradi, B.D. Steen, G. Saluta, G.L.

Kucera, C.S. Day, U. Bierbach, Effect of the diamine nonleaving group in

platinum-acridinylthiourea conjugates on DNA damage and cytotoxicity, Journal of

Medicinal Chemistry, 50 (2007) 2259-2263.

[75] M.C. Ackley, C.G. Barry, A.M. Mounce, M.C. Farmer, B.E. Springer, C.S. Day,

M.W. Wright, S.J. Berners-Price, S.M. Hess, U. Bierbach, Structure-activity

relationships in platinum-acridinylthiourea conjugates: effect of the thiourea

nonleaving group on drug stability, nucleobase affinity, and in vitro cytotoxicity,

Journal of Biological Inorganic Chemistry, 9 (2004) 453-461.

[76] Z. Ma, L. Rao, U. Bierbach, Replacement of a Thiourea-S with an Amidine-NH

Donor Group in a Platinum-Acridine Antitumor Compound Reduces the Metal's

Reactivity with Cysteine Sulfur, Journal of Medicinal Chemistry, 52 (2009) 3424-

3427.

[77] J.R. Choudhury, L. Rao, U. Bierbach, Rates of intercalator-driven platination of

DNA determined by a restriction enzyme cleavage inhibition assay, Journal of

Biological Inorganic Chemistry, 16 (2011) 373-380.

271

[78] H. Baruah, C.S. Day, M.W. Wright, U. Bierbach, Metal-intercalator-mediated self-

association and one-dimensional aggregation in the structure of the excised major

DNA adduct of a platinum-acridine agent, Journal of the American Chemical

Society, 126 (2004) 4492-4493.

[79] X. Wang, Z. Guo, Targeting and delivery of platinum-based anticancer drugs,

Chemical Society Reviews, 42 (2013) 202-224.

[80] J.S. Butler, P.J. Sadler, Targeted delivery of platinum-based anticancer complexes,

Current Opinion in Chemical Biology, 17 (2013) 175-188.

[81] C.M. Dawidczyk, C. Kim, J.H. Park, L.M. Russell, K.H. Lee, M.G. Pomper, P.C.

Searson, State-of-the-art in design rules for drug delivery platforms: Lessons

learned from FDA-approved nanomedicines, Journal of Controlled Release, 187

(2014) 133-144.

[82] Y. Barenholz, Doxil (R) - The first FDA-approved nano-drug: Lessons learned,

Journal of Controlled Release, 160 (2012) 117-134.

[83] A.A. Gabizon, H. Shmeeda, S. Zalipsky, Pros and cons of the liposome platform in

cancer drug targeting, Journal of Liposome Research, 16 (2006) 175-183.

[84] T.M. Allen, P.R. Cullis, Liposomal drug delivery systems: From concept to clinical

applications, Advanced Drug Delivery Reviews, 65 (2013) 36-48.

[85] G. Riess, Micellization of block copolymers, Progress in Polymer Science, 28 (2003)

1107-1170.

[86] V.P. Torchilin, Structure and design of polymeric surfactant-based drug delivery

systems, Journal of Controlled Release, 73 (2001) 137-172.

272

[87] C. Allen, D. Maysinger, A. Eisenberg, Nano-engineering block copolymer

aggregates for drug delivery, Colloids and Surfaces B: Biointerfaces, 16 (1999) 3-

27.

[88] X. Li, R. Lie, X. Qian, Y. Ding, Y. Tu, R. Guo, Y. Hu, X. Jiang, W. Guo, B. Liu,

Superior antitumor efficiency of cisplatin-loaded nanoparticles by intratumoral

delivery with decreased tumor metabolism rate, European Journal of Pharmaceutics

and Biopharmaceutics, 70 (2008) 726-734.

[89] T.C. Johnstone, S.J. Lippard, The Effect of Ligand Lipophilicity on the Nanoparticle

Encapsulation of Pt(IV) Prodrugs, Inorganic Chemistry, 52 (2013) 9915-9920.

[90] A. Lavasanifar, J. Samuel, G.S. Kwon, Poly(ethylene oxide)-block-poly(L-amino

acid) micelles for drug delivery, Advanced Drug Delivery Reviews, 54 (2002) 169-

190.

[91] N. Nishiyama, K. Kataoka, Preparation and characterization of size-controlled

polymeric micelle containing cis-dichlorodiammineplatinum(II) in the core, Journal

of Controlled Release, 74 (2001) 83-94.

[92] S. Bontha, A.V. Kabanov, T.K. Bronich, Polymer micelles with cross-linked ionic

cores for delivery of anticancer drugs, Journal of Controlled Release, 114 (2006)

163-174.

[93] H.S. Oberoi, F.C. Laquer, L.A. Marky, A.V. Kabanov, T.K. Bronich, Core cross-

linked block ionomer micelles as pH-responsive carriers for cis-

diamminedichloroplatinum(II), Journal of Controlled Release, 153 (2011) 64-72.

273

[94] N. Nishiyama, Y. Kato, Y. Sugiyama, K. Kataoka, Cisplatin-loaded polymer-metal

complex micelle with time-modulated decaying property as a novel drug delivery

system, Pharmaceutical Research, 18 (2001) 1035-1041.

[95] H. Ringsdorf, Structure and propertie of pharmacologically active polymers, Journal

of Polymer Science Part C-Polymer Symposium, (1975) 135-153.

[96] J. Kopecek, P. Kopeckova, T. Minko, Z.R. Lu, HPMA copolymer-anticancer drug

conjugates: design, activity, and mechanism of action, European Journal of

Pharmaceutics and Biopharmaceutics, 50 (2000) 61-81.

[97] R. Satchi-Fainaro, R. Duncan, C.M. Barnes, Polymer therapeutics for cancer:

Current status and future challenges, Polymer Therapeutics Ii: Polymers as Drugs,

Conjugates and Gene Delivery Systems2006, pp. 1-65.

[98] K.J. Haxton, H.M. Burt, Polymeric Drug Delivery of Platinum-Based Anticancer

Agents, Journal of Pharmaceutical Sciences, 98 (2009) 2299-2316.

[99] P. Sood, K.B. Thurmond, II, J.E. Jacob, L.K. Waller, G.O. Silva, D.R. Stewart, D.P.

Nowotnik, Synthesis and characterization of AP5346, a novel polymer-linked

diaminocyclohexyl platinum chemotherapeutic agent, Bioconjugate Chemistry, 17

(2006) 1270-1279.

[100] J.R. Rice, J.L. Gerberich, D.P. Nowotnik, S.B. Howell, Preclinical efficacy and

pharmacokinetics of AP5346, a novel diaminocyclohexane-platinum tumor-

targeting drug delivery system, Clinical Cancer Research, 12 (2006) 2248-2254.

[101] D. Stewart, J. Rice, P. Sood, J.S. John, S. Shevchuk, K.B. Thurmond, D. Nguyen,

G. Russell-Jones, D. Nowotnik, Preclinical development of AP5346: A

dachplatinum polymer conjugate, Journal of Controlled Release, 91 (2003) 255-256.

274

[102] D.P. Nowotnik, E. Cvitkovic, ProLindac (TM) (AP5346): A review of the

development of an HPMA DACH platinum Polymer Therapeutic, Advanced Drug

Delivery Reviews, 61 (2009) 1214-1219.

[103] Y.S. Sohn, Y.H. Cho, H. Baek, O.S. Jung, Synthesis and properties of low-

molecular-weight polyphosphazenes, Macromolecules, 28 (1995) 7566-7568.

[104] M. Ravera, E. Gabano, S. Bianco, G. Ermondi, G. Caron, M. Vallaro, G. Pelosi, I.

Zanellato, I. Bonarrigo, C. Cassino, D. Osella, Host-guest inclusion systems of

Pt(IV)-bis(benzoato) anticancer drug candidates and cyclodextrins, Inorganica

Chimica Acta, 432 (2015) 115-127.

[105] E.W. Neuse, G. Caldwell, A.G. Perlwitz, Carrier polymers for cisplatin-type

anticancer drug models, Polymers for Advanced Technologies, 7 (1996) 867-872.

[106] M. Li, Z. Tang, S. Lv, W. Song, H. Hong, X. Jing, Y. Zhang, X. Chen, Cisplatin

crosslinked pH-sensitive nanoparticles for efficient delivery. of doxorubicin,

Biomaterials, 35 (2014) 3851-3864.

[107] C.C. Lee, J.A. MacKay, J.M.J. Frechet, F.C. Szoka, Designing dendrimers for

biological applications, Nature Biotechnology, 23 (2005) 1517-1526.

[108] B.A. Howell, D. Fan, Poly(amidoamine) dendrimer-supported organoplatinum

antitumour agents, Proceedings of the Royal Society a-Mathematical Physical and

Engineering Sciences, 466 (2010) 1515-1526.

[109] V.K. Yellepeddi, A. Kumar, D.M. Maher, S.C. Chauhan, K.K. Vangara, S.

Palakurthi, Biotinylated PAMAM Dendrimers for Intracellular Delivery of

Cisplatin to Ovarian Cancer: Role of SMVT, Anticancer Research, 31 (2011) 897-

906.

275

[110] T. Kapp, A. Dullin, R. Gust, Platinum(II)-Dendrimer Conjugates: Synthesis and

Investigations on Cytotoxicity, Cellular Distribution, Platinum Release, DNA, and

Protein Binding, Bioconjugate Chemistry, 21 (2010) 328-337.

[111] G. Bozzuto, A. Molinari, Liposomes as nanomedical devices, International Journal

of Nanomedicine, 10 (2015) 975-999.

[112] B. Ozpolat, A.K. Sood, G. Lopez-Berestein, Liposomal siRNA nanocarriers for

cancer therapy, Advanced Drug Delivery Reviews, 66 (2014) 110-116.

[113] V.V. Mody, M.I. Nounou, M. Bikram, Novel nanomedicine-based MRI contrast

agents for gynecological malignancies, Advanced Drug Delivery Reviews, 61

(2009) 795-807.

[114] N. Hijnen, S. Langereis, H. Grull, Magnetic resonance guided high-intensity

focused ultrasound for image-guided temperature-induced drug delivery, Advanced

Drug Delivery Reviews, 72 (2014) 65-81.

[115] K.N.J. Burger, R. Staffhorst, H.C. de Vijlder, M.J. Velinova, P.H. Bomans, P.M.

Frederik, B. de Kruijff, Nanocapsules: lipid-coated aggregates of cisplatin with

high cytotoxicity, Nature Medicine, 8 (2002) 81-84.

[116] Z. Zhang, L. Mei, S.-S. Feng, Paclitaxel drug delivery systems, Expert Opinion on

Drug Delivery, 10 (2013) 325-340.

[117] S.H. Alavizadeh, A. Badiee, S. Golmohammadzadeh, M.R. Jaafari, The influence

of phospholipid on the physicochemical properties and anti-tumor efficacy of

liposomes encapsulating cisplatin in mice bearing C26 colon carcinoma,

International Journal of Pharmaceutics, 473 (2014) 326-333.

276

[118] A.S. Abu Lila, S. Kizuki, Y. Doi, T. Suzuki, T. Ishida, H. Kiwada, Oxaliplatin

encapsulated in PEG-coated cationic liposomes induces significant tumor growth

suppression via a dual-targeting approach in a murine solid tumor model, Journal of

Controlled Release, 137 (2009) 8-14.

[119] S. Zalba, I. Navarro, I.F. Troconiz, C. Tros de Ilarduya, M.J. Garrido, Application

of different methods to formulate PEG-liposomes of oxaliplatin: Evaluation in vitro

and in vivo, European Journal of Pharmaceutics and Biopharmaceutics, 81 (2012)

273-280.

[120] M. Hirai, H. Minematsu, Y. Hiramatsu, H. Kitagawa, T. Otani, S. Iwashita, T.

Kudoh, L. Chen, Y. Li, M. Okada, D.S. Salomon, K. Igarashi, M. Chikuma, M.

Seno, Novel and simple loading procedure of cisplatin into liposomes and targeting

tumor endothelial cells, International Journal of Pharmaceutics, 391 (2010) 274-

283.

[121] I. Vhora, N. Khatri, J. Desai, H.P. Thakkar, Caprylate-Conjugated Cisplatin for the

Development of Novel Liposomal Formulation, AAPS PharmSciTech, 15 (2014)

845-857.

[122] Q. Li, Y. Tian, D. Li, J. Sun, D. Shi, L. Fang, Y. Gao, H. Liu, The effect of

lipocisplatin on cisplatin efficacy and nephrotoxicity in malignant breast cancer

treatment, Biomaterials, 35 (2014) 6462-6472.

[123] Y. Kuang, J. Liu, Z. Liu, R. Zhuo, Cholesterol-based anionic long-circulating

cisplatin liposomes with reduced renal toxicity, Biomaterials, 33 (2012) 1596-1606.

277

[124] L.M. Ickenstein, M.C. Arfvidsson, D. Needham, L.D. Mayer, K. Edwards, Disc

formation in cholesterol-free lippsornes during phase transition, Biochimica Et

Biophysica Acta-Biomembranes, 1614 (2003) 135-138.

[125] J. Gubernator, G. Chwastek, M. Korycinska, M. Stasiuk, G. Grynkiewicz, F.

Lewrick, R. Suess, A. Kozubek, The encapsulation of idarubicin within liposomes

using the novel EDTA ion gradient method ensures improved drug retention in

vitro and in vivo, Journal of Controlled Release, 146 (2010) 68-75.

[126] R.M. Abra, R.B. Bankert, F. Chen, N.K. Egilmez, K. Huang, R. Saville, J.L. Slater,

M. Sugano, S.J. Yokota, The next generation of liposome delivery systems: Recent

experience with tumor-targeted, sterically-stabilized immunoliposomes and active-

loading gradients, Journal of Liposome Research, 12 (2002) 1-3.

[127] I.M. Hann, H.G. Prentice, Lipid-based amphotericin B: a review of the last 10 years

of use, International Journal of Antimicrobial Agents, 17 (2001) 161-169.

[128] C.E. Petre, D.P. Dittmer, Liposomal daunorubicin as treatment for Kaposi's

sarcoma, International Journal of Nanomedicine, 2 (2007) 277-288.

[129] G. Batist, G. Ramakrishnan, C.S. Rao, A. Chandrasekharan, J. Gutheil, T. Guthrie,

P. Shah, A. Khojasteh, M.K. Nair, K. Hoelzer, K. Tkaczuk, Y.C. Park, L.W. Lee, G.

Myocet Study, Reduced cardiotoxicity and preserved antitumor efficacy of

liposome-encapsulated doxorubicin and cyclophosphamide compared with

conventional doxorubicin and cyclophosphamide in a randomized, multicenter trial

of metastatic breast cancer, Journal of Clinical Oncology, 19 (2001) 1444-1454.

[130] M.J. Glantz, K.A. Jaeckle, M.C. Chamberlain, S. Phuphanich, L. Recht, L.J.

Swinnen, B. Maria, S. LaFollette, G.B. Schumann, B.F. Cole, S.B. Howell, A

278

randomized controlled trial comparing intrathecal sustained-release cytarabine

(DepoCyt) to intrathecal methotrexate in patients with neoplastic meningitis from

solid tumors, Clinical Cancer Research, 5 (1999) 3394-3402.

[131] V. D'Acremont, C. Herzog, B. Genton, Immunogenicity and safety of a virosomal

hepatitis A vaccine (Epaxal (R)) in the elderly, Journal of Travel Medicine, 13

(2006) 78-83.

[132] C. Herzog, K. Hartmann, V. Kuenzi, O. Kuersteiner, R. Mischler, H. Lazar, R.

Glueck, Eleven years of Inflexal (R) V-a virosomal adjuvanted influenza vaccine,

Vaccine, 27 (2009) 4381-4387.

[133] G.P. Stathopoulos, T. Boulikas, Lipoplatin formulation review article, Journal of

drug delivery, 2012 (2012) 581363-581363.

[134] M.L. Immordino, F. Dosio, L. Cattel, Stealth liposomes: review of the basic science,

rationale, and clinical applications, existing and potential, International Journal of

Nanomedicine, 1 (2006) 297-315.

[135] T. Dragovich, D. Mendelson, S. Kurtin, K. Richardson, D. Von Hoff, A. Hoos, A

phase 2 trial of the liposomal DACH platinum L-NDDP in patients with therapy-

refractory advanced colorectal cancer, Cancer Chemotherapy and Pharmacology,

58 (2006) 759-764.

[136] G.P. Stathopoulos, T. Boulikas, A. Kourvetaris, J. Stathopoulos, Liposomal

oxaliplatin in the treatment of advanced cancer: A phase I study, Anticancer

Research, 26 (2006) 1489-1493.

[137] K.K. Sankhala, A.C. Mita, R. Adinin, L. Wood, M. Beeram, S. Bullock, N.

Yamagata, K. Matsuno, T. Fujisawa, A. Phan, A phase I pharmacokinetic (PK)

279

study of MBP-426, a novel liposome encapsulated oxaliplatin, Journal of Clinical

Oncology, 27 (2009).

[138] D. Liu, C. He, A.Z. Wang, W. Lin, Application of liposomal technologies for

delivery of platinum analogs in oncology, International Journal of Nanomedicine, 8

(2013) 3309-3319.

[139] T. Daemen, G. Hofstede, M.T.T. Kate, I. Bakkerwoudenberg, G.L. Scherphof,

Liposomal doxorubicin-induced toxicity-depletion and impairment of phagocytic-

activity of liver macrophages, International Journal of Cancer, 61 (1995) 716-721.

[140] M.C. Woodle, K.K. Matthay, M.S. Newman, J.E. Hidayat, L.R. Collins, C.

Redemann, F.J. Martin, D. Papahadjopoulos, Versatility in lipid compositions

showing prolonged circulation with sterically stabilized liposomes, Biochimica Et

Biophysica Acta-Biomembranes, 1105 (1992) 193-200.

[141] P.I. Campbell, Toxicity of some charged lipids used in liposome preparations,

Cytobios, 37 (1983) 21-26.

[142] A.S. Abu Lila, H. Kiwada, T. Ishida, The accelerated blood clearance (ABC)

phenomenon: Clinical challenge and approaches to manage, Journal of Controlled

Release, 172 (2013) 38-47.

[143] B.S. Wong, S.L. Yoong, A. Jagusiak, T. Panczyk, H.K. Ho, W.H. Ang, G. Pastorin,

Carbon nanotubes for delivery of small molecule drugs, Advanced Drug Delivery

Reviews, 65 (2013) 1964-2015.

[144] C.R. Martin, P. Kohli, The emerging field of nanotube biotechnology, Nature

Reviews Drug Discovery, 2 (2003) 29-37.

280

[145] S.-T. Yang, X. Wang, G. Jia, Y. Gu, T. Wang, H. Nie, C. Ge, H. Wang, Y. Liu,

Long-term accumulation and low toxicity of single-walled carbon nanotubes in

intravenously exposed mice, Toxicology Letters, 181 (2008) 182-189.

[146] A. Guven, I.A. Rusakova, M.T. Lewis, L.J. Wilson, Cisplatin@US-tube carbon

nanocapsules for enhanced chemotherapeutic delivery, Biomaterials, 33 (2012)

1455-1461.

[147] J. Kolosnjaj-Tabi, K.B. Hartman, S. Boudjemaa, J.S. Ananta, G. Morgant, H.

Szwarc, L.J. Wilson, F. Moussa, In Vivo Behavior of Large Doses of Ultrashort

and Full-Length Single-Walled Carbon Nanotubes after Oral and Intraperitoneal

Administration to Swiss Mice, ACS Nano, 4 (2010) 1481-1492.

[148] S.C. Tsang, Y.K. Chen, P.J.F. Harris, M.L.H. Green, A simple chemical method of

opening and filling carbon nanotubes, Nature, 372 (1994) 159-162.

[149] J. Li, S.Q. Yap, C.F. Chin, Q. Tian, S.L. Yoong, G. Pastorin, W.H. Ang,

Platinum(IV) prodrugs entrapped within multiwalled carbon nanotubes: Selective

release by chemical reduction and hydrophobicity reversal, Chemical Science, 3

(2012) 2083-2087.

[150] R.P. Feazell, N. Nakayama-Ratchford, H. Dai, S.J. Lippard, Soluble single-walled

carbon nanotubes as longboat delivery systems for Platinum(IV) anticancer drug

design, Journal of the American Chemical Society, 129 (2007) 8438-8439.

[151] J. Li, S.Q. Yap, S.L. Yoong, T.R. Nayak, G.W. Chandra, W.H. Ang, T. Panczyk, S.

Ramaprabhu, S.K. Vashist, F.-S. Sheu, A. Tan, G. Pastorin, Carbon nanotube

bottles for incorporation, release and enhanced cytotoxic effect of cisplatin, Carbon,

50 (2012) 1625-1634.

281

[152] L. Lacerda, J. Russier, G. Pastorin, M. Antonia Herrero, E. Venturelli, H.

Dumortier, K.T. Al-Jamal, M. Prato, K. Kostarelos, A. Bianco, Translocation

mechanisms of chemically functionalised carbon nanotubes across plasma

membranes, Biomaterials, 33 (2012) 3334-3343.

[153] S. Matsumura, K. Ajima, M. Yudasaka, S. Iijima, K. Shiba, Dispersion of cisplatin-

loaded carbon nanohorns with a conjugate comprised of an artificial peptide

aptamer and polyethylene glycol, Molecular Pharmaceutics, 4 (2007) 723-729.

[154] S. Dhar, Z. Liu, J. Thomale, H. Dai, S.J. Lippard, Targeted Single-Wall Carbon

Nanotube-Mediated Pt(IV) Prodrug Delivery Using Folate as a Homing Device,

Journal of the American Chemical Society, 130 (2008) 11467-11476.

[155] A.A. Bhirde, V. Patel, J. Gavard, G. Zhang, A.A. Sousa, A. Masedunskas, R.D.

Leapman, R. Weigert, J.S. Gutkind, J.F. Rusling, Targeted Killing of Cancer Cells

in Vivo and in Vitro with EGF-Directed Carbon Nanotube-Based Drug Delivery,

ACS Nano, 3 (2009) 307-316.

[156] J. Liu, Y. Huang, A. Kumar, A. Tan, S. Jin, A. Mozhi, X.-J. Liang, pH-Sensitive

nano-systems for drug delivery in cancer therapy, Biotechnology Advances, 32

(2014) 693-710.

[157] S. Aryal, C.-M.J. Hu, L. Zhang, Polymer−Cisplatin Conjugate Nanoparticles for

Acid-Responsive Drug Delivery, ACS Nano, 4 (2009) 251-258.

[158] T. Thambi, V.G. Deepagan, H.Y. Yoon, H.S. Han, S.-H. Kim, S. Son, D.-G. Jo, C.-

H. Ahn, Y.D. Suh, K. Kim, I. Chan Kwon, D.S. Lee, J.H. Park, Hypoxia-

responsive polymeric nanoparticles for tumor-targeted drug delivery, Biomaterials,

35 (2014) 1735-1743.

282

[159] F. Perche, S. Biswas, T. Wang, L. Zhu, V.P. Torchilin, Hypoxia-Targeted siRNA

Delivery, Angewandte Chemie International Edition, 53 (2014) 3362-3366.

[160] Y. Kageyama, Y. Yamazaki, H. Okuno, Novel approaches to prodrugs of

anticancer diaminodichloroplatinum(II) complexes activated by stereoselective

enzymatic ester hydrolysis, Journal of Inorganic Biochemistry, 70 (1998) 25-32.

[161] J.R. Tauro, B.-S. Lee, S.S. Lateef, R.A. Gemeinhart, Matrix metalloprotease

selective peptide substrates cleavage within hydrogel matrices for cancer

chemotherapy activation, Peptides, 29 (2008) 1965-1973.

[162] M.A. Fuertes, C. Alonso, J.M. Pérez, Biochemical Modulation of Cisplatin

Mechanisms of Action:  Enhancement of Antitumor Activity and Circumvention of

Drug Resistance, Chemical Reviews, 103 (2003) 645-662.

[163] J. Reedijk, Metal-Ligand Exchange Kinetics in Platinum and Ruthenium

Complexes significance for effectiveness as anticancer drugs, Platinum Metals

Review, 52 (2008) 2-11.

[164] E. Gabano, M. Ravera, D. Osella, The Drug Targeting and Delivery Approach

Applied to Pt-Antitumour Complexes. A Coordination Point of View, Current

Medicinal Chemistry, 16 (2009) 4544-4580.

[165] J. Reedijk, Why Does Cisplatin Reach Guanine-N7 with Competing S-Donor

Ligands Available in the Cell?, Chemical Reviews, 99 (1999) 2499-2510.

[166] G. Ma, Y. Min, F. Huang, T. Jiang, Y. Liu, Thioether binding mediates

monofunctional platinum antitumor reagents to trans configuration in DNA

interactions, Chemical Communications, 46 (2010) 6938-6940.

283

[167] C. Li, R. Huang, Y. Ding, E. Sletten, F. Arnesano, M. Losacco, G. Natile, Y. Liu,

Effect of Thioethers on DNA Platination by trans-Platinum Complexes, Inorganic

Chemistry, 50 (2011) 8168-8176.

[168] M.E. Alberto, N. Russo, Methionine ligand selectively promotes monofunctional

adducts between trans-EE platinum anticancer drug and guanine DNA base,

Chemical Communications, 47 (2011) 887-889.

[169] S.S.G.E. van Boom, B.W. Chen, J.M. Teuben, J. Reedijk, Platinum−Thioether

Bonds Can Be Reverted by Guanine−N7 Bonds in Pt(dien)2+ Model Adducts,

Inorganic Chemistry, 38 (1999) 1450-1455.

[170] C. Li, Z. Li, E. Sletten, F. Arnesano, M. Losacco, G. Natile, Y. Liu, Methionine

Can Favor DNA Platination by trans-Coordinated Platinum Antitumor Drugs,

Angewandte Chemie International Edition, 48 (2009) 8497-8500.

[171] F. Arnesano, S. Scintilla, G. Natile, Interaction between Platinum Complexes and a

Methionine Motif Found in Copper Transport Proteins, Angewandte Chemie

International Edition, 46 (2007) 9062-9064.

[172] Y. Kasherman, S. Sturup, D. Gibson, Trans labilization of am(m)ine ligands from

platinum(II) complexes by cancer cell extracts, Journal of Biological Inorganic

Chemistry, 14 (2009) 387-399.

[173] J. Mendez-Arroyo, J. Barroso-Flores, A.M. Lifschitz, A.A. Sarjeant, C.L. Stern,

C.A. Mirkin, A Multi-State, Allosterically-Regulated Molecular Receptor With

Switchable Selectivity, Journal of the American Chemical Society, 136 (2014)

10340-10348.

284

[174] D. Pan, W. she, C. Guo, K. Luo, Q. Yi, Z. Gu, PEGylated dendritic

diaminocyclohexyl-platinum (II) conjugates as pH-responsive drug delivery

vehicles with enhanced tumor accumulation and antitumor efficacy, Biomaterials,

35 (2014) 10080-10092.

[175] H. Wu, H. Cabral, K. Toh, P. Mi, Y.-C. Chen, Y. Matsumoto, N. Yamada, X. Liu,

H. Kinoh, Y. Miura, M.R. Kano, H. Nishihara, N. Nishiyama, K. Kataoka,

Polymeric micelles loaded with platinum anticancer drugs target preangiogenic

micrometastatic niches associated with inflammation, Journal of Controlled

Release, 189 (2014) 1-10.

[176] U. Frey, J.D. Ranford, P.J. Sadler, Ring-opening reactions of the anticancer drug

carboplatin: NMR characterization of cis-[Pt(NH3)2(CBDCA-O)(5'-GMP-N7)] in

solution, Inorganic Chemistry, 32 (1993) 1333-1340.

[177] D.P. Bancroft, C.A. Lepre, S.J. Lippard, Platinum-195 NMR kinetic and

mechanistic studies of cis- and trans-diamminedichloroplatinum(II) binding to

DNA, Journal of the American Chemical Society, 112 (1990) 6860-6871.

[178] M.E.M. Dolman, K.M.A. van Dorenmalen, E.H.E. Pieters, M. Lacombe, J. Pato, G.

Storm, W.E. Hennink, R.J. Kok, Imatinib-ULS-lysozyme: A proximal tubular cell-

targeted conjugate of imatinib for the treatment of renal diseases, Journal of

Controlled Release, 157 (2012) 461-468.

[179] V.T. Huynh, J.Y. Quek, P.L. de Souza, M.H. Stenzel, Block Copolymer Micelles

with Pendant Bifunctional Chelator for Platinum Drugs: Effect of Spacer Length on

the Viability of Tumor Cells, Biomacromolecules, 13 (2012) 1010-1023.

285

[180] S.-M. Lee, T.V. O’Halloran, S.T. Nguyen, Polymer-Caged Nanobins for

Synergistic Cisplatin−Doxorubicin Combination Chemotherapy, Journal of the

American Chemical Society, 132 (2010) 17130-17138.

[181] N.P.E. Barry, P.J. Sadler, Challenges for Metals in Medicine: How

Nanotechnology May Help To Shape the Future, ACS Nano, 7 (2013) 5654-5659.

[182] N. Nishiyama, S. Okazaki, H. Cabral, M. Miyamoto, Y. Kato, Y. Sugiyama, K.

Nishio, Y. Matsumura, K. Kataoka, Novel Cisplatin-Incorporated Polymeric

Micelles Can Eradicate Solid Tumors in Mice, Cancer Research, 63 (2003) 8977-

8983.

[183] W. Cao, Y. Gu, M. Meineck, T. Li, H. Xu, Tellurium-Containing Polymer Micelles:

Competitive-Ligand-Regulated Coordination Responsive Systems, Journal of the

American Chemical Society, 136 (2014) 5132-5137.

[184] M. Johnson, E. Neuse, C.J. van Rensburg, E. Kreft, Cell Growth-Inhibiting

Properties of Selected Carrier-Bound, Monoamine-Coordinated Platinum(II)

Compounds, Journal of Inorganic and Organometallic Polymers, 13 (2003) 55-67.

[185] M. Callari, J.R. Aldrich-Wright, P.L. de Souza, M.H. Stenzel, Polymers with

platinum drugs and other macromolecular metal complexes for cancer treatment,

Progress in Polymer Science, 39 (2014) 1614-1643.

[186] G.J. Kirkpatrick, J.A. Plumb, O.B. Sutcliffe, D.J. Flint, N.J. Wheate, Evaluation of

anionic half generation 3.5–6.5 poly(amidoamine) dendrimers as delivery vehicles

for the active component of the anticancer drug cisplatin, Journal of Inorganic

Biochemistry, 105 (2011) 1115-1122.

286

[187] H. He, H. Xiao, H. Kuang, Z. Xie, X. Chen, X. Jing, Y. Huang, Synthesis of

mesoporous silica nanoparticle–oxaliplatin conjugates for improved anticancer

drug delivery, Colloids and Surfaces B: Biointerfaces, 117 (2014) 75-81.

[188] J. Gu, J. Liu, Y. Li, W. Zhao, J. Shi, One-Pot Synthesis of Mesoporous Silica

Nanocarriers with Tunable Particle Sizes and Pendent Carboxylic Groups for

Cisplatin Delivery, Langmuir, 29 (2012) 403-410.

[189] S.D. Brown, P. Nativo, J.-A. Smith, D. Stirling, P.R. Edwards, B. Venugopal, D.J.

Flint, J.A. Plumb, D. Graham, N.J. Wheate, Gold Nanoparticles for the Improved

Anticancer Drug Delivery of the Active Component of Oxaliplatin, Journal of the

American Chemical Society, 132 (2010) 4678-4684.

[190] B. Duncan, C. Kim, V.M. Rotello, Gold nanoparticle platforms as drug and

biomacromolecule delivery systems, Journal of Controlled Release, 148 (2010)

122-127.

[191] R. Plummer, R.H. Wilson, H. Calvert, A.V. Boddy, M. Griffin, J. Sludden, M.J.

Tilby, M. Eatock, D.G. Pearson, C.J. Ottley, Y. Matsumura, K. Kataoka, T. Nishiya,

A Phase I clinical study of cisplatin-incorporated polymeric micelles (NC-6004) in

patients with solid tumours, British Journal of Cancer, 104 (2011) 593-598.

[192] M. Baba, Y. Matsumoto, A. Kashio, H. Cabral, N. Nishiyama, K. Kataoka, T.

Yamasoba, Micellization of cisplatin (NC-6004) reduces its ototoxicity in guinea

pigs, Journal of Controlled Release, 157 (2012) 112-117.

[193] Y. Matsumura, Poly (amino acid) micelle nanocarriers in preclinical and clinical

studies, Advanced Drug Delivery Reviews, 60 (2008) 899-914.

287

[194] T. Ueno, K. Endo, K. Hori, N. Ozaki, A. Tsuji, S. Kondo, N. Wakisaka, S. Murono,

K. Kataoka, Y. Kato, T. Yoshizaki, Assessment of antitumor activity and acute

peripheral neuropathy of 1,2-diaminocyclohexane platinum (II)-incorporating

micelles (NC-4016), International Journal of Nanomedicine, 9 (2014) 3005-3012.

[195] H. Cabral, N. Nishiyama, K. Kataoka, Optimization of (1,2-diamino-

cyclohexane)platinum(II)-loaded polymeric micelles directed to improved tumor

targeting and enhanced antitumor activity, Journal of Controlled Release, 121

(2007) 146-155.

[196] H. Cabral, N. Nishiyama, S. Okazaki, H. Koyama, K. Kataoka, Preparation and

biological properties of dichloro(1,2-diaminocyclohexane)platinum(II) (DACHPt)-

loaded polymeric micelles, Journal of Controlled Release, 101 (2005) 223-232.

[197] R. Song, Y. Joo Jun, J. Ik Kim, C. Jin, Y.S. Sohn, Synthesis, characterization, and

tumor selectivity of a polyphosphazene–platinum(II) conjugate, Journal of

Controlled Release, 105 (2005) 142-150.

[198] P.G. Avaji, H.I. Joo, J.H. Park, K.S. Park, Y.J. Jun, H.J. Lee, Y.S. Sohn, Synthesis

and properties of a new micellar polyphosphazene–platinum(II) conjugate drug,

Journal of Inorganic Biochemistry, 140 (2014) 45-52.

[199] J. Xu, Q. Fu, J.M. Ren, G. Bryant, G.G. Qiao, Novel drug carriers: from grafted

polymers to cross-linked vesicles, Chemical Communications, 49 (2013) 33-35.

[200] K. Osada, H. Cabral, Y. Mochida, S. Lee, K. Nagata, T. Matsuura, M. Yamamoto,

Y. Anraku, A. Kishimura, N. Nishiyama, K. Kataoka, Bioactive Polymeric

Metallosomes Self-Assembled through Block Copolymer–Metal Complexation,

Journal of the American Chemical Society, 134 (2012) 13172-13175.

288

[201] S. Valiahdi, A. Egger, W. Miklos, U. Jungwirth, K. Meelich, P. Nock, W. Berger,

C. Hartinger, M. Galanski, M. Jakupec, B. Keppler, Influence of extracellular pH

on the cytotoxicity, cellular accumulation, and DNA interaction of novel pH-

sensitive 2-aminoalcoholatoplatinum(II) complexes, Journal of Biological

Inorganic Chemistry, 18 (2013) 249-260.

[202] R.M. Sawant, J.P. Hurley, S. Salmaso, A. Kale, E. Tolcheva, T.S. Levchenko, V.P.

Torchilin, “SMART” Drug Delivery Systems:  Double-Targeted pH-Responsive

Pharmaceutical Nanocarriers, Bioconjugate Chemistry, 17 (2006) 943-949.

[203] W. Friebolin, G. Schilling, M. Zöller, E. Amtmann, Synthesis and

Structure−Activity Relationship of Novel Antitumoral Platinum Xanthate

Complexes, Journal of Medicinal Chemistry, 47 (2004) 2256-2263.

[204] Y.Y. Scaffidi-Domianello, A.A. Legin, M.A. Jakupec, V.B. Arion, V.Y. Kukushkin,

M. Galanski, B.K. Keppler, Synthesis, Characterization, and Cytotoxic Activity of

Novel Potentially pH-Sensitive Nonclassical Platinum(II) Complexes Featuring

1,3-Dihydroxyacetone Oxime Ligands, Inorganic Chemistry, 50 (2011) 10673-

10681.

[205] C.-H. Lin, S.-H. Cheng, W.-N. Liao, P.-R. Wei, P.-J. Sung, C.-F. Weng, C.-H. Lee,

Mesoporous silica nanoparticles for the improved anticancer efficacy of cis-platin,

International Journal of Pharmaceutics, 429 (2012) 138-147.

[206] D.P. Nowotnik, E. Cvitkovic, ProLindac™ (AP5346): A review of the

development of an HPMA DACH platinum Polymer Therapeutic, Advanced Drug

Delivery Reviews, 61 (2009) 1214-1219.

289

[207] A.K. Varkouhi, M. Scholte, G. Storm, H.J. Haisma, Endosomal escape pathways

for delivery of biologicals, Journal of Controlled Release, 151 (2011) 220-228.

[208] A.D. Carvalho Júnior, F.P. Vieira, V.J. De Melo, M.T.P. Lopes, J.N. Silveira, G.A.

Ramaldes, A. Garnier-Suillerot, E.C. Pereira-Maia, M.C. De Oliveira, Preparation

and cytotoxicity of cisplatin-containing liposomes, Brazilian Journal of Medical

and Biological Research, 40 (2007) 1149-1157.

[209] M. Murakami, H. Cabral, Y. Matsumoto, S. Wu, M.R. Kano, T. Yamori, N.

Nishiyama, K. Kataoka, Improving Drug Potency and Efficacy by Nanocarrier-

Mediated Subcellular Targeting, Science Translational Medicine, 3 (2011).

[210] M. Li, Z. Tang, S. Lv, W. Song, H. Hong, X. Jing, Y. Zhang, X. Chen, Cisplatin

crosslinked pH-sensitive nanoparticles for efficient delivery of doxorubicin,

Biomaterials, 35 (2014) 3851-3864.

[211] K. Szaciłowski, W. Macyk, A. Drzewiecka-Matuszek, M. Brindell, G. Stochel,

Bioinorganic Photochemistry:  Frontiers and Mechanisms, Chemical Reviews, 105

(2005) 2647-2694.

[212] P.C. Bruijnincx, P.J. Sadler, Controlling Platinum, Ruthenium and Osmium

Reactivity for Anticancer Drug Design, Advances in Inorganic Chemistry, 61

(2009) 1-62.

[213] N.P.E. Barry, P.J. Sadler, Exploration of the medical periodic table: towards new

targets, Chemical Communications, 49 (2013) 5106-5131.

[214] P.J. Bednarski, F.S. Mackay, P.J. Sadler, Photoactivatable platinum complexes,

Anti-Cancer Agents in Medicinal Chemistry, 7 (2007) 75-93.

290

[215] Y. Zhao, G.M. Roberts, S.E. Greenough, N.J. Farrer, M.J. Paterson, W.H. Powell,

V.G. Stavros, P.J. Sadler, Two-Photon-Activated Ligand Exchange in Platinum(II)

Complexes, Angewandte Chemie-International Edition, 51 (2012) 11263-11266.

[216] Y. Zhao, G.M. Roberts, S.E. Greenough, N.J. Farrer, M.J. Paterson, W.H. Powell,

V.G. Stavros, P.J. Sadler, Two-Photon-Activated Ligand Exchange in Platinum(II)

Complexes, Angewandte Chemie International Edition, 51 (2012) 11263-11266.

[217] N.A. Kratochwil, M. Zabel, K.-J. Range, P.J. Bednarski, Synthesis and X-ray

Crystal Structure of trans,cis-[Pt(OAc)2I2(en)]:  A Novel Type of Cisplatin Analog

That Can Be Photolyzed by Visible Light to DNA-Binding and Cytotoxic Species

in Vitro, Journal of Medicinal Chemistry, 39 (1996) 2499-2507.

[218] H. Choy, C. Park, M. Yao, Current Status and Future Prospects for Satraplatin, an

Oral Platinum Analogue, Clinical Cancer Research, 14 (2008) 1633-1638.

[219] N.A. Kratochwil, J.A. Parkinson, P.J. Bednarski, P.J. Sadler, Nucleotide Platination

Induced by Visible Light, Angewandte Chemie International Edition, 38 (1999)

1460-1463.

[220] A. Vogler, A. Kern, J. Hüttermann, Photochemical Reductive trans-Elimination

from trans-Diazidotetracyanoplatinate(IV), Angewandte Chemie International

Edition, 17 (1978) 524-525.

[221] P.J. Bednarski, R. Grünert, M. Zielzki, A. Wellner, F.S. Mackay, P.J. Sadler, Light-

Activated Destruction of Cancer Cell Nuclei by Platinum Diazide Complexes,

Chemistry & Biology, 13 61-67.

[222] A.F. Westendorf, J.A. Woods, K. Korpis, N.J. Farrer, L. Salassa, K. Robinson, V.

Appleyard, K. Murray, R. Grünert, A.M. Thompson, P.J. Sadler, P.J. Bednarski,

291

Trans,trans,trans-[PtIV(N3)2(OH)2(py)(NH3)]: A Light-Activated Antitumor

Platinum Complex That Kills Human Cancer Cells by an Apoptosis-Independent

Mechanism, Molecular Cancer Therapeutics, 11 (2012) 1894-1904.

[223] Y. Zhao, N.J. Farrer, H. Li, J.S. Butler, R.J. McQuitty, A. Habtemariam, F. Wang,

P.J. Sadler, De Novo Generation of Singlet Oxygen and Ammine Ligands by

Photoactivation of a Platinum Anticancer Complex, Angewandte Chemie-

International Edition, 52 (2013) 13633-13637.

[224] L. Ronconi, P.J. Sadler, Photoreaction pathways for the anticancer complex

trans,trans,trans-[Pt(N3)2(OH)2(NH3)2], Dalton Transactions, 40 (2011) 262-268.

[225] F.S. Mackay, J.A. Woods, P. Heringová, J. Kašpárková, A.M. Pizarro, S.A.

Moggach, S. Parsons, V. Brabec, P.J. Sadler, A potent cytotoxic photoactivated

platinum complex, Proceedings of the National Academy of Sciences of the U.S.A.,

104 (2007) 20743-20748.

[226] A.F. Westendorf, L. Zerzankova, L. Salassa, P.J. Sadler, V. Brabec, P.J. Bednarski,

Influence of pyridine versus piperidine ligands on the chemical, DNA binding and

cytotoxic properties of light activated trans,trans,trans- Pt(N3)(2)(OH)2(NH3)(L),

Journal of Inorganic Biochemistry, 105 (2011) 652-662.

[227] H.-C. Tai, Y. Zhao, N.J. Farrer, A.E. Anastasi, G. Clarkson, P.J. Sadler, R.J. Deeth,

A Computational Approach to Tuning the Photochemistry of Platinum(IV)

Anticancer Agents, Chemistry – A European Journal, 18 (2012) 10630-10642.

[228] J. Pracharova, L. Zerzankova, J. Stepankova, O. Novakova, N.J. Farrer, P.J. Sadler,

V. Brabec, J. Kasparkova, Interactions of DNA with a New Platinum(IV) Azide

292

Dipyridine Complex Activated by UVA and Visible Light: Relationship to Toxicity

in Tumor Cells, Chemical Research in Toxicology, 25 (2012) 1099-1111.

[229] H.-C. Tai, R. Brodbeck, J. Kasparkova, N.J. Farrer, V. Brabec, P.J. Sadler, R.J.

Deeth, Combined Theoretical and Computational Study of Interstrand DNA

Guanine-Guanine Cross-Linking by trans- Pt(pyridine)2 Derived from the

Photoactivated Prodrug trans,trans,trans- Pt(N3)2(OH)2(pyridine)2, Inorganic

Chemistry, 51 (2012) 6830-6841.

[230] A.M. Pizarro, R.J. McQuitty, F.S. Mackay, Y. Zhao, J.A. Woods, P.J. Sadler,

Cellular Accumulation, Lipophilicity and Photocytotoxicity of Diazido

Platinum(IV) Anticancer Complexes, ChemMedChem, 9 (2014) 1169-1175.

[231] Z. Zhu, X. Wang, T. Li, S. Aime, P.J. Sadler, Z. Guo, Platinum(II)-Gadolinium(III)

Complexes as Potential Single-Molecular Theranostic Agents for Cancer Treatment,

Angewandte Chemie-International Edition, 53 (2014) 13225-13228.

[232] P. Müller, B. Schröder, J.A. Parkinson, N.A. Kratochwil, R.A. Coxall, A. Parkin, S.

Parsons, P.J. Sadler, Nucleotide Cross-Linking Induced by Photoreactions of

Platinum(IV)–Azide Complexes, Angewandte Chemie International Edition, 42

(2003) 335-339.

[233] F.S. Mackay, J.A. Woods, H. Moseley, J. Ferguson, A. Dawson, S. Parsons, P.J.

Sadler, A Photoactivated trans-Diammine Platinum Complex as Cytotoxic as

Cisplatin, Chemistry – A European Journal, 12 (2006) 3155-3161.

[234] N.J. Farrer, J.A. Woods, V.P. Munk, F.S. Mackay, P.J. Sadler, Photocytotoxic

trans-Diam(m)ine Platinum(IV) Diazido Complexes More Potent than Their cis

Isomers, Chemical Research in Toxicology, 23 (2009) 413-421.

293

[235] Y. Zhao, J.A. Woods, N.J. Farrer, K.S. Robinson, J. Pracharova, J. Kasparkova, O.

Novakova, H. Li, L. Salassa, A.M. Pizarro, G.J. Clarkson, L. Song, V. Brabec, P.J.

Sadler, Diazido Mixed-Amine Platinum(IV) Anticancer Complexes Activatable by

Visible-Light Form Novel DNA Adducts, Chemistry – A European Journal, 19

(2013) 9578-9591.

[236] N.J. Farrer, J.A. Woods, L. Salassa, Y. Zhao, K.S. Robinson, G. Clarkson, F.S.

Mackay, P.J. Sadler, A Potent Trans-Diimine Platinum Anticancer Complex

Photoactivated by Visible Light, Angewandte Chemie International Edition, 49

(2010) 8905-8908.

[237] Y. Zhao, N.J. Farrer, H. Li, J.S. Butler, R.J. McQuitty, A. Habtemariam, F. Wang,

P.J. Sadler, De Novo Generation of Singlet Oxygen and Ammine Ligands by

Photoactivation of a Platinum Anticancer Complex, Angewandte Chemie

International Edition, 52 (2013) 13633-13637.

[238] N. Graf, S.J. Lippard, Redox activation of metal-based prodrugs as a strategy for

drug delivery, Advanced Drug Delivery Reviews, 64 (2012) 993-1004.

[239] P. Sova, A. Mistr, A. Kroutil, F. Zak, P. Pouckova, M. Zadinova, Preclinical anti-

tumor activity of a new oral platinum(IV) drug LA-12, Anti-Cancer Drugs, 16

(2005) 653-657.

[240] K.S. Lovejoy, S.J. Lippard, Non-traditional platinum compounds for improved

accumulation, oral bioavailability, and tumor targeting, Dalton Transactions, (2009)

10651-10659.

294

[241] A.R. Khokhar, Y. Deng, Y. Kido, Z.H. Siddik, Preparation, characterization, and

antitumor activity of new ethylenediamine platinum(IV) complexes containing

mixed carboxylate ligands, Journal of Inorganic Biochemistry, 50 (1993) 79-87.

[242] M.D. Hall, S. Amjadi, M. Zhang, P.J. Beale, T.W. Hambley, The mechanism of

action of platinum(IV) complexes in ovarian cancer cell lines, Journal of Inorganic

Biochemistry, 98 (2004) 1614-1624.

[243] Y.-R. Zheng, K. Suntharalingam, T.C. Johnstone, H. Yoo, W. Lin, J.G. Brooks, S.J.

Lippard, Pt(IV) Prodrugs Designed to Bind Non-Covalently to Human Serum

Albumin for Drug Delivery, Journal of the American Chemical Society, 136 (2014)

8790-8798.

[244] S. Choi, C. Filotto, M. Bisanzo, S. Delaney, D. Lagasee, J.L. Whitworth, A. Jusko,

C. Li, N.A. Wood, J. Willingham, A. Schwenker, K. Spaulding, Reduction and

Anticancer Activity of Platinum(IV) Complexes, Inorganic Chemistry, 37 (1998)

2500-2504.

[245] L. Ellis, H. Er, T. Hambley, The Influence of the Axial Ligands of a Series of

Platinum(IV) Anti-Cancer Complexes on Their Reduction to Platinum(II) and

Reaction With DNA, Australian Journal of Chemistry, 48 (1995) 793-806.

[246] C.K.J. Chen, J.Z. Zhang, J.B. Aitken, T.W. Hambley, Influence of Equatorial and

Axial Carboxylato Ligands on the Kinetic Inertness of Platinum(IV) Complexes in

the Presence of Ascorbate and Cysteine and within DLD-1 Cancer Cells, Journal of

Medicinal Chemistry, 56 (2013) 8757-8764.

[247] J.Z. Zhang, E. Wexselblatt, T.W. Hambley, D. Gibson, Pt(IV) analogs of

oxaliplatin that do not follow the expected correlation between electrochemical

295

reduction potential and rate of reduction by ascorbate, Chemical Communications,

48 (2012) 847-849.

[248] S. Dhar, S.J. Lippard, Mitaplatin, a potent fusion of cisplatin and the orphan drug

dichloroacetate, Proceedings of the National Academy of Sciences of the U.S.A.,

106 (2009) 22199-22204.

[249] E. Wexselblatt, E. Yavin, D. Gibson, Platinum(IV) Prodrugs with Haloacetato

Ligands in the Axial Positions can Undergo Hydrolysis under Biologically

Relevant Conditions, Angewandte Chemie International Edition, 52 (2013) 6059-

6062.

[250] J. Yang, W. Liu, M. Sui, J. Tang, Y. Shen, Platinum (IV)-coordinate polymers as

intracellular reduction-responsive backbone-type conjugates for cancer drug

delivery, Biomaterials, 32 (2011) 9136-9143.

[251] X. Wang, C. Yang, Y. Zhang, X. Zhen, W. Wu, X. Jiang, Delivery of platinum(IV)

drug to subcutaneous tumor and lung metastasis using bradykinin-potentiating

peptide-decorated chitosan nanoparticles, Biomaterials, 35 (2014) 6439-6453.

[252] S. Dhar, W.L. Daniel, D.A. Giljohann, C.A. Mirkin, S.J. Lippard, Polyvalent

Oligonucleotide Gold Nanoparticle Conjugates as Delivery Vehicles for

Platinum(IV) Warheads, Journal of the American Chemical Society, 131 (2009)

14652-14653.

[253] A. Kumar, S. Huo, X. Zhang, J. Liu, A. Tan, S. Li, S. Jin, X. Xue, Y. Zhao, T. Ji, L.

Han, H. Liu, X. Zhang, J. Zhang, G. Zou, T. Wang, S. Tang, X.-J. Liang,

Neuropilin-1-Targeted Gold Nanoparticles Enhance Therapeutic Efficacy of

Platinum(IV) Drug for Prostate Cancer Treatment, ACS Nano, 8 (2014) 4205-4220.

296

[254] S.L. Yoong, B.S. Wong, Q.L. Zhou, C.F. Chin, J. Li, T. Venkatesan, H.K. Ho, V.

Yu, W.H. Ang, G. Pastorin, Enhanced cytotoxicity to cancer cells by mitochondria-

targeting MWCNTs containing platinum(IV) prodrug of cisplatin, Biomaterials, 35

(2014) 748-759.

[255] Y. Shi, S.-A. Liu, D.J. Kerwood, J. Goodisman, J.C. Dabrowiak, Pt(IV) complexes

as prodrugs for cisplatin, Journal of Inorganic Biochemistry, 107 (2012) 6-14.

[256] W. Neumann, B.C. Crews, L.J. Marnett, E. Hey-Hawkins, Conjugates of Cisplatin

and Cyclooxygenase Inhibitors as Potent Antitumor Agents Overcoming Cisplatin

Resistance, ChemMedChem, 9 (2014) 1150-1153.

[257] V. Novohradsky, L. Zerzankova, J. Stepankova, O. Vrana, R. Raveendran, D.

Gibson, J. Kasparkova, V. Brabec, Antitumor platinum(IV) derivatives of

oxaliplatin with axial valproato ligands, Journal of Inorganic Biochemistry, 140

(2014) 72-79.

[258] K.R. Barnes, A. Kutikov, S.J. Lippard, Synthesis, Characterization, and

Cytotoxicity of a Series of Estrogen-Tethered Platinum(IV) Complexes, Chemistry

& Biology, 11 (2004) 557-564.

[259] S. Abramkin, S.M. Valiahdi, M.A. Jakupec, M. Galanski, N. Metzler-Nolte, B.K.

Keppler, Solid-phase synthesis of oxaliplatin-TAT peptide bioconjugates, Dalton

Transactions, 41 (2012) 3001-3005.

[260] D.Y.Q. Wong, C.H.F. Yeo, W.H. Ang, Immuno-Chemotherapeutic Platinum(IV)

Prodrugs of Cisplatin as Multimodal Anticancer Agents, Angewandte Chemie

International Edition, 53 (2014) 6752-6756.

297

[261] N. Graf, T.E. Mokhtari, I.A. Papayannopoulos, S.J. Lippard, Platinum(IV)-

chlorotoxin (CTX) conjugates for targeting cancer cells, Journal of Inorganic

Biochemistry, 110 (2012) 58-63.

[262] Simon P. Wisnovsky, Justin J. Wilson, Robert J. Radford, Mark P. Pereira, Maria R.

Chan, Rebecca R. Laposa, Stephen J. Lippard, Shana O. Kelley, Targeting

Mitochondrial DNA with a Platinum-Based Anticancer Agent, Chemistry &

Biology, 20 (2013) 1323-1328.

[263] K. Suntharalingam, Y. Song, S.J. Lippard, Conjugation of vitamin E analog [small

alpha]-TOS to Pt(IV) complexes for dual-targeting anticancer therapy, Chemical

Communications, 50 (2014) 2465-2468.

[264] S. Dhar, N. Kolishetti, S.J. Lippard, O.C. Farokhzad, Targeted delivery of a

cisplatin prodrug for safer and more effective prostate cancer therapy in vivo,

Proceedings of the National Academy of Sciences of the U.S.A., 108 (2011) 1850-

1855.

[265] S. Dhar, F.X. Gu, R. Langer, O.C. Farokhzad, S.J. Lippard, Targeted delivery of

cisplatin to prostate cancer cells by aptamer functionalized Pt(IV) prodrug-PLGA–

PEG nanoparticles, Proceedings of the National Academy of Sciences of the U.S.A.,

105 (2008) 17356-17361.

[266] D. Montagner, S.Q. Yap, W.H. Ang, A Fluorescent Probe for Investigating the

Activation of Anticancer Platinum(IV) Prodrugs Based on the Cisplatin Scaffold,

Angewandte Chemie International Edition, 52 (2013) 11785-11789.

298

[267] W. Scarano, H.T.T. Duong, H. Lu, P.L. De Souza, M.H. Stenzel, Folate

Conjugation to Polymeric Micelles via Boronic Acid Ester to Deliver Platinum

Drugs to Ovarian Cancer Cell Lines, Biomacromolecules, 14 (2013) 962-975.

[268] H. Xiao, H. Song, Q. Yang, H. Cai, R. Qi, L. Yan, S. Liu, Y. Zheng, Y. Huang, T.

Liu, X. Jing, A prodrug strategy to deliver cisplatin(IV) and paclitaxel in

nanomicelles to improve efficacy and tolerance, Biomaterials, 33 (2012) 6507-6519.

[269] J. Graham, M. Mushin, P. Kirkpatrick, Oxaliplatin, Nature Reviews Drug

Discovery, 3 (2004) 11-12.

[270] G. Momekov, A. Bakalova, M. Karaivanova, Novel approaches towards

development of non-classical platinum-based antineoplastic agents: Design of

platinum complexes characterized by an alternative DNA-binding pattern and/or

tumor-targeted cytotoxicity, Current Medicinal Chemistry, 12 (2005) 2177-2191.

[271] R. Guddneppanavar, U. Bierbach, Adenine-N3 in the DNA minor groove - an

emerging target for platinum containing anticancer pharmacophores, Anticancer

Agents Med Chem, 7 (2007) 125-138.

[272] S. Komeda, A. Casini, Next-generation anticancer metallodrugs, Current Topics in

Medicinal Chemistry, 12 (2012) 219-235.

[273] S. Komeda, Unique platinum-DNA interactions may lead to more effective

platinum-based antitumor drugs, Metallomics, 3 (2011) 650-655.

[274] H. Baruah, M.W. Wright, U. Bierbach, Solution structural study of a DNA duplex

containing the guanine-N7 adduct formed by a cytotoxic platinum-acridine hybrid

agent, Biochemistry, 44 (2005) 6059-6070.

299

[275] X. Qiao, A.E. Zeitany, M.W. Wright, A.S. Essader, K.E. Levine, G.L. Kucera, U.

Bierbach, Analysis of the DNA damage produced by a platinum-acridine antitumor

agent and its effects in NCI-H460 lung cancer cells, Metallomics, 4 (2012) 645-652.

[276] C.L. Smyre, G. Saluta, T.E. Kute, G.L. Kucera, U. Bierbach, Inhibition of DNA

Synthesis by a Platinum-Acridine Hybrid Agent Leads to Potent Cell Kill in Non-

Small Cell Lung Cancer, ACS Med Chem Lett, 2 (2011) 870-874.

[277] L.A. Graham, G.M. Wilson, T.K. West, C.S. Day, G.L. Kucera, U. Bierbach,

Unusual Reactivity of a Potent Platinum-Acridine Hybrid Antitumor Agent, ACS

Medicinal Chemistry Letters, 2 (2011) 687-691.

[278] J.R. Choudhury, L. Rao, U. Bierbach, Rates of intercalator-driven platination of

DNA determined by a restriction enzyme cleavage inhibition assay, J Biol Inorg

Chem, 16 (2011) 373-380.

[279] Z. Ma, J.R. Choudhury, M.W. Wright, C.S. Day, G. Saluta, G.L. Kucera, U.

Bierbach, A non-cross-linking platinum-acridine agent with potent activity in non-

small-cell lung cancer, Journal of Medicinal Chemistry, 51 (2008) 7574-7580.

[280] S.C. Alley, N.M. Okeley, P.D. Senter, Antibody-drug conjugates: targeted drug

delivery for cancer, Current Opinion in Chemical Biology, 14 (2010) 529-537.

[281] R. Srinivasan, J. Li, S.L. Ng, K.A. Kalesh, S.Q. Yao, Methods of using click

chemistry in the discovery of enzyme inhibitors, Nature Protocols, 2 (2007) 2655-

2664.

[282] M. Bahta, F. Liu, S.E. Kim, A.G. Stephen, R.J. Fisher, T.R. Burke, Jr., Oxime-

based linker libraries as a general approach for the rapid generation and screening

of multidentate inhibitors, Nature Protocols, 7 (2012) 686-702.

300

[283] V.Y. Kukushkin, A.J. Pombeiro, Additions to metal-activated organonitriles,

Chemical Reviews, 102 (2002) 1771-1802.

[284] J.M. Baskin, J.A. Prescher, S.T. Laughlin, N.J. Agard, P.V. Chang, I.A. Miller, A.

Lo, J.A. Codelli, C.R. Bertozzi, Copper-free click chemistry for dynamic in vivo

imaging, Proceedings of the National Academy of Sciences of the U.S.A., 104

(2007) 16793-16797.

[285] R.E. Wang, F. Costanza, Y. Niu, H. Wu, Y. Hu, W. Hang, Y. Sun, J. Cai,

Development of self-immolative dendrimers for drug delivery and sensing, Journal

of Controlled Release, 159 (2012) 154-163.

[286] M.R. Karver, R. Weissleder, S.A. Hilderbrand, Bioorthogonal reaction pairs enable

simultaneous, selective, multi-target imaging, Angew Chem Int Ed Engl, 51 (2012)

920-922.

[287] S.C. Dhara, A rapid method for the synthesis of cis-[Pt(NH3)2Cl2], Indian Journal

of Chemistry, 8 (1970) 193-194.

[288] T.M. Augustus, J. Anderson, S.M. Hess, U. Bierbach,

Bis(acridinylthiourea)platinum(II) Complexes: Synthesis, DNA affinity, and

biological activity in ghoblastoma cells, Bioorganic & Medicinal Chemistry Letters,

13 (2003) 855-858.

[289] Z. Ma, J.R. Choudhury, M.W. Wright, C.S. Day, G. Saluta, G.L. Kucera, U.

Bierbach, A Non-Cross-Linking Platinum−Acridine Agent with Potent Activity in

Non-Small-Cell Lung Cancer, Journal of Medicinal Chemistry, 51 (2008) 7574-

7580.

301

[290] D. Wang, S.J. Lippard, Cellular processing of platinum anticancer drugs, Nature

Reviews Drug Discovery, 4 (2005) 307-320.

[291] A.V. Klein, T.W. Hambley, Platinum Drug Distribution in Cancer Cells and

Tumors, Chemical Reviews, 109 (2009) 4911-4920.

[292] L.E. Wedlock, S.J. Berners-Price, Recent Advances in Mapping the Sub-cellular

Distribution of Metal-Based Anticancer Drugs, Australian Journal of Chemistry, 64

(2011) 692-704.

[293] S. Trapp, R.W. Horobin, A predictive model for the selective accumulation of

chemicals in tumor cells, European Biophysics Journal with Biophysics Letters, 34

(2005) 959-966.

[294] X.J. Liang, D.W. Shen, K.G. Chen, S.M. Wincovitch, S.H. Garfield, M.M.

Gottesman, Trafficking and localization of platinum complexes in cisplatin-

resistant cell lines monitored by fluorescence-labeled platinum, Journal of Cellular

Physiology, 202 (2005) 635-641.

[295] J. Suryadi, U. Bierbach, DNA Metalating-Intercalating Hybrid Agents for the

Treatment of Chemoresistant Cancers, Chemistry – A European Journal, 18 (2012)

12926-12934.

[296] K. Cheung-Ong, K.T. Song, Z.D. Ma, D. Shabtai, A.Y. Lee, D. Gallo, L.E. Heisler,

G.W. Brown, U. Bierbach, G. Giaever, C. Nislow, Comparative Chemogenomics

To Examine the Mechanism of Action of DNA-Targeted Platinum-Acridine

Anticancer Agents, ACS Chemical Biology, 7 (2012) 1892-1901.

302

[297] D. Fornasiero, T. Kurucsev, The binding of 9-aminoacridine to calf thymus DNA

in aqueous solution-electronic spectral studies, Biophysical Chemistry, 23 (1985)

31-37.

[298] Z.P. Demko, K.B. Sharpless, A click chemistry approach to tetrazoles by Huisgen

1,3-dipolar cycloaddition: Synthesis of 5-sulfonyl tetrazoles from azides and

sulfonyl cyanides, Angewandte Chemie-International Edition, 41 (2002) 2110-2113.

[299] S.B. Buck, J. Bradford, K.R. Gee, B.J. Agnew, S.T. Clarke, A. Salic, Detection of

S-phase cell cycle progression using 5-ethynyl-2 '-deoxyuridine incorporation with

click chemistry an alternative to using 5-bromo-2 '-deoxyuridine antibodies,

Biotechniques, 44 (2008) 927-929.

[300] S.A. Kane, S.J. Lippard, Photoreactivity of platinum(II) in cisplatin-modified DNA

affords specific cross-links to HMG domain proteins, Biochemistry, 35 (1996)

2180-2188.

[301] H. Kostrhunova, J. Malina, A.J. Pickard, J. Stepankova, M. Vojtiskova, J.

Kasparkova, T. Muchova, M.L. Rohlfing, U. Bierbach, V. Brabec, Replacement of

a Thiourea with an Amidine Group in a Monofunctional Platinum-Acridine

Antitumor Agent. Effect on DNA Interactions, DNA Adduct Recognition and

Repair, Molecular Pharmaceutics, 8 (2011) 1941-1954.

[302] A. Saraste, K. Pulkki, Morphologic and biochemical hallmarks of apoptosis,

Cardiovascular Research, 45 (2000) 528-537.

[303] Y.W. Lam, L. Trinkle-Mulcahy, A.I. Lamond, The nucleolus, Journal of Cell

Science, 118 (2005) 1335-1337.

303

[304] X. Ning, J. Guo, M.A. Wolfert, G.J. Boons, Visualizing metabolically labeled

glycoconjugates of living cells by copper-free and fast huisgen cycloadditions,

Angew Chem Int Ed Engl, 47 (2008) 2253-2255.

[305] C.L. Smyre, G. Saluta, T.E. Kute, G.L. Kucera, U. Bierbach, Inhibition of DNA

Synthesis by a Platinum-Acridine Hybrid Agent Leads to Potent Cell Kill in

Nonsmall Cell Lung Cancer, ACS Medicinal Chemistry Letters, 2 (2011) 870-874.

[306] D. Drygin, W.G. Rice, I. Grummt, The RNA Polymerase I Transcription

Machinery: An Emerging Target for the Treatment of Cancer, Annual Review of

Pharmacology and Toxicology2010, pp. 131-156.

[307] D. Drygin, A. Lin, J. Bliesath, C.B. Ho, S.E. O'Brien, C. Proffitt, M. Omori, M.

Haddach, M.K. Schwaebe, A. Siddiqui-Jain, N. Streiner, J.E. Quin, E. Sanij, M.J.

Bywater, R.D. Hannan, D. Ryckman, K. Anderes, W.G. Rice, Targeting RNA

Polymerase I with an Oral Small Molecule CX-5461 Inhibits Ribosomal RNA

Synthesis and Solid Tumor Growth, Cancer Research, 71 (2011) 1418-1430.

[308] J.M. Baskin, J.A. Prescher, S.T. Laughlin, N.J. Agard, P.V. Chang, I.A. Miller, A.

Lo, J.A. Codelli, C.R. Bertozzi, Copper-free click chemistry for dynamic in vivo

imaging, Proceedings of the National Academy of Sciences of the United States of

America, 104 (2007) 16793-16797.

[309] L.A. Graham, J. Suryadi, T.K. West, G.L. Kucera, U. Bierbach, Synthesis, aqueous

reactivity, and biological evaluation of carboxylic Acid ester-functionalized

platinum-acridine hybrid anticancer agents, J Med Chem, 55 (2012) 7817-7827.

[310] S. Ding, X. Qiao, J. Suryadi, G.S. Marrs, G.L. Kucera, U. Bierbach, Using

Fluorescent Post-Labeling To Probe the Subcellular Localization of DNA-Targeted

304

Platinum Anticancer Agents, Angewandte Chemie-International Edition, 52 (2013)

3350-3354.

[311] A.C. Society, Cancer Facts & Figures 2012, American Cancer Society, Atlanta, GA,

2012.

[312] M.J. Higgins, V. Stearns, Understanding Resistance to Tamoxifen in Hormone

Receptor-Positive Breast Cancer, Clinical Chemistry, 55 (2009) 1453-1455.

[313] M.P. Decatris, S. Sundar, K.J. O'Byrne, Platinum-based chemotherapy in

metastatic breast cancer: current status, Cancer Treatment Reviews, 30 (2004) 53-

81.

[314] C.L. Smith, B.W. O'Malley, Coregulator function: A key to understanding tissue

specificity of selective receptor modulators, Endocrine Reviews, 25 (2004) 45-71.

[315] E. Kim, P.T. Rye, J.M. Essigmann, R.G. Croy, A bifunctional platinum(II)

antitumor agent that forms DNA adducts with affinity for the estrogen receptor,

Journal of Inorganic Biochemistry, 103 (2009) 256-261.

[316] R. Gust, W. Beck, G. Jaouen, H. Schoenenberger, Optimization of cisplatin for the

treatment of hormone dependent tumoral diseases Part 1: Use of steroidal ligands,

Coordination Chemistry Reviews, 253 (2009) 2742-2759.

[317] R. Gust, W. Beck, G. Jaouen, H. Schoenenberger, Optimization of cisplatin for the

treatment of hormone-dependent tumoral diseases Part 2: Use of non-steroidal

ligands, Coordination Chemistry Reviews, 253 (2009) 2760-2779.

[318] K. Mitra, J.C. Marquis, S.M. Hillier, P.T. Rye, B. Zayas, A.S. Lee, J.M. Essigmann,

R.G. Croy, A rationally designed genotoxin that selectively destroys estrogen

305

receptor-positive breast cancer cells, Journal of the American Chemical Society,

124 (2002) 1862-1863.

[319] P.J. Burke, T.H. Koch, Design, synthesis, and biological evaluation of doxorubicin

- Formaldehyde conjugates targeted to breast cancer cells, Journal of Medicinal

Chemistry, 47 (2004) 1193-1206.

[320] K.L. Dao, R.R. Sawant, J.A. Hendricks, V. Ronga, V.P. Torchilin, R.N. Hanson,

Design, Synthesis, and Initial Biological Evaluation of a Steroidal Anti-Estrogen-

Doxorubicin Bioconjugate for Targeting Estrogen Receptor-Positive Breast Cancer

Cells, Bioconjugate Chemistry, 23 (2012) 785-795.

[321] K.W. Peng, H. Wang, Z. Qin, G.T. Wijewickrama, M. Lu, Z. Wang, J.L. Bolton,

G.R.J. Thatcher, Selective Estrogen Receptor Modulator Delivery of Quinone

Warheads to DNA Triggering Apoptosis in Breast Cancer Celts, ACS Chemical

Biology, 4 (2009) 1039-1049.

[322] M.D. Johnson, H. Zuo, K.H. Lee, J.P. Trebley, J.M. Rae, R.V. Weatherman, Z.

Desta, D.A. Flockhart, T.C. Skaar, Pharmacological characterization of 4-hydroxy-

N-desmethyl tamoxifen, a novel active metabolite of tamoxifen, Breast Cancer

Research and Treatment, 85 (2004) 151-159.

[323] A.K. Shiau, D. Barstad, P.M. Loria, L. Cheng, P.J. Kushner, D.A. Agard, G.L.

Greene, The structural basis of estrogen receptor/coactivator recognition and the

antagonism of this interaction by tamoxifen, Cell, 95 (1998) 927-937.

[324] B.S. Katzenellenbogen, M.J. Norman, R.L. Eckert, S.W. Peltz, W.F. Mangel,

Bioactivities, estrogen-receptor interactions, and plasminogen activator-inducing

306

activities of tamoxifen an hydroxytamoxifen isomers in MCF-7 human-breast

cancer-cells, Cancer Research, 44 (1984) 112-119.

[325] S.C. Alley, N.M. Okeley, P.D. Senter, Antibody-drug conjugates: targeted drug

delivery for cancer, Current Opinion in Chemical Biology, 14 (2010) 529-537.

[326] M. Lacroix, G. Leclercq, Relevance of breast cancer cell lines as models for breast

tumours: an update, Breast Cancer Research and Treatment, 83 (2004) 249-289.

[327] M.J. Abedin, D. Wang, M.A. McDonnell, U. Lehmann, A. Kelekar, Autophagy

delays apoptotic death in breast cancer cells following DNA damage, Cell Death

and Differentiation, 14 (2007) 500-510.

[328] R.E. Wang, F. Costanza, Y. Niu, H. Wu, Y. Hu, W. Hang, Y. Sun, J. Cai,

Development of self-immolative dendrimers for drug delivery and sensing, Journal

of Controlled Release, 159 (2012) 154-163.

[329] E.A. Musgrove, R.L. Sutherland, Biological determinants of endocrine resistance in

breast cancer, Nature Reviews Cancer, 9 (2009) 631-643.

[330] D.D. Yu, B.M. Forman, Simple and Efficient Production of (Z)-4-

Hydroxytamoxifen, a Potent Estrogen Receptor Modulator, The Journal of Organic

Chemistry, 68 (2003) 9489-9491.

[331] A.H. Fauq, G.M. Maharvi, D. Sinha, A convenient synthesis of (Z)-4-hydroxy-N-

desmethyltamoxifen (endoxifen), Bioorganic & Medicinal Chemistry Letters, pp.

3036-3038.

[332] K.M. Huttunen, H. Raunio, J. Rautio, Prodrugs-from Serendipity to Rational

Design, Pharmacological Reviews, 63 (2011) 750-771.

307

[333] T. Lammers, F. Kiessling, W.E. Hennink, G. Storm, Drug targeting to tumors:

Principles, pitfalls and (pre-) clinical progress, Journal of Controlled Release, 161

(2012) 175-187.

[334] E.H. Kerns, L. Di, Pharmaceutical profiling in drug discovery, Drug Discovery

Today, 8 (2003) 316-323.

[335] K.M. H. Lullmann, A. Ziegler, D. Bieger, , Color Atlas of Pharmacology, 2ednd ed,

Thieme, Stuttgart (2000) pp.32-42.

[336] S.E. Pratt, S. Durland-Busbice, R.L. Shepard, K. Heinz-Taheny, P.W. Iversen, A.H.

Dantzig, Human Carboxylesterase-2 Hydrolyzes the Prodrug of Gemcitabine

(LY2334737) and Confers Prodrug Sensitivity to Cancer Cells, Clinical Cancer

Research, 19 (2013) 1159-1168.

[337] T. Imai, M. Taketani, M. Shii, M. Hosokawa, K. Chiba, Substrate specificity of

carboxylesterase isozymes and their contribution to hydrolase activity in human

liver and small intestine, Drug Metabolism and Disposition, 34 (2006) 1734-1741.

[338] S. Ding, X. Qiao, G.L. Kucera, U. Bierbach, Using a Build-and-Click Approach for

Producing Structural and Functional Diversity in DNA-Targeted Hybrid Anticancer

Agents, Journal of Medicinal Chemistry, 55 (2012) 10198-10203.

[339] G. Natile, F.P. Intini, R. Bertani, R.A. Michelin, M. Mozzon, S.M. Sbovata, A.

Venzo, R. Seraglia, Synthesis and characterisation of the amidine complexes trans-

PtCl(NH3) {HN=C(NH2)R}(2) Cl (R = Me, Ph, CH2Ph) derived from addition of

NH3 to the coordinated nitriles in trans- PtCl2(N CR)(2), Journal of

Organometallic Chemistry, 690 (2005) 2121-2127.

308

[340] R.A. Michelin, P. Sgarbossa, S.M. Sbovata, V. Gandin, C. Marzano, R. Bertani,

Chemistry and Biological Activity of Platinum Amidine Complexes,

ChemMedChem, 6 (2011) 1172-1183.

[341] J. Wang, E.T. Williams, J. Bourgea, Y.N. Wong, C.J. Patten, Characterization of

Recombinant Human Carboxylesterases: Fluorescein Diacetate as a Probe

Substrate for Human Carboxylesterase 2, Drug Metabolism and Disposition, 39

(2011) 1329-1333.

[342] S.M. Valiahdi, A.E. Egger, W. Miklos, U. Jungwirth, K. Meelich, P. Nock, W.

Berger, C.G. Hartinger, M. Galanski, M.A. Jakupec, B.K. Keppler, Influence of

extracellular pH on the cytotoxicity, cellular accumulation, and DNA interaction of

novel pH-sensitive 2-aminoalcoholatoplatinum(II) complexes, Journal of

Biological Inorganic Chemistry, 18 (2013) 249-260.

[343] D.A. Weaver, E.L. Crawford, K.A. Warner, F. Elkhairi, S.A. Khuder, J.C. Willey,

ABCC5, ERCC2, XPA and XRCC1 transcript abundance levels correlate with

cisplatin chemoresistance in non-small cell lung cancer cell lines, Molecular

Cancer, 4 (2005).

[344] B.L. Barthel, Z.Y. Zhang, D.L. Rudnicki, C.D. Coldren, M. Polinkovsky, H.R. Sun,

G.G. Koch, D.C.F. Chan, T.H. Koch, Preclinical Efficacy of a Carboxylesterase 2-

Activated Prodrug of Doxazolidine, Journal of Medicinal Chemistry, 52 (2009)

7678-7688.

[345] Z. Zhang, H. Hatta, K. Tanabe, S. Nishimoto, A new class of 5-fluoro-2 '-

deoxyuridine prodrugs conjugated with a tumor-homing cyclic peptide CNGRC by

309

ester linkers: Synthesis, reactivity, and tumor-cell-selective cytotoxicity,

Pharmaceutical Research, 22 (2005) 381-389.

[346] E.T. Martins, H. Baruah, J. Kramarczyk, G. Saluta, C.S. Day, G.L. Kucera, U.

Bierbach, Design, synthesis, and biological activity of a novel non-cisplatin-type

platinum-acridine pharmacophore, J Med Chem, 44 (2001) 4492-4496.

[347] J.B. Gibbs, Mechanism-based target identification and drug discovery in cancer

research, Science, 287 (2000) 1969-1973.

[348] J. Suryadi, U. Bierbach, DNA Metalating-Intercalating Hybrid Agents for the

Treatment of Chemoresistant Cancers, Chemistry-a European Journal, 18 (2012)

12926-12934.

[349] K. Cheung-Ong, K.T. Song, Z. Ma, D. Shabtai, A.Y. Lee, D. Gallo, L.E. Heisler,

G.W. Brown, U. Bierbach, G. Giaever, C. Nislow, Comparative Chemogenomics

To Examine the Mechanism of Action of DNA-Targeted Platinum-Acridine

Anticancer Agents, ACS Chemical Biology, 7 (2012) 1892-1901.

[350] S. Ding, A.J. Pickard, G.L. Kucera, U. Bierbach, Design of Enzymatically

Cleavable Prodrugs of a Potent Platinum-Containing Anticancer Agent, Chemistry

– A European Journal, 20 (2014) 16164-16173.

[351] S. Ding, X. Qiao, G.L. Kucera, U. Bierbach, Design of a platinum-acridine-

endoxifen conjugate targeted at hormone-dependent breast cancer, Chemical

Communications, 49 (2013) 2415-2417.

[352] R. Morphy, Z. Rankovic, Designed multiple ligands. An emerging drug discovery

paradigm, Journal of Medicinal Chemistry, 48 (2005) 6523-6543.

310

[353] J. Yang, X. Sun, W. Mao, M. Sui, J. Tang, Y. Shen, Conjugate of Pt(IV)-Histone

Deacetylase Inhibitor as a Prodrug for Cancer Chemotherapy, Molecular

Pharmaceutics, 9 (2012) 2793-2800.

[354] S. Dhar, S.J. Lippard, Mitaplatin, a potent fusion of cisplatin and the orphan drug

dichloroacetate, Proceedings of the National Academy of Sciences of the United

States of America, 106 (2009) 22199-22204.

[355] R.K. Pathak, S. Marrache, J.H. Choi, T.B. Berding, S. Dhar, The Prodrug Platin-A:

Simultaneous Release of Cisplatin and Aspirin, Angewandte Chemie-International

Edition, 53 (2014) 1963-1967.

[356] C.T. Keith, A.A. Borisy, B.R. Stockwell, Multicomponent therapeutics for

networked systems, Nature Reviews Drug Discovery, 4 (2005) 71-U10.

[357] S.A. Eisen, D.K. Miller, R.S. Woodward, E. Spitznagel, T.R. Przybeck, The effect

of prescribled daily dose frequency on patient medication compliance, Archives of

Internal Medicine, 150 (1990) 1881-1884.

[358] A.S. Reddy, S. Zhang, Polypharmacology: drug discovery for the future, Expert

Review of Clinical Pharmacology, 6 (2013) 41-47.

[359] R. Morphy, C. Kay, Z. Rankovic, From magic bullets to designed multiple ligands,

Drug Discovery Today, 9 (2004) 641-651.

[360] Y. Bansal, O. Silakari, Multifunctional compounds: Smart molecules for

multifactorial diseases, European Journal of Medicinal Chemistry, 76 (2014) 31-42.

[361] A. El-Faham, F. Albericio, Peptide Coupling Reagents, More than a Letter Soup,

Chemical Reviews, 111 (2011) 6557-6602.

311

[362] P. Thirumurugan, D. Matosiuk, K. Jozwiak, Click Chemistry for Drug

Development and Diverse Chemical-Biology Applications, Chemical Reviews, 113

(2013) 4905-4979.

[363] A.J. Pickard, F. Liu, T.F. Bartenstein, L.G. Haines, K.E. Levine, G.L. Kucera, U.

Bierbach, Redesigning the DNA-Targeted Chromophore in Platinum-Acridine

Anticancer Agents: A Structure-Activity Relationship Study, Chemistry – A

European Journal, 20 (2014) 16174-16187.

[364] G. Dirscherl, B. Koenig, The use of solid-phase synthesis techniques for the

preparation of peptide-metal complex conjugates, European Journal of Organic

Chemistry, (2008) 597-634.

[365] E. Valeur, M. Bradley, Amide bond formation: beyond the myth of coupling

reagents, Chemical Society Reviews, 38 (2009) 606-631.

[366] A. El-Faham, F. Albericio, COMU: A third generation of uronium-type coupling

reagents, Journal of Peptide Science, 16 (2010) 6-9.

[367] C. Najera, From alpha-amino acids to peptides: All you need for the journey,

Synlett, (2002) 1388-1403.

[368] J. Lu, C. Liu, P. Wang, M. Ghazwani, J. Xu, Y. Huang, X. Ma, P. Zhang, S. Li,

The self-assembling camptothecin-tocopherol prodrug: An effective approach for

formulating camptothecin, Biomaterials, 62 (2015) 176-187.

[369] S. Aryal, C.-M.J. Hu, L. Zhang, Synthesis of Ptsome: a platinum-based liposome-

like nanostructure, Chemical Communications, 48 (2012) 2630-2632.

312

[370] J.R. Vargas, E.G. Stanzl, N.N.H. Teng, P.A. Wender, Cell-Penetrating,

Guanidinium-Rich Molecular Transporters for Overcoming Efflux-Mediated

Multidrug Resistance, Molecular Pharmaceutics, 11 (2014) 2553-2565.

[371] T. Ekblad, E. Camaioni, H. Schuler, A. Macchiarulo, PARP inhibitors:

polypharmacology versus selective inhibition, FEBS Journal, 280 (2013) 3563-

3575.

[372] M. Yang, A.J. Pickard, X. Qiao, M.J. Gueble, C.S. Day, G.L. Kucera, U. Bierbach,

Synthesis, Reactivity, and Biological Activity of Gold(I) Complexes Modified with

Thiourea-Functionalized Tyrosine Kinase Inhibitors, Inorganic Chemistry, 54

(2015) 3316-3324.

[373] L. Milane, M. Trivedi, A. Singh, M. Talekar, M. Amiji, Mitochondrial biology,

targets, and drug delivery, Journal of Controlled Release, 207 (2015) 40-58.

[374] V.K. Yellepeddi, K.K. Vangara, S. Palakurthi, Poly(amido)amine (PAMAM)

dendrimer-cisplatin complexes for chemotherapy of cisplatin-resistant ovarian

cancer cells, Journal of Nanoparticle Research, 15 (2013).

[375] D.M. Turner, C.T.M.B. Tom, A.R. Renslo, Simple Plate-Based, Parallel Synthesis

of Disulfide Fragments using the CuAAC Click Reaction, Acs Combinatorial

Science, 16 (2014) 661-664.

[376] O. Moukha-Chafiq, R.C. Reynolds, Parallel Solution-Phase Synthesis and General

Biological Activity of a Uridine Antibiotic Analog Library, Acs Combinatorial

Science, 16 (2014) 232-237.

313

[377] S. Mukhopadhyay, C.M. Barnes, A. Haskel, S.M. Short, K.R. Barnes, S.J. Lippard,

Conjugated platinum(IV)-peptide complexes for targeting angiogenic tumor

vasculature, Bioconjugate Chemistry, 19 (2008) 39-49.

[378] W. Bannwarth, R. Knorr, Formation of carboxamides with N,N,N',N'-tetramethyl

(succinimido) uronium tetrafluoroborate in aqueous / organic solvent systems,

Tetrahedron Letters, 32 (1991) 1157-1160.

[379] M.J. Wigglesworth, D.C. Murray, C.J. Blackett, M. Kossenjans, J.W.M. Nissink,

Increasing the delivery of next generation therapeutics from high throughput

screening libraries, Current Opinion in Chemical Biology, 26 (2015) 104-110.

[380] S. van Zutphen, E.A. Stone, S. van Rijt, M.S. Robillard, G.A. van der Marel, H.S.

Overkleeft, H. den Dulk, J. Brouwer, J. Reedijk, Combinatorial discovery of new

asymmetric cis platinum anticancer complexes is made possible with solid-phase

synthetic methods, Journal of Inorganic Biochemistry, 99 (2005) 2032-2038.

[381] C.J. Ziegler, A.P. Silverman, S.J. Lippard, High-throughput synthesis and

screening of platinum drug candidates, Journal of Biological Inorganic Chemistry,

5 (2000) 774-783.

[382] A. Cern, A. Golbraikh, A. Sedykh, A. Tropsha, Y. Barenholz, A. Goldblum,

Quantitative structure - property relationship modeling of remote liposome loading

of drugs, Journal of Controlled Release, 160 (2012) 147-157.

[383] A.A. Kale, V.P. Torchilin, "Smart" drug carriers: PEGylated TATp-Modified pH-

Sensitive Liposomes, Journal of Liposome Research, 17 (2007) 197-203.

[384] C. Allen, N. Dos Santos, R. Gallagher, G.N.C. Chiu, Y. Shu, W.M. Li, S.A.

Johnstone, A.S. Janoff, L.D. Mayer, M.S. Webb, M.B. Bally, Controlling the

314

physical behavior and biological performance of liposome formulations through

use of surface grafted poly(ethylene glycol), Bioscience Reports, 22 (2002) 225-

250.

[385] V.P. Torchilin, V.G. Omelyanenko, M.I. Papisov, A.A. Bogdanov, V.S. Trubetskoy,

J.N. Herron, C.A. Gentry, Poly(ethylene glycol) on the liposome surface-on the

mechanism of polymer-coated liposome longevity, Biochimica Et Biophysica Acta-

Biomembranes, 1195 (1994) 11-20.

[386] M.-L. Briuglia, C. Rotella, A. McFarlane, D.A. Lamprou, Influence of cholesterol

on liposome stability and on in vitro drug release, Drug Delivery and Translational

Research, 5 (2015) 231-242.

[387] N. Dos Santos, L.D. Mayer, S.A. Abraham, R.C. Gallagher, R.A.K. Cox, P.G.

Tardi, M.B. Bally, Improved retention of idarubicin after intravenous injection

obtained for cholesterol-free liposomes, Biochimica Et Biophysica Acta-

Biomembranes, 1561 (2002) 188-201.

[388] J.-P. Colletier, B. Chaize, M. Winterhalter, D. Fournier, Protein encapsulation in

liposomes: efficiency depends on interactions between protein and phospholipid

bilayer, BMC biotechnology, 2 (2002) 9-9.

[389] A.P. Costa, X. Xu, D.J. Burgess, Freeze-Anneal-Thaw Cycling of Unilamellar

Liposomes: Effect on Encapsulation Efficiency, Pharmaceutical Research, 31 (2014)

97-103.

[390] N. Dos Santos, C. Allen, A.-M. Doppen, M. Anantha, K.A.K. Cox, R.C. Gallagher,

G. Karlsson, K. Edwards, G. Kenner, L. Samuels, M.S. Webb, M.B. Bally,

Influence of poly(ethylene glycol) grafting density and polymer length on

315

liposomes: Relating plasma circulation lifetimes to protein binding, Biochimica Et

Biophysica Acta-Biomembranes, 1768 (2007) 1367-1377.

[391] K.N. Burger, R.W. Staffhorst, H.C. de Vijlder, M.J. Velinova, P.H. Bomans, P.M.

Frederik, B. de Kruijff, Nanocapsules: lipid-coated aggregates of cisplatin with

high cytotoxicity, Nature Medicine, 8 (2002) 81-84.

[392] H. Shmeeda, Y. Amitay, D. Tzemach, J. Gorin, A. Gabizon, Liposome

encapsulation of zoledronic acid results in major changes in tissue distribution and

increase in toxicity, Journal of Controlled Release, 167 (2013) 265-275.

[393] K. Riviere, H.M. Kieler-Ferguson, K. Jerger, F.C. Szoka, Jr., Anti-tumor activity of

liposome encapsulated fluoroorotic acid as a single agent and in combination with

liposome irinotecan, Journal of Controlled Release, 153 (2011) 288-296.

[394] C. He, K. Lu, D. Liu, W. Lin, Nanoscale Metal-Organic Frameworks for the Co-

Delivery of Cisplatin and Pooled siRNAs to Enhance Therapeutic Efficacy in

Drug-Resistant Ovarian Cancer Cells, Journal of the American Chemical Society,

136 (2014) 5181-5184.

[395] H.K. Shete, R.H. Prabhu, V.B. Patravale, Endosomal Escape: A Bottleneck in

Intracellular Delivery, Journal of Nanoscience and Nanotechnology, 14 (2014) 460-

474.

[396] T. Daemen, G. Hofstede, M.T.T. Kate, I. Bakkerwoudenberg, G.L. Scherphof,

Liposomal doxorubicin-induced toxicity-depletion and impairment of phagocytic-

activity of liver macrophages, International Journal of Cancer, 61 (1995) 716-721.

[397] S. Kim, D.J. Kim, S.B. Howell, Modulation of the peritoneal clearance of

liposomal cytarabine by blank liposomes, Clinical Research, 35 (1987) A524-A524.

316

APPENDIX

APPENDIX A. NMR SPECTROSCOPY

Figure A.1. 1H NMR spectrum of compound A2 in CDCl3.

317

Figure A.2. 1H NMR (top) and 13C NMR (bottom) spectra of A3 in CDCl3.

318

Figure A.3. 1H NMR (top) and 13C NMR (bottom) spectra of A4 in CDCl3.

319

Figure A.4. 1H NMR of A5 in MeOH-d4.

320

Figure A.5. 1H NMR (top) and 13C NMR (bottom) spectra of A6 in CDCl3.

321

Figure A.6. 1H NMR (top) and 13C NMR (bottom) spectra of A7 in D2O.

322

Figure A.7. 1H NMR (top) and 13C NMR (bottom) spectra of A8 in CDCl3.

323

Figure A.8. 1H NMR (top) and 13C NMR (bottom) spectra of A9 in CDCl3.

324

Figure A.9. 1H NMR (top) and 13C NMR (bottom) spectra of A10 in CDCl3.

325

Figure A.10. 1H NMR spectra of P1 in D2O.

326

Figure A.11. 1H NMR spectra of P2 in D2O.

327

Figure A.12. 1H NMR spectra of P3 in D2O.

328

Figure A.13. 1H NMR spectra of P4 in D2O.

329

Figure A.14. 1H NMR spectra of P5 in D2O.

330

Figure A.15. 1H NMR spectra of P6 in D2O.

331

Figure A.16. 1H NMR (top) and 13C NMR (bottom) spectra of P1-A3 in DMF-d7.

332

Figure A.17. 1H NMR (top) and 13C NMR (bottom) spectra of P4-A3 in DMF-d7.

333

Figure A.18. 1H NMR (top) and 13C NMR (bottom) spectra of P6-A1 in DMF-d7.

334

Figure A.19. 1H NMR (top) and 13C NMR (bottom) spectra of P1-A8 in DMF-d7.

335

Figure A.20. 1H NMR spectra of P3-A7 in D2O.

336

Figure A.21. 1H NMR (top) and 13C NMR (bottom) spectra of 3-3 in CDCl3.

337

Figure A.22. 1H NMR (top) and 13C NMR (bottom) spectra of 3-5 in CDCl3.

338

Figure A.23. 1H NMR (top) and 13C NMR (bottom) spectra of P1-A9 in MeOH-d4.

339

Figure A.24. 1H NMR spectra of 4-6 in DMSO-d6.

340

Figure A.25. 1H NMR spectra of 4-7 in DMSO-d6.

341

Figure A.26. 1H NMR spectra of 4-3 in DMSO-d6.

342

Figure A.27. 1H NMR (top) and 13C NMR (bottom) spectra of 4-9 in CDCl3.

343

Figure A.28. 1H NMR (top) and 13C NMR (bottom) spectra of A11 in CDCl3.

344

Figure A.29. 1H NMR spectra of Boc-A3 in CDCl3.

345

Figure A.30. 1H NMR spectra of A12 in CDCl3.

346

Figure A.31. 1H NMR spectra of A13 in CDCl3.

347

Figure A.32. Variable Temperature 1H NMR spectrum of compound A12 in CDCl3 at -40 °C

(red), 0 °C (yellow), 25 °C (green), 35 °C (blue), 60 °C (purple), confirming interconversion of

the E- and Z-isomeric forms. Coalescence of endoxifen based peaks due to rapid interconversion

of E- and Z- isomers is highlighted by red, green and blue box.

a b

c

348

Figure A.33. Variable Temperature 1H NMR spectrum of compound A13 in CDCl3 at 25 °C

(red), 45 °C (green), 60 °C (blue).

349

Figure A.34. 1H NMR (top) and 13C NMR (bottom) spectra of P4-A3 in MeOH-d4.

350

Figure A.35. 1H NMR (top) and 13C NMR (bottom) spectra of P4-A12 in DMF-d7.

351

Figure A.36. 2-D 1H–13C HMBC spectra for compound P4-A12 in DMF-d7 giving

connectivities in the carbamate linker. The two sets of cross-peaks observed for several of the

proton and carbon nuclei is due to the presence of E- and Z-isomers, which were assigned in the

expanded view of the spectrum. No attempts were made to assign the two sets of signals to the

specific isomeric forms.

352

Figure A.37. 1H NMR (top) and 13C NMR (bottom) spectra of P4-A13 in DMF-d7.

353

Figure A.38. 1H NMR (top) and 13C NMR (bottom) spectra of 5-7 in CDCl3.

354

Figure A.39. 1H NMR (top) and 13C NMR (bottom) spectra of 5-8 in CDCl3.

355

Figure A.40. 1H NMR (top) and 13C NMR (bottom) spectra of 5-9 in CDCl3.

356

Figure A.41. 1H NMR (top) and 13C NMR (bottom) spectra of 5-10 in CDCl3.

357

Figure A.42. 1H NMR spectra of P9 in D2O.

358

Figure A.43. 1H NMR spectra of P10 in D2O.

359

Figure A.44. 1H NMR spectra of P11 in D2O.

360

Figure A.45. 1H NMR spectra of P12 in MeOH-d4.

361

Figure A.46. 1H NMR spectra of P13 in MeOH-d4.

362

Figure A.45. 1H NMR spectra of P14 in MeOH-d4.

363

Figure A.46. 1H NMR spectra of P15 in MeOH-d4.

364

Figure A.47. 1H NMR (top) and 13C NMR (bottom) spectra of P9-A1 in CDCl3.

365

Figure A.48. HMBC spectrum of compound P9-A1 and assignment of isomers.

366

Figure A.49. COSY spectrum of compound P9-A1.

367

Figure A.50. 1H NMR (top) and 13C NMR (bottom) spectra of P10-A1 in MeOH-d4.

368

Figure A.51. 1H NMR (top) and 13C NMR (bottom) spectra of P11-A1 in MeOH-d4.

369

Figure A.52. 1H NMR (top) and 13C NMR (bottom) spectra of P12-A1 in MeOH-d4.

370

Figure A.53. 1H NMR (top) and 13C NMR (bottom) spectra of P13-A1 in MeOH-d4.

371

Figure A.54. 1H NMR (top) and 13C NMR (bottom) spectra of P14-A1 in MeOH-d4.

372

Figure A.55. 1H NMR (top) and 13C NMR (bottom) spectra of P15-A1 in MeOH-d4.

373

Figure A.56. 1H NMR (top) and 13C NMR (bottom) spectra of P8-A1 in DMF-d7.

374

Figure A.57. Kinetic study using time-dependent 1H NMR spectra for the hydrolysis reaction of

complex P9-A1 in 10 mM PB (pH* 7.0) incubated at 37 ºC. Protons A, B and C of complex P9-

A1 were monitored and used for signal integrations.

375

Figure A.58. 1H NMR (top) and 13C NMR (bottom) spectra of P16-A1 in MeOH-d4.

376

Figure A.59. 1H NMR (top) and 13C NMR (bottom) spectra of P16-B1 in MeOH-d4.

377

Figure A.60. 1H NMR (top) and 13C NMR (bottom) spectra of 6-16 in DMF-d7.

378

Figure A.61. COSY spectrum of compound 6-16.

379

Figure A.62. HMBC spectrum of compound 6-16.

380

APPENDIX B. LC-MS SPECTROSCOPY

APPENDIX B1. LC-MS SPECTROSCOPY

Figure B1.1. LC-ESMS analysis of reaction P1 + A1.

381

Figure B1.2. LC-ESMS analysis of reaction P1 + A2.

382

Figure B1.3. LC-ESMS analysis of reaction P1 + A3.

383

Figure B1.4. LC-ESMS analysis of reaction P1 + A4.

384

Figure B1.5. LC-ESMS analysis of reaction P1 + A5.

385

Figure B1.6. LC-ESMS analysis of reaction P1 + A6.

386

Figure B1.7. LC-ESMS analysis of reaction P1 and A7.

387

Figure B1.8. LC-ESMS analysis of reaction P1 + A8.

388

Figure B1.9. LC-ESMS analysis of reaction P1 + A9.

389

Figure B1.10. LC-ESMS analysis of reaction P1 + A10.

390

Figure B1.11. LC-ESMS analysis of reaction P2 + A1.

391

Figure B1.12. LC-ESMS analysis of reaction P2 + A2.

392

Figure B1.13. LC-ESMS analysis of reaction P2 + A3.

393

Figure B1.14. LC-ESMS analysis of reaction P2 + A4.

394

Figure B1.15. LC-ESMS analysis of reaction P2 + A5.

395

Figure B1.16. LC-ESMS analysis of reaction P2 + A6.

396

Figure B1.17. LC-ESMS analysis of reaction P2 + A7.

397

Figure B1.18. LC-ESMS analysis of reaction P2 + A8.

398

Figure B1.19. LC-ESMS analysis of reaction P2 + A9.

399

Figure B1.20. LC-ESMS analysis of reaction P2 + A10.

400

Figure B1.21. LC-ESMS analysis of reaction P3 + A1.

401

Figure B1.22. LC-ESMS analysis of reaction P3 + A2.

402

Figure B1.23. LC-ESMS analysis of reaction P3 + A3.

403

Figure B1.24. LC-ESMS analysis of reaction P3 + A4.

404

Figure B1.25. LC-ESMS analysis of reaction P3 + A5.

405

Figure B1.26. LC-ESMS analysis of reaction P3 + A6.

406

Figure B1.27. LC-ESMS analysis of reaction P3 + A7.

407

Figure B1.28. LC-ESMS analysis of reaction P3 + A8.

408

Figure B1.29. LC-ESMS analysis of reaction P3 + A9.

409

Figure B1.30. LC-ESMS analysis of reaction P3 + A10.

410

Figure B1.31. LC-ESMS analysis of reaction P4 + A1.

411

Figure B1.32. LC-ESMS analysis of reaction P4 + A2.

412

Figure B1.33. LC-ESMS analysis of reaction P4 + A3.

413

Figure B1.34. LC-ESMS analysis of reaction P4 + A4.

414

Figure B1.35. LC-ESMS analysis of reaction P4 + A5.

415

Figure B1.36. LC-ESMS analysis of reaction P4 + A6.

416

Figure B1.37. LC-ESMS analysis of reaction P4 + A7.

417

Figure B1.38. LC-ESMS analysis of reaction P4 + A8.

418

Figure B1.39. LC-ESMS analysis of reaction P4 + A9.

419

Figure B1.40. LC-ESMS analysis of reaction P4 + A10.

420

Figure B1.41. LC-ESMS analysis of reaction P5 + A1.

421

Figure B1.42. LC-ESMS analysis of reaction P5 + A2.

422

Figure B1.43. LC-ESMS analysis of reaction P5 + A3.

423

Figure B1.44. LC-ESMS analysis of reaction P5 + A4.

424

Figure B1.45. LC-ESMS analysis of reaction P5 + A5.

425

Figure B1.46. LC-ESMS analysis of reaction P5 + A6.

426

Figure B1.47. LC-ESMS analysis of reaction P5 + A7.

427

Figure B1.48. LC-ESMS analysis of reaction P5 + A8.

428

Figure B1.49. LC-ESMS analysis of reaction P5 + A9.

429

Figure B1.50. LC-ESMS analysis of reaction P5 + A10.

430

Figure B1.51. LC-ESMS analysis of reaction P6 + A1.

431

Figure B1.52. LC-ESMS analysis of reaction P6 + A2.

432

Figure B1.53. LC-ESMS analysis of reaction P6 + A3.

433

Figure B1.54. LC-ESMS analysis of reaction P6 + A4.

434

Figure B1.55. LC-ESMS analysis of reaction P6 + A5.

435

Figure B1.56. LC-ESMS analysis of reaction P6 + A6.

436

Figure B1.57. LC-ESMS analysis of reaction P6 + A7.

437

Figure B1.58. LC-ESMS analysis of reaction P6 + A8.

438

Figure B1.59. LC-ESMS analysis of reaction P6 + A9.

439

Figure B1.60. LC-ESMS analysis of reaction P6 + A10.

440

Figure B1.61. Percent conversion in ‘click’ reactions for platinum-acridines. Compounds are

sorted and color-coded by common acridine moieties.

441

APPENDIX B2. LC-ESMS ANALYSIS OF PURIFIED COMPOUND

Figure B2.1. LC-MS analysis of compound A1.

1

04271219.D: UV Chromatogram, 363-463 nm

0

25

50

75

100

125

Intens.

mAU

0 5 10 15 20 25 Time [min]

195.2

252.2

+MS, 10.9min #575

0

1

2

3

4

7x10

Intens.

150 175 200 225 250 275 300 325 350 375 m/z

442

Figure B2.2. LC-MS analysis of compound A2.

1

05021201.D: UV Chromatogram, 363-463 nm

0

20

40

60

80

Intens.

mAU

0 5 10 15 20 25 Time [min]

266.2

+MS, 11.9min #661

0

1

2

3

7x10

Intens.

220 240 260 280 300 320 340 m/z

443

Figure B2.3. LC-MS analysis of compound A3.

1

04271212.D: UV Chromatogram, 363-463 nm

0

50

100

150

Intens.

mAU

0 5 10 15 20 25 Time [min]

195.3

282.3

+MS, 11.0min #616

0.0

0.2

0.4

0.6

0.8

1.0

8x10

Intens.

150 200 250 300 350 400 m/z

444

Figure B2.4. LC-MS analysis of compound A4.

1

04271213.D: UV Chromatogram, 363-463 nm

0

20

40

60

80

Intens.

mAU

0 5 10 15 20 25 Time [min]

195.2

296.3

+MS, 11.3min #625

0

2

4

6

7x10

Intens.

150 200 250 300 350 400 m/z

445

Figure B2.5. LC-MS analysis of compound A5.

1

04271214.D: UV Chromatogram, 363-463 nm

0

10

20

30

40

Intens.

mAU

0 5 10 15 20 25 Time [min]

296.3

+MS, 13.4min #734

0.0

0.5

1.0

1.5

7x10

Intens.

150 200 250 300 350 400 m/z

446

Figure B2.6. LC-MS analysis of compound A6.

1

04271221.D: UV Chromatogram, 363-463 nm

0

5

10

15

20

Intens.

mAU

0 5 10 15 20 25 Time [min]

259.3

310.2

+MS, 12.4min #659

0.00

0.25

0.50

0.75

1.00

1.25

6x10

Intens.

150 200 250 300 350 400 m/z

447

Figure B2.7. LC-MS analysis of compound A7.

1

04271216.D: UV Chromatogram, 363-463 nm

0

5

10

15

20

Intens.

mAU

0 5 10 15 20 25 Time [min]

259.3

267.3 275.3

279.3

283.3

297.3

310.3

325.4 335.3 341.3 355.3

+MS, 14.6min #794

0

1

2

3

4

5

5x10

Intens.

260 280 300 320 340 m/z

448

Figure B2.8. LC-MS analysis of compound A8.

1

04271217.D: UV Chromatogram, 363-463 nm

0

5

10

15

20

25

Intens.

mAU

0 5 10 15 20 25 Time [min]

364.3

+MS, 13.3min #715

0

2

4

6

7x10

Intens.

200 250 300 350 400 450 500 m/z

449

Figure B2.9. LC-MS analysis of compound A9.

1

05021202.D: UV Chromatogram, 363-463 nm

0

10

20

30

40

Intens.

mAU

0 5 10 15 20 25 Time [min]

378.3

+MS, 13.1min #728

0.0

0.2

0.4

0.6

0.8

1.0

8x10

Intens.

200 250 300 350 400 450 500 m/z

450

Figure B2.10. LC-MS analysis of compound A10.

451

Figure B2.11. LC-MS analysis of compound P1-A3.

1

05261207.D: UV Chromatogram, 363-463 nm

0.0

2.5

5.0

7.5

10.0

12.5

Intens.

mAU

0 5 10 15 20 25 Time [min]

169.3

195.2

207.2

221.2

250.2264.2

282.2

295.3

320.2

337.2

363.3

396.2

440.1 473.1

499.2

529.2589.3

627.3

670.2

+MS, 14.0min #756

0.0

0.2

0.4

0.6

0.8

1.0

1.2

6x10

Intens.

200 300 400 500 600 700 m/z

452

Figure B2.12. LC-MS analysis of compound P4-A3.

1

05261205.D: UV Chromatogram, 363-463 nm

0

5

10

15

20

Intens.

mAU

0 5 10 15 20 25 Time [min]

159.2

195.2

207.2

221.2

239.2

252.2

265.6279.2

305.2

327.3

344.3

363.2

391.3

413.3

453.3

463.2

474.2

497.3

513.2

524.3

539.2

561.2

585.2

616.7632.4

653.1

680.7

698.2

718.6

+MS, 10.4min #567

0.0

0.2

0.4

0.6

0.8

1.0

5x10

Intens.

200 300 400 500 600 700 m/z

453

Figure B2.13. LC-MS analysis of compound P6-A1.

1

06011211.D: UV Chromatogram, 363-463 nm

0

10

20

30

Intens.

mAU

0 5 10 15 20 25 Time [min]

195.2

221.2

252.2

288.2

307.2

361.1

391.4 415.1428.3

485.2

514.2575.3

612.3

+MS, 13.4min #737

0.0

0.5

1.0

1.5

2.0

2.5

6x10

Intens.

200 300 400 500 600 700 m/z

454

Figure B2.14. LC-MS analysis of compound P1-A8.

1

02091211.D: UV Chromatogram, 363-463 nm

0.0

2.5

5.0

7.5

10.0

12.5

Intens.

mAU

0 5 10 15 20 25 Time [min]

195.1 221.1279.2

333.4

354.8

391.3

419.4

708.5

+MS, 16.6min #949

0.0

0.2

0.4

0.6

0.8

1.0

1.2

7x10

Intens.

200 300 400 500 600 700 800 900 m/z

455

Figure B2.15. LC-MS analysis of compound P3-A7.

1

0 5 10 15 20 25 Time [min]

0

5

10

15

20

25

Intens.

mAU

06121202.D: UV Chromatogram, 363-463 nm

159.2

195.2

207.2221.2

248.2264.1

279.2

288.2

310.2

320.1 399.1

415.1

442.1458.1

485.1

501.1

557.2

575.2611.2

629.2

+MS, 14.3min #774

0.00

0.25

0.50

0.75

1.00

1.25

6x10

Intens.

200 300 400 500 600 700 m/z

456

Figure B2.16. LC-MS analysis of compound P1-A9.

102091207.D: UV Chromatogram, 363-463 nm

0

10

20

30

40

Intens.

mAU

0 5 10 15 20 25 Time [min]

195.1 221.2293.3

329.7343.8

361.8

433.4

722.4

+MS, 15.1min #870

0

1

2

3

4

5

6

7x10

Intens.

200 300 400 500 600 700 m/z

457

Figure B2.17. LC-ESMS analysis of compound P4-A3. (A) Reverse-phase HPLC trace of

compound P4-A3. (B) ESMS spectrum of compound 5 (HPLC fraction with retention time 10.0

min) recorded in positive-ion mode. Characteristic molecular and fragment ions: [M]+ (m/z

585.2).

1

05261205.D: UV Chromatogram, 363-463 nm

0

5

10

15

20

Intens.

mAU

0 5 10 15 20 25 Time [min]

159.2

195.2

207.2

221.2

239.2

252.2

265.6279.2

305.2

327.3

344.3

363.2

391.3

413.3

453.3

463.2

474.2

497.3

513.2

524.3

539.2

561.2

585.2

616.7632.4

653.1

680.7

698.2

718.6

+MS, 10.4min #567

0.0

0.2

0.4

0.6

0.8

1.0

5x10

Intens.

200 300 400 500 600 700 m/z

A

B

B

458

Figure B2.18. LC-ESMS analysis of compound A12. (A) Reverse-phase HPLC trace of

compound A12. (B) ESMS spectrum of compound 9 (HPLC fraction with retention time 34.0

min) recorded in positive-ion mode. Characteristic molecular and fragment ions: [M]+ (m/z

681.5).

681.5

459.4

552.4 283.4

A

B

459

A

460

Figure B2.19. LC-ESMS analysis of compound P4-A12. (A) Reverse-phase HPLC trace of

compound P4-A12. (B) ESMS spectrum of compound P4-A12 (HPLC fraction with retention

time 33.8 min) recorded in positive-ion mode. Characteristic molecular and fragment ions: [M-

2H]+ (m/z 985.6), [M-2NH3-Pt-Cl-MeCN]+ (m/z 681.5), [M-H]2+ (m/z 493.3, z = 2), [M-2NH3-

Cl]2+ (m/z 457.8, z = 2), [M-2NH3-Cl-Et-OH]2+ (m/z 436.9, z =2).

436.9

457.8

493.3

681.5

+MS, 33.9min #1897

0

1

2

3

4

5

7x10

Intens.

300 400 500 600 700 800 900 1000 m/z

985.6

B

461

Figure B2.20. LC-ESMS analysis of compound A13. (A) Reverse-phase HPLC trace of

compound A13. (B) ESMS spectrum of compound A13 (HPLC fraction with retention time 35.6

min) recorded in positive-ion mode. Characteristic molecular and fragment ions: [M+H]+ (m/z

769.8), [M+H]2+ (m/z 384.5, z=2).

A

B

462

1

463

Figure B2.21 continued.

464

Figure B2.21 continued.

465

Figure B2.21. LC-ESMS analysis of compound P4-A13. (1) Reverse-phase HPLC trace of

compound P4-A13. Peaks 2–5 show identical masses and can be assigned to rotational isomers

as a result of hindered rotation within the amidine and (E,Z)-endoxifen moieties. Peak 1 is an

unidentified impurity. Combined % purity for peaks 2–5 (rotamers): 97%. (2) ESMS spectrum

of one species of compound 10’ (HPLC fraction with retention time 35.2 min) recorded in

positive-ion mode. Characteristic molecular and fragment ions: [M+H]+ (m/z 1074.8), [M+H]2+

(m/z 537.5, z = 2), [M-2NH3-Cl]2+ (m/z 502.4, z = 2). (3) ESMS spectrum of one species of

compound 10’ (HPLC fraction with retention time 36.4 min) recorded in positive-ion mode.

Characteristic molecular and fragment ions: [M+H]+ (m/z 1074.8), [M]2+ (m/z 537.5, z = 2), [M-

2NH3-Cl]2+ (m/z 502.4, z = 2). (4) ESMS spectrum of one species of compound 10’ (HPLC

fraction with retention time 37.2 min) recorded in positive-ion mode. Characteristic molecular

and fragment ions: [M]+ (m/z 1073.8), [M+H]2+ (m/z 537.8, z = 2), [M-2NH3-Cl]2+ (m/z 502.4, z

= 2). (5) ESMS spectrum of one species of compound 10’ (HPLC fraction with retention time

37.2 min) recorded in positive-ion mode. Characteristic molecular and fragment ions: [M]+ (m/z

1073.8), [M]2+ (m/z 537.4, z = 2), [M-2NH3-Cl]2+ (m/z 502.4, z = 2).

466

Figure B2.22. LC-MS analysis of compound P9-A1.

467

Figure B2.23. LC-MS analysis of compound P10-A1.

468

Figure B2.24. LC-MS analysis of compound P11-A1.

469

Figure B2.25. LC-MS analysis of compound P12-A1.

470

Figure B2.26. LC-MS analysis of compound P13-A1.

471

Figure B2.27. LC-MS analysis of compound P14-A1.

472

Figure B2.28. LC-MS analysis of compound P15-A1.

473

Figure B2.29. LC-MS analysis of compound P8-A1.

474

Figure B2.30. LC-MS analysis of compound P16-A1.

475

Figure B2.31. LC-MS analysis of compound P16-B1.

476

Figure B2.32. LC-MS analysis of compound 6-16.

477

APPENDIX B3. LC-MS ANALYSIS EVALUATION OF STABILITY OF CONJUGATES IN

BUFFERS

Figure B3.1. LC-ESMS analysis of the reaction between 3-3 and P1 suggesting decomposition

of the target compound (peak 4) via loss of N2 and N-chelate formation (peak 1) along with an

unidentified platinum-containing product (peak 3).

478

Figure B3.2. LC-ESMS analysis of the reaction mixture between 3-5 and P1 suggesting

decomposition of the target compound (peak 4) via loss of N2 and N-chelate formation (peak 1)

along with unidentified platinum-containing products (peaks 3 and 5).

479

Figure B3.3. LC-ESMS analysis of the mixture of compound P1-A9 in 10 mM PBS (pH 7.4)

after incubation for 24 h at 37 °C, showing minor aquation of the chloro leaving group. The

azide group and amide linkage are stable under these conditions.

480

Figure B3.4. LC-ESMS analysis of the mixture of compound P1-A9 in 10 mM PB (pH 7.4)

after incubation for 24 h at 37 °C, showing aquation and substitution of chloride by phosphate.

481

Figure B3.5. LC-ESMS analysis of the mixture of compound P1-A9 in 10 mM acetate buffer

(pH 5.0) incubating for 24 h at 37 °C.

482

Figure B3.6. LC-ESMS analysis of the mixture of compound P4-A12 in 10 mM acetate buffer

(pH = 5.0) incubating for 48 h at 37 °C. ESMS spectrum was recorded in positive-ion mode.

Characteristic molecular and fragment ions: [M-2NH3-Pt-Cl]+ (m/z 722.4), [M-Cl+MeCOO]2+

(m/z 505.3, z =2), [M-H]2+ (m/z 493.3, z = 2), [M-2NH3-Cl+H]2+ (m/z 458.2, z = 2), [M-2NH3-

Cl-Et-OH]2+ (m/z 436.9, z =2). (C) Compound P4-A12 is chemically stable in the incubation

condition. Only the displacement of chloride ligand with buffer was observed.

195.2436.3

458.2

505.3

722.4

+MS, 12.5min #692

0.0

0.5

1.0

1.5

2.0

7x10

Intens.

200 300 400 500 600 700 800 900 1000 m/z

C

A

B

483

Figure B3.7. LC-ESMS analysis of the mixture of compound P4-A12 in 10 mM phosphate

buffer saline (pH = 7.4) incubating for 48 h at 37 °C. ESMS spectrum was recorded in positive-

ion mode. Characteristic molecular and fragment ions: [M-Cl+H2PO4]2+ (m/z 524.3, z =2), [M-

H]2+ (m/z 493.3, z = 2), [M-2NH3-Cl]2+ (m/z 458.3, z = 2), [M-2NH3-Cl-Et-OH]2+ (m/z 436.2, z

=2). (C) Compound P4-A12 is chemically stable in the incubation condition. The displacement

of chloride ligand with buffer was observed, while dramatically inhibited by the presence of

chloride ions in buffer.

436.2

458.3

475.3

493.7

524.2

+MS, 12.6min #693

0.00

0.25

0.50

0.75

1.00

1.25

7x10

Intens.

200 300 400 500 600 700 800 900 1000 m/z

A

B

C

484

Figure B3.8. LC-ESMS analysis of the mixture of compound P4-A12 in 10 mM phosphate

buffer (pH = 7.4) incubating for 48 h at 37 °C. ESMS spectrum was recorded in positive-ion

mode. Characteristic molecular and fragment ions : [M-2NH3-Pt-Cl-MeCN]+ (m/z 681.4), [M-

Cl+H2PO4]2+ (m/z 524.3, z =2), [M-H]2+ (m/z 493.3, z = 2), [M-2NH3-Cl]2+ (m/z 457.8, z = 2),

[M-2NH3-Cl-Et-OH]2+ (m/z 436.2, z =2). (C) Compound P4-A12 is chemically stable in the

incubation condition. Only the displacement of chloride ligand with buffer was observed.

391.3419.4

436.2

457.8

475.3 493.3

524.3

537.7550.2

681.4

+MS, 12.7min #697

0.0

0.5

1.0

1.5

2.0

2.5

3.0

6x10

Intens.

350 400 450 500 550 600 650 700 750 800 m/z

A

B

C

485

APPENDIX B4. METAL-ASSISTED ESTER HYDROLYSIS

Figure B4.1. LC-ESMS analysis of the mixture of compound P10-A1 in phosphate buffer (PB,

pH 7.4) incubated at 37 °C.

486

Figure

B4.2. LC-ESMS analysis of the mixture of compound P11-A1 in phosphate buffer (PB, pH 7.4)

incubated at 37 °C.

487

Figure B4.3. LC-ESMS analysis of the mixture of compound P12-A1 in phosphate buffer (PB,

pH 7.4) incubated at 37 °C.

488

Figure B4.4. LC-ESMS analysis of the mixture of compound P13-A1 in phosphate buffer (PB,

pH 7.4) incubated at 37 °C.

489

Figure B4.5. LC-ESMS analysis of the mixture of compound P14-A1 in phosphate buffer (PB,

pH 7.4) incubated at 37 °C.

490

Figure B4.6. LC-ESMS analysis of the mixture of compound P15-A1 in phosphate buffer (PB,

pH 7.4) incubated at 37 °C.

491

Figure B4.7. LC-ESMS analysis of the mixture of compound 6-12 in phosphate buffer (PB, pH

7.4) incubated at 37 °C.

492

Figure B4.8. LC-ESMS analysis of the mixture of compound 6-14 in phosphate buffer (PB, pH

7.4) incubated at 37 °C.

493

Figure B4.9. LC-ESMS analysis of the mixture of compound 6-16 in phosphate buffer (PB, pH

7.4) incubated at 37 °C.

494

Figure B4.10. LC-ESMS analysis of the mixture of compound 6-17 in phosphate buffer (PB,

pH 7.4) incubated at 37 °C.

495

Figure B4.11. LC-ESMS analysis of the mixture of compound 6-18 in phosphate buffer (PB,

pH 7.4) incubated at 37 °C.

496

Figure B4.12. LC-ESMS analysis of the mixture of compound 6-22 in phosphate buffer (PB,

pH 7.4) incubated at 37 °C.

497

Figure B4.13. LC-ESMS analysis of the mixture of compound P10-A1 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

498

Figure B4.14. LC-ESMS analysis of the mixture of compound P11-A1 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

499

Figure B4.15. LC-ESMS analysis of the mixture of compound P12-A1 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

500

Figure B4.16. LC-ESMS analysis of the mixture of compound P13-A1 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

501

FigureB4.17. LC-ESMS analysis of the mixture of compound P14-A1 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

502

Figure B4.18. LC-ESMS analysis of the mixture of compound P15-A1 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

503

Figure B4.19. LC-ESMS analysis of the mixture of compound 6-12 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

504

Figure B4.20. LC-ESMS analysis of the mixture of compound 6-14 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

505

Figure B4.21. LC-ESMS analysis of the mixture of compound 6-16 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

506

Figure B4.22. LC-ESMS analysis of the mixture of compound 6-17 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

507

Figure B4.23. LC-ESMS analysis of the mixture of compound 6-18 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

508

Figure B4.24. LC-ESMS analysis of the mixture of compound 6-22 in phosphate buffered

saline (PBS, pH 7.4) incubated at 37 °C.

509

APPENDIX B5. hCES-2 MEDIATED ESTER HYDROLYSIS

Figure B5.1. LC-ESMS analysis of the enzymatic cleavage of compound P9-A1 in phosphate

buffered saline (PBS, pH 7.4) incubated at 37 °C.

510

Figure B5.2. LC-ESMS analysis of the enzymatic cleavage of compound P10-A1 in phosphate

buffered saline (PBS, pH 7.4) incubated at 37 °C.

511

Figure B5.3. LC-ESMS analysis of the enzymatic cleavage of compound P11-A1 in phosphate

buffered saline (PBS, pH 7.4) incubated at 37 °C.

512

Figure B5.4. LC-ESMS analysis of the enzymatic cleavage of compound P12-A1 in phosphate

buffered saline (PBS, pH 7.4) incubated at 37 °C.

513

Figure B5.5. LC-ESMS analysis of the enzymatic cleavage of compound P13-A1 in phosphate

buffered saline (PBS, pH 7.4) incubated at 37 °C.

514

Figure B5.6. LC-ESMS analysis of the enzymatic cleavage of compound P14-A1 in phosphate

buffered saline (PBS, pH 7.4) incubated at 37 °C.

515

Figure B5.7. LC-ESMS analysis of the enzymatic cleavage of compound P15-A1 in phosphate

buffered saline (PBS, pH 7.4) incubated at 37 °C.

516

APPENDIX B6. LC-MS ANALYSIS OF THE COUPLING REACTIONS FOR

MULTIFUNCTIONAL PLATINUM-ACRIDINES

Figure B6.1. LC-MS analysis of the reaction mixture for the preparation of 6-11.

517

Figure B6.2. LC-MS analysis of the reaction mixture for the preparation of 6-12.

518

Figure B6.3. LC-MS analysis of the reaction mixture for the preparation of 6-13.

519

Figure B6.4. LC-MS analysis of the reaction mixture for the preparation of 6-14.

520

Figure B6.5. LC-MS analysis of the reaction mixture for the preparation of 6-15.

521

Figure B6.6. LC-MS analysis of the reaction mixture for the preparation of 6-17.

522

Figure B6.7. LC-MS analysis of the reaction mixture for the preparation of 6-18.

523

Figure B6.8. LC-MS analysis of the reaction mixture for the preparation of 6-19.

524

Figure B6.9. LC-MS analysis of the reaction mixture for the preparation of 6-20.

525

Figure B6.10. LC-MS analysis of the reaction mixture for the preparation of 6-21.

526

Figure B6.11. LC-MS analysis of the reaction mixture for the preparation of 6-22.

527

Figure B6.12. LC-MS analysis of the reaction mixture for the preparation of 6-23.

528

Figure B6.13. LC-MS analysis of the reaction mixture for the preparation of 6-24.

529

APPENDIX C. CD SPECTRA

Figure C.1. Circular dichroism (CD) spectra (ICD region) of calf thymus DNA modified

with platinum-acridines P1-A1 (blue trace) and P1-A9 (red trace). CD spectra were

recorded for ~1 mM (n.t) DNA in 10 mM phosphate buffer (25 °C, pH 7.4) at a drug-to-

nucleotide ratio (rb) of 0.05.

530

APPENDIX D. CONFOCAL MICROSCOPIC IMAGES

Figure D.1. Acridine fluorescence in NCI-H460 cells (excitation wavelength: 405 nm,

emission wavelength range: 414-470 nm) after treatment with 5 µM compound P1-A9

(“Pt-Azide”) for 3 h.

Figure D.2. Total background-corrected fluorescent intensities in cells treated with

compound P1-A9 (1, 5, 25 μM) and in untreated cells. Plotted data are the mean

fluorescence intensities (± S.D.) determined for n > 50 cells. The data was analyzed with

one-way ANOVA (***P < 0.001; P < 0.05 was considered significant).

******

***

0

500

1000

1500

2000

2500

3000

3500

control 1 uM 5 uM 25 uM

Flu

ore

sce

nce

at

48

8 n

m (

arb

itr.

un

its)

531

Figure D.3. (a) Background-corrected fluorescent intensities in nuclei of cells treated

with compound P1-A9 (5 μM) and in untreated cells. (b) Nucleolus-to-nucleus

fluorescence intensity ratios (Fnucleolus/Fnucleus) in cells treated with compound P1-A9 (5

μM) and in untreated cells. Plotted data are the mean fluorescence intensities (± S.D.)

determined for n = 40–50 cells. The data was analyzed with one-way ANOVA (***P <

0.001; P < 0.05 was considered significant).

532

Figure D.4. Enlarged view of a colocalization image of a single NCI-H460 cancer cell

treated with P1-A9 and P1-A9* (green), followed by post-labeling with CuAAC/Alexa

Fluor 647-alkyne (red). Hoechst 33342 (blue), merged, and bright field channel images

are also shown. Scale bar = 5 µM. Note the high degree of colocalization of the green-

and red-fluorescent dyes in the nucleolus (and in two unidentified cytosolic structures),

and the markedly reduced relative fluorescence intensity in the nucleus in the green

channel vs the red channel.

533

APPENDIX E. CELL PROLIFERATION ASSAYS

Figure E.1. Drug-response curves for cell proliferation assays in NCI-H460 cells treated

with selected compounds. Error bars indicate ± standard deviations from the mean for

two independent experiments performed in triplicate.

Figure E.2. Drug-response curves for cell proliferation assays in NCI-H460 cells treated

with P3-A7 and the corresponding acridine ligand A7. Error bars indicate ± standard

deviations from the mean for two independent experiments performed in triplicate.

534

Figure E.3. Dose-response curves for cell proliferation assays in non-small cell lung

cancer cell line (A) NCI-A549 and (B) NCI-H460 treated with compound P8-A1. Error

bars indicate ± standard deviations from the mean for two independent experiments

performed in triplicate.

535

APPENDIX F. COMPUTATION STUDIES

Figure F.1. Molecular models (AMBER) of the monofunctional–intercalative adducts at

guanine–N7 in the DNA major groove formed by compound 1 (left) and compound 2

(right) based on the NMR solution structure of the hybrid adduct (PDB 1XRW). The

platinum–acridine and the newly installed azide linker are shown in red and element-

specific colors, respectively.

536

APPENDIX G. MTD STUDIES

Figure G.1. Maximum tolerated dose (MTD) study of compound P15-A1 in 420 Swiss-

Webster (SW) mice (6-8 weeks, 18-25 g, female, n = 5). Compound P15-A1 was

administered i.p. at the dose of 1.6 mg/kg on (A) days 0, 4, 8, and 12, and (B) days 0, 1, 2,

3, and 4. No statistically significant weight loss is observed in both treatment groups.

537

APPENDIX H. SIZE DISTRIBUTION OF LIPOSOMES

Figure H.1. Size distribution of Lipo-1 determined by DLS.

Figure H.2. Size distribution of Lipo-2 determined by DLS.

538

Figure H.3. Size distribution of Lipo-3 determined by DLS.

Figure H.4. Size distribution of Lipo-4 determined by DLS.

539

APPENDIX I.CHEMICAL STRUCTURES OF COUPLING REAGENTS USED IN

CHAPTER 6

540

APPENDIX J.CHEMICAL STRUCTURES OF LIPIDS USED IN CHAPTER 7

541

SCHOLASTIC VITA

Song Ding

BORN: January 6, 1985

Nanjing, China

UNDERGRADUATE STUDY: China Pharmaceutical University

Nanjing, China

B.S. Basic Pharmacy, 2007

GRADUATE STUDY: China Pharmaceutical University

Nanjing, China

M.S. Medicinal Chemistry, 2010

GRADUATE STUDY: Wake Forest University

Winston-Salem, North Carolina

Ph.D. candidate, 2015

SCHOLASTIC AND PROFESSIONAL EXPERIENCE:

Teaching Assistant, Wake Forest University 2014-

2015

Winston-Salem, North Carolina

Research Assistant, Wake Forest University 2010-

2014

Winston-Salem, North Carolina

HORNORS AND AWARDS:

GlaxoSmithKline Graduate Research Award 2014

542

Department of Chemistry, Wake Forest University

Wake Forest Innovation and Commercialization Services, SPARK Award (for

development of a liposomal formulation for delivery of platinum–acridine anticancer

agents, Co-PI with U. Bierbach)

8/2013– 7/2014

PUBLICATIONS:

1. Ding, S.; Bierbach,U. et al, Development of a chemical toolbox for discovery of

multifunctional platinum-acridine agents. In preparation.

2. Fahrenholtz C.D., Ding S., Bernish B.W., Wright M.L., Bierbach U., Singh R.N.,

Self-assembling Platinum-acridine loaded carbon nanotubes for cancer chemotherapy.

Submitted to Scientific reports.

3. Ding, S.; Bierbach,U. et al, Liposomal delivery of platinum-acridine anticancer

agents for non-small cell lung cancer. In preparation.

4. Ding, S.; Bierbach, U. Target-selective delivery and activation of platinum-based

anticancer

agents. Future Medicinal Chemistry 2015,7(7), 911-927.

5. Ding, S.; Pickard A.J.; Kucera G.L.; Bierbach U., Design of Enzymatically Cleavable

Prodrugs of a Potent Platinum-Containing Anticancer Agent. Chemistry – A European

Journal 2014, 20 (49), 16164-16173.

6. Qiao, X.; Ding, S.; Liu, F.; Kucera, G. L.; Bierbach, U., Investigating the cellular fate

of a DNA-targeted platinum-based anticancer agent by orthogonal double-click

chemistry. Journal of Biological Inorganic Chemistry 2014, 19 (3), 415-26.

7. Ding, S.; Qiao, X.; Kucera, G. L.; Bierbach, U., Design of a platinum-acridine-

endoxifen conjugate targeted at hormone-dependent breast cancer. Chemical

Communications 2013, 49 (24), 2415-2417;

8. Ding, S.; Qiao, X.; Suryadi, J.; Marrs, G. S.; Kucera, G. L.; Bierbach, U., Using

Fluorescent Post-Labeling To Probe the Subcellular Localization of DNA-Targeted

Platinum Anticancer Agents. Angewandte Chemie-International Edition 2013, 52

(12), 3350-3354;

9. Ding, S.; Qiao, X.; Kucera, G. L.; Bierbach, U., Using a Build-and-Click Approach

for Producing Structural and Functional Diversity in DNA-Targeted Hybrid

Anticancer Agents. Journal of Medicinal Chemistry 2012, 55 (22), 10198-10203;

PATENTS:

1. Bierbach U., Ding S.: Targeted delivery and prodrug designs for platinum-acridine

anti-cancer compounds and methods thereof. PCT/US2012/053189 (filed 08/30/2012).

Nationalization in US: #US 14/241,900, July 10, 2014.

2. Bierbach, U., Ding, S.: Enzymatically activatable prodrugs of DNA-targeted

platinum-acridine agents and methods thereof. Provisional patent application #

62/015,478 (filed 06/22/2014).

543

PRESENTATION:

1. Ding, S.; Qiao, X.; Kucera, G. L.; Bierbach, U., Development and evaluation of

novel conjugation chemistries for the targeted delivery and cellular imaging of

potent platinum-acridine anticancer agents. 243th ACS National Meeting, San

Diego, CA, March 28, 2012. (Poster Presentation)

2. Ding, S.; Qiao, X.; Kucera, G. L.; Bierbach, U., Novel conjugation chemistries for

the functionalization and cellular imaging of potent platinum- acridine anticancer

agents. SERMACS 2012, Raleigh, NC, November 15, 2012. (Oral presentation)

3. Ding, S.; Kucera, G. L.; Bierbach, U., Novel strategies for improving the

pharmacological properties of platinum-acridine anticancer agents. 249th ACS

National Meeting, San Diego, CA, March 25, 2015. (Poster Presentation)