synthesis, structural characterisation and cytotoxicity - nova

377
SYNTHESIS, STRUCTURAL CHARACTERISATION AND CYTOTOXICITY STUDY OF TIN(IV) COMPOUNDS CONTAINING ONS SCHIFF BASES By ENIS NADIA BINTI MD YUSOF Thesis Submitted to the School of Graduate Studies, Universiti Putra Malaysia, Malaysia and The University of Newcastle, Australia in Fulfilment of the Requirements for the Degree of Doctor of Philosophy December 2019

Upload: khangminh22

Post on 11-Apr-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

SYNTHESIS, STRUCTURAL CHARACTERISATION AND CYTOTOXICITY STUDY OF TIN(IV) COMPOUNDS CONTAINING ONS SCHIFF BASES

By

ENIS NADIA BINTI MD YUSOF

Thesis Submitted to the School of Graduate Studies, Universiti Putra Malaysia, Malaysia and The University of Newcastle, Australia in Fulfilment of the

Requirements for the Degree of Doctor of Philosophy

December 2019

All material contained within the thesis, including without limitation text, logos, icons, photographs and all other artwork, is copyright material of Universiti Putra Malaysia unless otherwise stated. Use may be made of any material contained within the thesis for non-commercial purposes from the copyright holder. Commercial use of material may only be made with the express, prior, written permission of Universiti Putra Malaysia. Copyright © Universiti Putra Malaysia

i

Abstract of thesis presented to the Senate of Universiti Putra Malaysia and The University of Newcastle in fulfilment of the requirement for the degree of Doctor of

Philosophy

SYNTHESIS, STRUCTURAL CHARACTERISATION AND CYTOTOXICITY STUDY OF TIN(IV) COMPOUNDS CONTAINING ONS SCHIFF BASES

By

ENIS NADIA BINTI MD YUSOF

December 2019

Chair : Associate Professor Thahira Begum, PhD (UPM) Associate Professor Alister J. Page, PhD (UON) Faculty : Science There is an urgent need for substantial investigation of non-platinum drugs with higher

activity and improved selectivity to address the problem associated with the use of

platinum-based compounds as therapeutic agents. In light of this, diphenyltin(IV),

dimethyltin(IV) and tin(IV) compounds were synthesised from the Schiff bases of three

series of dithiocarbazate (S-2-methylbenzyldithiocarbazate (S1), S-4-methylbenzyl

dithiocarbazate (S2), S-benzyldithiocarbazate (S3)) and two series of thiosemicarbazides

(4-methyl-3-thiosemicarbazide and 4-phenyl-3-thiosemicarbazide) with aldehydes, 2-

hydroxy-3-methoxybenzaldehyde (oVa) or 2,3-dihydroxybenzaldehyde (catechol). The

tin(IV) compounds formed were found to have a general formula of [R2Sn(ONS)] and

[Sn(ONS)2] (where R = Me and Ph). The compounds were fully characterised by physico-

chemical and spectroscopic methods. The spectroscopic results supported the coordination

geometry in which the Schiff bases behaved as tridentate ONS donor ligands coordinating

via azomethine nitrogen, thiolo sulphur and phenoxide oxygen atoms. A total of 11 crystal

ii

structures of the expected compounds were solved in this work. In order to verify the

experimental data, the compounds were optimised using the density functional theory

(DFT) method with the B3LYP hybrid exchange correlation functional with LanL2DZ

pseudopotential on tin and 6-311G(d,p) Pople basis set for all other atoms. Diphenyltin(IV)

compounds showed the most promising cytotoxicity with IC50 values ranging between

0.016 – 4.40 μM against a panel of twelve cancer cell lines (RT-112, EJ-28 (bladder), HT29

(colon), U87, SJ-G2, SMA (glioblastoma), MCF-7 (breast), A2780 (ovarian), H460 (lung),

A431 (skin), Du145 (prostate), BE2-C (neuroblastoma) and MIA (pancreatic)). The three

diphenyltin(IV) compounds of the oVa series were able to induce the production of

Reactive Oxygen Species (ROS) and acted as a cell apoptosis inducer. Good binding

interactions for all the diphenyltin(IV) compound series were observed and supported by

molecular docking analysis, where hydrogen, electrostatic and hydrophobic binding

interactions were observed. This highlights the important of two phenyl groups coordinated

directly to the tin ion to enhance the cytotoxicity by strong π-π stacking interactions to

biomacromolecules. Diphenyltin(IV) compounds could bring hope in the field of drug

development against various diseases including cancers.

iii

Abstrak tesis yang dikemukakan kepada Senat Universiti Putra Malaysia dan The University of Newcastle sebagai memenuhi keperluan untuk Ijazah Doktor Falsafah

SINTESIS, PENCIRIAN STRUKTUR DAN KAJIAN SITOTOSIK BAGI SEBATIAN TIN(IV) YANG MENGANDUNGI BES SCHIFF ONS

Oleh

ENIS NADIA BINTI MD YUSOF

Disember 2019

Pengerusi : Professor Madya Thahira Begum, PhD (UPM) Professor Madya Alister J. Page, PhD (UON) Fakulti : Sains Terdapat keperluan segera bagi penyiasatan penting ke atas ubat yang tidak mengandungi

platinum dengan aktitviti yang tinggi dan meningkatkan pemilihan untuk menyelesaikan

masalah berkaitan dengan penggunaan sebatian daripada platinum sebagai agen

terapeutik. Disebabkan keadaan ini, sebatian difenilstanum(IV), dimetilstanum(IV) dan

stanum(IV) telah disintesis daripada bes Schiff yang terdiri daripada tiga siri ditiokarbazat

(S-2-metilbenzilditiokarbazat (S1), S-4-metilbenzilditiokarbazat (S2), S-

benzilditiokarbazat (S3) dan dua siri tiosemikarbazid (4-metil-3-thiosemikarbazid dan 4-

fenil-3-tiosemikarbazid) dengan aldehid, 2-hidroksil-3-metoksibenzaldehid (oVa) atau

2,3-dihidroksibenzaldehid (katekol). Sebatian stanum(IV) yang dihasilkan telah ditemui

mempunyai formula umum [R2Sn(ONS)] dan [Sn(ONS)2] (di mana R = Me dan Ph).

Sebatian tersebut telah dicirikan sepenuhnya dengan kaedah fiziko-kimia dan

spektroskopi. Hasil spektroskopi menyokong geometri pengkoordinatan di mana bes

Schiff bertindak sebagai ligan penderma tridentat ONS berkoordinat melalui atom-atom

nitrogen azometina, sulfur tiolo, dan oksigen fenoksid. Sebanyak 11 struktur hablur

iv

sebatian yang dijangkakan telah diselesaikan dalam kajian ini. Bagi mengesahkan data

eksperimen, sebatian-sebatian itu telah dioptimum menggunakan kaedah teori berfungsi

ketumpatan (DFT) dengan fungsian korelasi pertukaran hibrid B3LYP dengan keupayaan

pseudo LanL2DZ ke atas stanum dan set asas Pople 6-311G(d,p) bagi semua atom-atom

yang lain. Sebatian difenilstanum(IV) menunjukkan kesitotoksikan yang paling

memberangsangkan dengan nilai IC50 diantara 0.016 – 4.40 μM terhadap satu panel

daripada dua belas siri sel kanser (RT-112, EJ-28 (pundi kencing), HT29 (usus besar),

U87, SJ-G2, SMA (glioblastoma), MCF-7 (payudara), A2780 (ovari), H460 (paru-paru),

A431 (kulit), Du145 (prostat), BE2-C (neuroblastoma) dan MIA (pankreas)). Tiga

sebatian difenilstanum(IV) daripada siri oVa berkebolehan untuk mendorong penghasilan

Spesis Oksigen Reaktif (ROS) dan bertindak sebagai pendorong sel apoptosis. Ikatan

interaksi yang baik bagi kesemua siri sebatian difenilstanum(IV) telah diperhatikan dan

disokong oleh analisis molekul docking, di mana ikatan interaksi hidrogen, elektrostatik

dan hidrofobik telah diperhatikan. Ini menunjukkan bahawa pentingnya dua kumpulan

fenil yang berkoordinat secara terus kepada ion stannum untuk meningkatkan sifat

sitotosik melalui interaksi π-π yang kuat kepada biomakromolekul. Difenilstannum(IV)

memberi harapan dalam bidang pembangunan dadah terhadap pelbagai penyakit

termasuklah kanser.

v

ACKNOWLEGEMENTS First of all, I would like thank to Allah for giving His blessing to complete my PhD

research project. I wish to express my deepest gratitude and appreciation to my principal

supervisors, Associate Professor Dr. Thahira Begum and Associate Professor Dr. Alister J.

Page as well as my co-supervisors, Professor Adam McCluskey, Dr. Mohamed Ibrahim

Mohamed Tahir, Professor Abhimanyu Veerakumarasivam and Dr. Muhammad Alif bin

Muhammad Latif, whose encouragement, guidance and support from the initial to the final

level. They spent countless hours clearing my doubts and problems with regards to my

research project. I would also like to extend my appreciation to Professor Karen Anne

Crouse for meaningful discussion, Dr. Michela Simone for her grateful NMR discussion,

Dr. Robert Burns for his kind suggestions about 119Sn NMR experiments, Dr. Jennette

Sakoff for anticancer screening and Professor Edward R. T. Tiekink for single crystal X-

ray structure analysis.

My special thanks to the wonderful computational chemistry group members, Ben,

Josh, Tilly, Kas, Krishna, Simone, Gareth, Izaac, Xinyu, Babu, Ryan, Thom, who had

guided me patiently and motivated me to get better in research. My special thanks also goes

to Inorganic Chemistry members, Nabihah, Nazhirah, Chee Keong, Lee Chin, Nabeel and

Ali for their kind assistance in helping me to complete my research. My sincere

appreciation goes to all lecturers and staff at Department of Chemistry, Faculty of Science,

Medical Genetic Laboratory, Faculty of Medicine and School of Chemistry, Faculty of

Environmental and Life Sciences and the Calvary Mater Hospital, Newcastle, Australia for

being helpful and cooperative throughout this research.

And finally, I would like to thank family members, especially my husband for his

endless love and encouragement, my parents, thank you for always loving, supporting and

wishing me the best for the whole of my life. To my siblings, thank you for always making

vi

me smile and releasing my tension during writing this thesis. Lastly, I offer my best wishes

to all of those who supported me in any aspect during the completion of this research.

vii

I certify that a Thesis Examination Committee has met on 2 December 2019 to conduct the final examination of Enis Nadia binti Md Yusof on her thesis entitled “Synthesis, Structural Characterisation and Cytotoxicity Study of Tin(IV) Compounds containing ONS Schiff Bases” in accordance with the Universities and University Colleges Act 1971 and the Constitution of the Universiti Putra Malaysia [P.U.(A) 106] 15 March 1998. The Committee recommends that the student be awarded the Doctoral of Philosophy.

Members of the Thesis Examination Committee were as follows: Lim Hong Ngee, PhD Associate Professor Faculty of Science Universiti Putra Malaysia (Chairman) Tan Kar Ban, PhD Associate Professor Faculty of Science Universiti Putra Malaysia (Internal Examiner) Christopher Scarlett, PhD Professor School of Environmental and Life Sciences The University of Newcastle Australia (Internal Examiner) Jagadese J. Vittal, PhD Professor Department of Chemistry National University of Singapore Singapore (External Examiner)

__________________________________ ZURIATI AHMAD ZUKARNAIN, PhD Professor and Deputy Dean School of Graduate Studies Universiti Putra Malaysia Date: 2 January 2020

viii

This thesis was submitted to the Senate of Universiti Putra Malaysia and has been accepted as fulfilment of the requirement for the degree of Doctor of Philosophy. The members of the Supervisory Committee were as follows:

Thahira Begum, PhD Associate Professor Faculty of Science Universiti Putra Malaysia (Chairman) Alister J. Page, PhD Associate Professor School of Environmental and Life Sciences The University of Newcastle, Australia (Member) Mohamed Ibrahim Mohamed Tahir, PhD Senior Lecturer Faculty of Science Universiti Putra Malaysia, Malaysia (Member) Abhimanyu Veerakumarasivam, PhD Professor Department of Biological Sciences School of Science and Technology Sunway University, Malaysia (Member) Muhammad Alif bin Muhammad Latif, PhD Senior Lecturer Centre of Foundation Studies for Agricultural Sciences Universiti Putra Malaysia, Malaysia (Member) Adam McCluskey, PhD Professor School of Environmental and Life Sciences The University of Newcastle, Australia (Member)

______________________________ ZALILAH MOHD SHARIFF, PhD Professor and Deputy Dean School of Graduate Studies Universiti Putra Malaysia Date:

ix

Declaration by graduate student

I hereby confirm that: this thesis is my original work; the thesis contains no material which has been accepted, or is being examined, for

the award of any other degree or diploma in any university or other tertiaryinstitution;

quotations, illustrations and citations have been duly acknowledged and appropriatecopyright permissions have been obtained for the content to be made availableelectronically;

ownership of intellectual property from the thesis is as stipulated in the Memorandumof Agreement (MoA), or as according to the Universiti Putra Malaysia (Research)Rules 2012, in the event where the MoA is absent;

permission from supervisor and the office of Deputy Vice-Chancellor (Research andInnovation) are required prior to publishing it (in the form of written, printed or inelectronic form) including books, journals, modules, proceedings, popular writings,seminar papers, manuscripts, posters, reports, lecture notes, learning modules or anyother materials as stated in the Universiti Putra Malaysia (Research) Rules 2012;

there is no plagiarism or data falsification/fabrication in the thesis, and scholarlyintegrity is upheld as according to the Universiti Putra Malaysia (Graduate Studies)Rules 2003 (Revision 2012-2013) and the Universiti Putra Malaysia (Research)Rules 2012 and the University of Newcastle Rules Governing Higher Degrees byResearch. The thesis has undergone plagiarism detection software;

i give consent to the final version of my thesis being made available worldwide whendeposited in the University’s Digital Repository, subject to the provisions of theCopyright Act 1968 and any approved embargo;

the thesis has been submitted to Universiti Putra Malaysia and the University ofNewcastle as part of a Jointly Awarded Doctoral Degree.

Signature: ___________________________ Date: 9/7/2019__________________

Name and Matric No.: Enis Nadia binti Md Yusof (GS44510 (UPM) and 3285394 (UON))

x

Declaration by Members of Supervisory Committee

This is to confirm that: the research conducted and the writing of this thesis was under our supervision; supervision responsibilities as stated in the Universiti Putra Malaysia (Graduate

Studies) Rules 2003 (Revision 2012-2013) and the University of Newcastle RulesGoverning Higher Degrees by Research and University of Newcastle Code ofPractice are adhered to.

Signature: Name of Chairman of Supervisory Committee: Thahira Begum

Signature: Name of Member of Supervisory Committee: Alister J. Page

Signature: Name of Member of Supervisory Committee: Mohamed Ibrahim Mohamed Tahir

Signature: Name of Member of Supervisory Committee: Abhi Veerakumarasivam

Signature: Name of Member of Supervisory Committee: Muhammad Alif bin Muhammad Latif

Signature: Name of Member of Supervisory Committee: Adam McCluskey

xi

TABLE OF CONTENTS Page ABSTRACT i ABSTRAK iii ACKNOWLEDGEMENTS v APPROVAL vii DECLARATION ix LIST OF TABLES xvi LIST OF FIGURES xviii LIST OF SCHEMES xxiii LIST OF ABBREVIATIONS xxiv CHAPTER

1 COORDINATION CHEMISTRY AND CYTOTOXICITY OF TIN(IV) COMPOUNDS

1

1.1 Introduction 1 1.2 Long-Standing Chemotherapeutic Agents 4 1.3 Non-Platinum-Based Preclinical Investigations for

Anticancer Drug Applications 7

1.3.1 Titanium Compound 8 1.3.2 Ruthenium Compounds 9 1.3.3 Gallium Compounds 10 1.3.4 Iron Compounds 12 1.3.5 Cobalt Compounds 13 1.3.6 Gold Compounds 14 1.4 Tin(IV) compounds as potential anticancer agents 15 1.4.1 Tin(IV) Compounds of Oxygen and/or Nitrogen

Donor Ligands 17

1.4.2 Tin(IV) Compounds of Sulphur Donor Ligands 23 1.5 Chemistry and Bioactivities of Schiff Bases and Their

Tin(IV) Compounds 27

1.5.1 Anticancer Activity of Dithiocarbazate Schiff Bases and Their Tin(IV) Compounds

28

1.5.2 Anticancer Activity of Thiosemicarbazone Schiff Bases and Their Tin(IV) Compounds

31

1.6 Mechanism of Action 36 1.6.1 DNA Binding 36 1.6.2 Apoptosis 38 1.7 Knowledge Gap 40 1.8 Project Aims 41 1.9 Thesis Outline 42 2 SYNTHESIS, STRUCTURAL EVALUATION AND

CYTOTOXICITY STUDIES OF O-VANILLIN DERIVED DITHIOCARBAZATE SCHIFF BASES

43

2.1 Introduction 43 2.2 Experimental and Computational Methods 44 2.2.1 Materials 44

xii

2.2.2 Synthesis 45 2.2.2.1 S-N-R-Benzyldithiocarbazates (N = 2,

3, R = Methyl) 45

2.2.2.2 S-2-Methybenzyl-β-N-(2-hydroxy-3-methoxybenzyl methylene) dithiocarbazate (1)

46

2.2.2.3 S-4-Methybenzyl-β-N-(2-hydroxy-3-methoxybenzyl methylene) dithiocarbazate (2)

47

2.2.2.4 S-Benzyl-β-N-(2-hydroxy-3-methoxybenzylmethylene) dithiocarbazate (3)

47

2.2.3 Physical Measurements 48 2.2.4 Single Crystal X-ray Structure Determination 48 2.2.5 Density Functional Theory (DFT) Calculations 49 2.2.6 MTT Assays 50 2.3 Results and Discussion 52 2.3.1 Synthesis 52 2.3.2 IR Spectral Analysis 53 2.3.3 NMR Spectroscopic Analysis 54 2.3.4 Mass Spectral Analysis 55 2.3.5 UV-vis Absorption Spectroscopy 55 2.3.6 X-ray Crystallography 56 2.3.7 In vitro Cytotoxicity 58 2.4 Conclusions 61 3 ORGANOTIN(IV) COMPOUNDS OF O-VANILLIN

SCHIFF BASES: SYNTHESIS, STRUCTURAL CHARACTERISATION, IN-SILICO STUDIES AND CYTOTOXICITY

62

3.1 Introduction 62 3.2 Experimental 64 3.2.1 Materials 64 3.2.2 Synthesis 65 3.2.2.1 Diphenyltin(IV) Compounds 65 3.2.2.2 Dimethyltin(IV) Compounds 66 3.2.3 Physical Measurements 68 3.2.4 Single Crystal X-ray Structure Determination 68 3.2.5 Computational Methods 69 3.2.5.1 DFT Calculations 69 3.2.5.2 Molecular Docking Studies 70 3.2.6 Biological Assays 71 3.2.6.1 MTT Assays 71 3.2.6.2 Quantification of Apoptosis by

Annexin V 71

3.2.6.3 Measurement of Reactive Oxygen Species (ROS)

72

3.2.6.4 DNA Binding Studies 73 3.3 Results and Discussion 73 3.3.1 Synthesis 73

xiii

3.3.2 IR Spectral Analysis 74 3.3.3 NMR Spectroscopic Analysis 75 3.3.4 X-ray Crystallography 76 3.3.5 UV-vis Absorption Spectroscopy 85 3.3.6 Biological Assays 86 3.3.6.1 Cytotoxicity 86 3.3.6.2 Annexin V Assays of 4, 5 and 6-

Treated RT-112 Cells 88

3.3.6.3 Measurement of Reactive Oxygen Species (ROS) in RT-112 Cells Treated with 4 and 5

89

3.3.6.4 DNA Interaction Studies 92 3.3.7 Molecular Docking Studies 93 3.4 Conclusion 98 4 SELECTIVE CYTOTOXICITY OF ORGANOTIN(IV)

COMPOUNDS CONTAINING CATECHOL DITHIOCARBAZATE SCHIFF BASES

101

4.1 Introduction 101 4.2 Experimental 102 4.2.1 Materials and Physical Measurements 102 4.2.2 Synthesis 102 4.2.2.1 S-N-R-benzyldithiocarbazate

(N = 2, 3; R = methyl) 102

4.2.2.2 Schiff bases 102 4.2.2.3 Diphenyltin(IV) Compounds 104 4.2.2.4 Dimethyltin(IV) Compounds 106 4.2.3 X-ray Crystallographic Analysis 107 4.2.4 DFT Calculations and Molecular Docking

Studies 108

4.2.5 MTT Assay 108 4.3 Results and Discussion 108 4.3.1 Synthesis 108 4.3.2 IR Spectral Analysis 111 4.3.3 NMR Spectroscopic Analysis 112 4.3.4 Mass Spectral Analysis 114 4.3.5 UV-vis Absorption Spectroscopy 114 4.3.6 Structure Descriptions for 12, 14, 15, 17 and 18 115 4.3.7 In vitro Cytotoxicity 125 4.3.8 DNA Binding Analysis 126 4.3.9 Molecular Docking Analysis 127 4.4 Conclusions 134 5 HOMOLEPTIC TIN(IV) COMPOUNDS OF

DINEGATIVE ONS DITHIOCARBAZATE SCHIFF BASES: SYNTHESIS, X-RAY CRYSTALLOGRAPHY, DFT AND CYTOTOXICITY STUDIES

136

5.1 Introduction 136 5.2 Experimental 137 5.2.1 Materials and Physical Measurements 137

xiv

5.2.2 Synthesis 138 5.2.2.1 Synthesis of Schiff Bases 138 5.2.2.2 Synthesis of Tin(IV) Compounds (19-

24) 138

5.2.3 Single Crystal X-ray Structure Determination 141 5.2.4 DFT Calculations and Molecular Docking

Studies 141

5.2.5 MTT Assays 142 5.3 Results and Discussion 142 5.3.1 Synthesis 142 5.3.2 IR Spectral Analysis 143 5.3.3 NMR Spectroscopic Analysis 144 5.3.4 UV-vis Absorption Spectral Analysis 145 5.3.5 X-ray Crystallography 145 5.3.6 In vitro Cytotoxic Activity 151 5.3.7 Molecular Docking Analysis 151 5.4 Conclusions 156 6 TIN(IV) COMPOUNDS OF THIOSEMICARBAZONE

SCHIFF BASES: SYNTHESIS, STRUCTURAL CHARACTERISATION AND IN VITRO CYTOTOXICITY

158

6.1 Introduction 158 6.2 Experimental 159 6.2.1 Materials and Physical Measurements 159 6.2.2 Synthesis 160 6.2.2.1 Schiff Bases 160 6.2.2.2 Tin(IV) Compounds derived from 25

and 27 162

6.2.2.3 Tin(IV) Compounds derived from 26 165 6.2.2.4 Tin(IV) Compounds derived from 28 166 6.2.3 DFT Calculations and Molecular Docking

Simulations 168

6.2.4 MTT Assay 168 6.3 Results and Discussion 169 6.3.1 Synthesis 169 6.3.2 IR Spectral Analysis 170 6.3.3 NMR Spectroscopic Analysis 171 6.3.4 Mass Spectral Analysis 173 6.3.5 UV-vis Absorption Spectral Analysis 173 6.3.6 In vitro Cytotoxic Activity 174 6.3.7 DNA Binding Studies 176 6.3.8 Molecular Docking Analysis 177 6.4 Conclusions 182 7 CONCLUSIONS AND FUTURE RECOMMENDATIONS 184 7.1 Conclusions 184 7.2 Future Recommendations 187

xv

REFERENCES 189 APPENDICES 215 BIODATA OF STUDENT 344 LIST OF PUBLICATIONS 345

xvi

LIST OF TABLES Table

Page

1.1 Cytotoxic activity of tin(IV) compounds against a panel cancer cells and survival of MRC-5 cells, compared to cisplatin.

20

2.1 Physical data of synthesised S-substituted dithiocarbazate derivatives.

46

2.2 Crystal data and refinement details for Schiff base 3

50

2.3 Hydrogen-bond geometry (Å, °) of 3 obtained from single crystal X-ray diffraction analysis. The structure is shown in Figure 2.2.

57

2.4 Summary of the in vitro cytotoxicity of compounds 1-3 in several cell lines, determined by MTT assay and expressed as GI50 values with standard errors. (GI50 is the concentration at which cell growth is inhibited by 50% over 72 h).

60

3.1 Crystal data and refinement details for compounds 7, 8 and 9.

70

3.2 Selected geometric parameters (Å, °) for 7, 8 and 9.

82

3.3 Summary of the in vitro cytotoxicity of the Schiff bases and their organotin(IV) compounds.

90

3.4 Binding constants (Kb), hypochromism (%), bathochromic shifts (nm) and Gibbs free energy (kJ mol−1) values for the interaction of organotin(IV) compounds with calf thymus DNA (CT-DNA).

94

3.5 Molecular docking data for the organotin(IV) compounds with B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2.

96

4.1 Crystallographic data and refinement details for 12, 14, 15, 17 and 18.

110

4.2 Selected geometric data (Å, °) characterising 12, 14, 15, 17 and 18.

120

4.3 Summary of the in vitro cytotoxicity of the Schiff bases (10-12) and their organotin(IV) compounds (13-18) in several cell lines, determined by MTT assay, expressed as GI50 values with standard errors. GI50 is the concentration at which cell growth was inhibited by 50% over 72 h.

129

4.4 Binding constants (Kb), hypochromism (%), bathochromic shifts (nm) and Gibbs free energy (kJ mol-1) values for the interaction of organotin(IV) compounds with CT-DNA.

131

xvii

4.5 Molecular docking data for the organotin(IV) compounds with B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2.

133

5.1 Crystal data and refinement details for complexes 20 and 21.

142

5.2 Selected geometric parameters (Å, °) for 20 and 21.

147

5.3 Summary of the in vitro cytotoxicity of tin(IV) compounds in several cell lines, determined by the MTT assay and expressed as a GI50 value with standard error. GI50 is the concentration of tin(IV) compounds at which cell growth is inhibited at 50% over 72 hours.

153

5.4 Molecular docking data for the tin(IV) compounds with B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2.

154

6.1 In vitro cytotoxicity of tin(IV) compounds (29-30, 32-40) derived thiosemicarbazone Schiff bases (25-28) in several cell lines, determined by the MTT assay and expressed as a GI50 value with standard error. GI50 is the concentration at which cell growth is inhibited by 50% over 72 hours.

178

6.2 Molecular docking data for 27, 32, 33 and 34 with B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2.

181

xviii

LIST OF FIGURES

Figure

Page

1.1 Estimated number of new cancer cases and mortality in 2018, worldwide, all cancers, both sexes and all ages.

2

1.2 Structure of different platinum-based anticancer drugs.

5

1.3 Schematic diagram showing the cytotoxic pathway for cisplatin. After entering the tumour cells, cisplatin is equated, then binds to cellular DNA.

6

1.4 The new cancer drug allows for low-dose cancer chemotherapy and promises fewer side effects.

7

1.5 The titanium-based compounds, budotitane (left) and titanocene dichloride (right).

9

1.6 The ruthenium-based compounds, NAMI-A (left) and KP1019 (right).

10

1.7 Structure of gallium-based compound, KP46 which was designed and used in clinical trials.

11

1.8 Schematic representation of the mechanism of action of gallium-based compounds. Tf = transferrin; NDP = nucleoside diphosphate; dNDP = desoxynucleoside diphosphate; BAX = a proapoptotic protein.

12

1.9 The iron-based compounds, ferrocenium picrate (left) and ferrocenium trichloroacetate (right).

13

1.10 The cobalt-based compounds, hexacarbonyldicobalt complex of the propargylic ester of acetylsalicylic acid (Co-ASS).

14

1.11 The gold-based compounds, auranofin (left) and [Au(dppe)2]Cl (right).

15

1.12 The structure of tin(IV) compounds containing oxygen and/or nitrogen donor ligands.

21

1.13 The structures of tin(IV) compounds containing sulphur donor ligands.

25

1.14 Tautomerism in (a) DTC and (b) TSC Schiff bases.

28

1.15 Different substituents at position R in dithiocarbazate derivatives.

30

xix

1.16 The structure of tin(IV) compounds derived from dithiocarbazate Schiff bases.

31

1.17 Chemical structure of (a) Methisazone and (b) Triapine.

32

1.18 The structure of tin(IV) compounds derived from thiosemicarbazone Schiff bases.

35

1.19 Molecular docked model of diphenyltin(IV) complex with DNA dodecamer duplex of sequence d(CGCGAATTCGCG)2 (PDB ID: 1BNA). The image provides the side view of the docked model of complexes.

37

1.20 a) Molecular docking simulation of diethyltin(IV) complex-DNA complex (binding site of topoisomerase II) b) detailed molecular interactions of diethyltin(IV) complex with amino acid residues.

38

1.21 Mechanism of extrinsic and intrinsic apoptotic pathway.

40

2.1 The structures of (a) vanillin and (b) o-vanillin.

43

2.2 The (a) thione and (b) thiol tautomeric forms of the Schiff bases 1-3.

54

2.3 HOMO-LUMO of (a) 1, (b) 2 and (c) 3.

56

2.4 The molecular structure of 3, showing the atom- labelling scheme and displacement ellipsoids at the 50% probability level.

57

2.5 Molecular packing in 3: (a) a perspective view of the supramolecular layer sustained by thioamide-N-H…S(thione) and phenyl-C-H…O(hydroxy) interactions and, (b) a view of the unit-cell contents shown in projection down the b axis, highlighting one layer in space-filling mode. The N-H…S and C-H…O interactions are shown as blue and orange dashed lines, respectively. For (a), non-interacting H atoms have been omitted.

58

3.1 The structures of Schiff bases (1-3) and organotin(IV) compounds (4-9).

64

3.2 Molecular structures of the first independent molecules of (a) 7, (b) 8 and (c) 9 showing atom labelling schemes. (d) Overlay diagram.

79

3.3 Crystallographic diagrams for 7: (a) Molecular structure of the second independent molecule, molecule b, (b) overlay diagram of molecules a (red image) and inverted-b (green) drawn so the SnC2 atoms of the Me2Sn moiety are overlapped and (c) a view of the supramolecular layer in the ab-plane (left image) and a view of the unit cell contents in projection down the a-axis with one layer

80

xx

highlighted in space-filling mode (right image). The C‒H…O and C‒H…π interactions are shown as orange and purple dashed lines, respectively.

3.4 Crystallographic diagrams for 8: (a) Molecular structure of the second independent molecule, molecule b, (b) overlay diagram of molecules a (blue image) and inverted-b (pink) drawn so the SnC2 atoms of the Me2Sn moiety are overlapped and (c) a view of the supramolecular dimer sustained by C‒H…π (chelate) interactions (left image; non-participating hydrogen atoms have been omitted) and a view of the unit cell contents in projection down the a-axis. The C‒H…O and C‒H…π interactions are shown as orange and purple dashed lines, respectively.

80

3.5 Crystallographic diagrams for 9: (a) Molecular structure of the second independent molecule, molecule b, (b) overlay diagram of molecules a (yellow image) and b (aqua) drawn so the five-membered rings are overlapped and (c) supramolecular dimer sustained by edge-to-edge chelate ring…benzene interactions (upper image) between Sn1-containing molecules, supramolecular layer sustained by edge-to-edge π (chelate ring)…π (ethoxybenzene) and phenyl-C‒H…π (methoxybenzene) interactions occurring between Sn2-containing molecules and a view of the unit cell contents in projection down the b-axis. The edge-to-edge π (chelate ring)…π (ethoxybenzene) and C‒H…π interactions are shown as blue and purple dashed lines, respectively.

84

3.6 HOMO-LUMO of (a) 4, (b) 5, (c) 6, (d) 7, (e) 8 and (f) 9.

86

3.7 Apoptosis detection through fluorescence microscopy. Cells were treated for 24 h with 4 (0.31 µM), 5 (1.66 µM) and 6 (0.58 µM), and the negative control (DMSO) in complete media. After staining with Annexin V and PI, necrotic and apoptotic cells were detected by fluorescence microscopy (20×).

91

3.8 Percent reactive oxygen species (ROS) production in RT-112 cells treated with (a) 4 (0.31 μM) (b) 5 (1.66 μM) for 24 h and stained with 1 mM DCFH-DA for 60 minutes at 37 °C. DMSO and H2O2 acted as negative and positive controls, respectively.

92

3.9 (a) Electronic absorption spectra of (i) 4, (ii) 5 and (iii) 6; (b) Plot of [DNA]/εa − εf vs [DNA] for absorption titration of DNA with (i) 4, (ii) 5 and (iii) 6. The arrow indicates the change in absorbance in tandem with increasing DNA concentration.

95

3.10 (a) Schematic representation of organotin(IV) compounds that fit well in the grooves of the DNA strand obtained by docking simulations.

97

xxi

4.1 The structures of catechol dithiocarbazate Schiff bases (10-12) and

their organotin(IV) compounds (13-18).

102

4.2 (a) Thione and (b) thiol tautomerism of Schiff bases 10-12.

109

4.3 Frontier molecular orbitals LUMO and HOMO of the optimised (a) 10, (b) 11, (c) 12, (d) 13, (e) 14, (f) 15, (g) 16, (h) 17 and (i) 18 using B3LYP functional.

116

4.4 Molecular structure of 12 and atom labelling scheme.

117

4.5 ORTEP structures showing atom labelling scheme and overlay diagrams for (a) the two independent molecules of 14, (b) the two independent molecules of 15, (c) 17, (d) 18 and (e) overlay diagram for 17 and 18. For the overlay diagrams, molecules have been overlapped so that the CS2 residues are coincident.

119

4.6 A view of the linear supramolecular chain along [1 0 4] in the crystal of 12. The hydroxyl-O‒H…O(hydroxyl) and thioamide-N‒H…S(thione) hydrogen bonds are shown as orange and green dashed lines, respectively.

123

4.7 A view of the zig-zag chain along [0 0 1] in the crystal of 14. The hydroxyl-O‒H…O(hydroxyl) and thioamide-N‒H…S(thione) hydrogen bonds are shown as orange and green dashed lines, respectively. For reasons of clarity, the hydrogen atoms have been removed and only the ipso-carbon atoms of the tin-bound phenyl rings shown.

125

4.8 (a) Electronic absorption spectra of (i) 16, (ii) 17 and (iii) 18; (b) Plot of [DNA]/εa - εf vs [DNA] for absorption titration of DNA with (i) 16, (ii) 17 and (iii) 18. (The arrow indicates the change in absorbance in tandem with increasing DNA concentration).

130

4.9 (a) Schematic representation of organotin(IV) compounds that fit well in the grooves of the DNA strand obtained by docking simulations. The two double-stranded DNA comprised of the phosphate deoxyribose backbone with guanine (DG, red), cytosine (DC, blue), adenine (DA, pink) and thymine (DT, orange). (b) Molecular interactions of organotin(IV) compounds within the grooves of double stranded DNA residues.

132

5.1 The structures of homoleptic tin(IV) compounds (19-24).

137

5.2 Frontier MOs of (a) 19, (b) 20, (c) 21, (d) 22, (e) 23 and (f) 24.

146

5.3 Molecular structures of the molecules in (a) 20 (first independent molecule; the structure for the second independent molecule is

148

xxii

shown in Appendix Figure A5.1) and (b) 21, showing atom labelling schemes and 50% displacement ellipsoids. (c) Overlay diagram of the molecules in 20 (red image for the first independent molecule), 20a (green, inverted second molecule) and 21 (blue). Molecules have been overlapped so the Sn,S1,N2 chelate rings are coincident.

5.4 Homoleptic tin(IV) compounds that have been studied.

150

5.5 (a) Schematic representation of 19 (cyan), 21 (yellow), 22 (blue), 23 (grey) and 24 (green) that fit well in the grooves of the DNA strand obtained by docking simulations. The two double-stranded DNA comprised of the phosphate deoxyribose backbone (grey) with guanine (green), cytosine (purple), adenine (DA, red) and thymine (cyan). (b) Molecular interactions of 19, 21, 22, 23 and 24 within the grooves of double stranded DNA residues. Hydrogen bond, electrostatic and hydrophobic interactions are depicted by green, orange and pink dot lines, respectively.

155

5.6 The small modification (blue) of backbone of (a) dithiocarbazate Schiff bases to produce (b) thiosemicarbazone Schiff bases.

157

6.1 The structures of (a) dithiocarbazate and (b) thiosemicarbazone Schiff bases.

159

6.2 Frontier MOs (a) 25, (b) 26, (c) 27, (d) 28, (e) 29, (f) 30, (g) 31, (h) 32, (i) 33, (j) 34, (k) 35, (l) 36, (m) 37, (n) 38, (o) 39 and (p) 40.

174

6.3 Electronic absorption spectra of (i) 27, (ii) 32, (iii) 33 and (iv) 34. The arrow indicates the change in absorbance in tandem with increasing DNA concentration.

180

6.4 Molecular interactions of (a) 27, (b) 32, (c) 33 and (d) 34 within the grooves of double stranded DNA residues. Hydrogen bond, electrostatic and hydrophobic interactions are depicted by green, orange and pink dot lines, respectively.

180

xxiii

LIST OF SCHEMES

Scheme

Page

2.1 Synthetic pathway for the formation of 1-3.

53

3.1 Synthesis of organotin(IV) compounds (4-9).

74

4.1 Synthetic pathway for the formation of Schiff bases 10, 11 and 12.

109

4.2 Synthetic pathway for the synthesis of organotin compounds 13-18.

111

5.1 Synthetic pathway for the synthesis of 19-24.

143

6.1 Synthetic pathway of thiosemicarbazone Schiff bases 25-28.

169

6.2 Synthetic pathway of tin(IV) compounds (29-40) of thiosemicarbazone Schiff bases.

170

xxiv

LIST OF ABBREVIATIONS

3677 Human melanoma cell line A2780 Human ovarian carcinoma cancer cell line A431 Human skin cancer cell line A549 Human lung carcinoma line B16F10 Murine melanoma cell line B3LYP Berke's three-parameter exchange potential and the

Lee–Yang–Parr correlation functional Bcap37 Mammary tumour cell line BE2-C Neuroblastoma cancer cell line Bel-7402 Hepatocellular carcinoma line BGC-823 Gastric carcinoma cell line BSA Bovine serum albumins CDCl3 Deuterated chloroform CH3OH Methanol CHO Chinese hamster ovary cell line CIF Crystallographic Information File CML-T1 Chronic myeloid leukemic cell line CoLo205 Colon carcinoma cell line CT-DNA Calf thymus deoxyribonucleic acid DCF 2′,7′-Dichlorodihydrofluorescein DCF-DA Dichloro-dihydro-fluorescein diacetate DCM Dichloromethane DFT Density Functional Theory DMEM Dulbecco’s modified Eagle’s medium DMSO-d6 Deuterated dimethylfulfoxide DNA Deoxyribonucleic acid DTC Dithiocarbazate Du145 Human prostate carcinoma cancer cell line EJ-28 Muscle invasive bladder cancer cell line Et3N Triethylamine FT-IR Fourier transform infrared spectroscopy GI50 Concentration of the anti-cancer drug that inhibits the growth of

cancer cells by 50% relative to untreated control H2981 Human lung adenocarcinoma cell line H460 Human non-small lung carcinoma cell line HCT116 Human colorectal carcinoma cell line HCT-8 Colon carcinoma cell line HeLa Human cervical epithelioid carcinoma cell line HepG2 Human liver carcinoma cell line HL-60 Human acute myeloid leukemia cell line HT29 Human colon adenocarcinoma cell line IC50 Concentration of compound that inhibits a biological activity by

50% K562 Human myelogenous leukemia cell line KB Nasopharyngeal carcinoma cell line KP1019 Indazolium trans-[tetrachlorobis(1H-indazole)ruthenate(III)] KP46 Tris(8-quinolinolato)gallium(III)

xxv

LanL2DZ Los Alamos National Laboratory 2-double-z LMCT Ligand-metal charge transfer LMS Leiomysarcoma cell line LOX Lipoxygenase m/z Mass-to-charge ratio MCF-10A Normal breast cell line MCF-7 Human breast carcinoma with positive oestrogen receptor cell line MDA-MB-231 Human breast carcinoma with negative oestrogen receptor cell line Me2SnCl2 Dimethyltin(IV) dichloride MeOH Methanol MIA Pancreatic cancer cell line MKT4 Titanocene dichloride MRC-5 Normal foetal lung fibroblast cell line MTT 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide NAMI-A Imidazolium trans-[tetrachloro(dimethylsulfoxide)(1H-

imidazole)ruthenate(III)] NCI National Cancer Institute NMR Nuclear magnetic resonance ONS Oxygen-nitrogen-sulphur oVa o-Vanillin; 2-hydroxy-3-methoxybenzaldehyde P388 Lymphocytic leukaemia cell line PDB Protein Data Bank Ph Phenyl Ph2SnCl2 Diphenyltin(IV) dichloride PI Propidium Iodide r.m.s Root-mean-square ROS Reactive Oxygen Species RR Ribonucleotide reductase RT-112 Minimum muscle invasive bladder cancer cell line S1 S-2-Methylbenzyldithiocarbazate S2 S-4-Methylbenzyldithiocarbazate S3 S-Benzyldithiocarbazate SJ-G2 Pediatric glioblastoma cell line SMMC-7721 Human hepatocellular carcinoma cell line TiO2 Titanium oxide Tris-HCl Tris(hydroxymethyl)aminomethane hydrochloride U2-OS Human osteosarcoma cell line U87 Human glioblastoma cell line UV-vis Ultraviolet-visible WiDr Human colon carcinoma cell line WRL-68 Normal liver cell ORTEP Oak Ridge thermal ellipsoid plot PBS Phosphate buffer saline PEG Polyethylene glycol

1

CHAPTER 1

COORDINATION CHEMISTRY AND CYTOTOXICITY OF TIN(IV) COMPOUNDS

1.1 Introduction

Cancer is the proliferation of abnormal cells which can occur in various body parts.

Globally, the International Agency for Research on Cancer has reported 18.1 million new

cancer cases (Figure 1.1) and 9.6 million deaths due to cancer in 2018.2 The types of

cancer with highest mortality rates are lung (1.76 million deaths), colorectum (880 792),

stomach (782 685), liver (781 631 deaths) and breast (626 679 deaths) cancer.2 Cancer

starts by certain changes to genes that control the cells function, especially on how the

cells grow and divide. Genes are essential and carry the instructions to make proteins in

cells. The changes or damages in genes called mutated genes, where they do not work

properly in giving instructions to the cells and ultimately the cells grow out of control,

which lead to cancer. Cancer cells are also immature and do not develop into mature cells

with their specific functions, able to avoid the immune system and capable to ignore

signals that tell them to stop dividing. Furthermore, they also can spread to other body

parts through circulatory and lymphatic systems.1 Cancer can be caused by person’s

genetic factors if the genetic changes present in germ cells, and as well as external agents

such as a high exposure rate to ultraviolet, ionising radiation and/or cigarette smoke,

particularly at an advanced age. Occupational exposure to chemical carcinogens is also a

factor, especially for those who are regularly exposed to aromatic amines such as beta-

naphthylamine, xenylamine and benzidine.3 Furthermore, infections from certain viruses,

bacteria and parasites also can initiate the development of cancer. Undoubtedly, ageing

is another factor for the development of cancer as the frequency of cancer increases with

2

age. Over time, continuous errors accumulate in our genes and the risk of developing

cancer increases.

Figure 1.1: Estimated number of new cancer cases and mortality in 2018, worldwide, all cancers, both sexes and all ages (Reproduced with permission from ref. 2). The transformation of benign cells to malignant cells involves changes in cell

signalling, cell metabolism, the cell’s ability to avoid apoptosis and its ability to spread

or metastasise to different sites in the body.1 There are many cancer treatment options

available, depending on the type and stage of cancer. Cancer treatments include surgery,

radiation therapy, chemotherapy, immunotherapy, targeted therapy, hormone therapy and

stem cell transplant, however most cancer patients receive a combination of these

treatments.4 Nevertheless, there are also many cancers where there are effectively no

viable therapies until now.

Targeted therapy is one of the most prominent cancer treatments and uses specific

molecular targets, such as antibodies as the main treatment source for hematologic

malignancies.5 Antibodies are part of the immune system and thus play a major role in

defending the body from disease and infection. An antibody is a protein that binds to a

specific protein called an antigen. As a result, antibodies automatically alert the immune

3

system to destroy cells containing antigens.6 To note here that this approach only works

if the targeted receptor is present on the cell surface and appropriately clustered. They are

also problematic as this approach shuts down a wide array of other signalling pathways.

Antibody-dependent cellular cytotoxicity is an effective treatment, under the right

conditions, but it is also a treatment with high numbers of level 5 adverse events which

are Type I immediate reactions (anaphylaxis, urticaria); Type II reactions (immune

thrombocytopenia, neutopenia, hemolytic anemia); Type III responses (vasculitis, serum

sickness; some pulmonary adverse events); and Type IV delayed mucocutaneous

reactions as well as infusion reactions/cytokine release syndrome, tumor lysis syndrome,

progressive multifocal leukoencephalopathy and cardiac events, especially the new

programmed cell death 1 (PD1) and anti-PD1 mAbs (nivolumab and pembrolizumab)

based approaches.7,8 Researchers have been successful in designing and developing many

copies of antibodies that specifically target cancer cell antigens, known as monoclonal

antibodies (mAbs).9 There are different types of mAbs used in cancer treatment, for

instance naked mAbs, bispecific mAbs and conjugated mAbs. Naked mAbs are

antibodies that work by themselves without the presence of drugs or radioactive materials

and are commonly used to treat cancer. However, antibodies can also bind to the antigen

of non-cancerous cells, as well as free-floating proteins. Alemtuzumab (Campath®) was

the first humanised mAb therapeutic, depleting lymphocytes, monocytes and dendritic

cells, where this antibody is very commonly used to treat chronic lymphocytic leukemia

patients.10 Bispecific mAbs work with two different proteins at the same time. For

example, blinatumomab (Blincyto) is used to treat some types of acute lymphocytic

leukemia, where the mAbs attached to the CD19 protein found on some leukemic and

lymphoma cells as well as to the CD3 protein found on immune cells (T cells).10

Conjugated mAbs consist of antibodies joined together with chemotherapy drugs or

4

radioactive particles. Even though new technologies have been developed to treat cancer,

these approaches do not guarantee a cure or remission, thus, chemotherapy remains of

much interest due to the effectiveness of the drugs that can work throughout the body.

1.2 Long-Standing Chemotherapeutic Agents

Chemotherapy is one of the most powerful tools used in cancer treatment. Ongoing

research continues to find new and effective low cost chemotherapy drugs with fewer side

effects as compared to existing ones. Chemotherapy can kill cancer cells that have spread

or metastasised to other parts of the body from their original primary tumour site.

Inorganic compounds offer potential advantages for chemotherapy over organic

compounds. The great bioactivity of inorganic compounds can be explained on the basis

of Tweedy’s chelation theory, where chelation reduces the polarity of the metal ion due

to partial sharing of its positive charge with donor atoms and also due to electron

delocalisation over the whole chelate rings. Such chelation increases the lipophilic

character of the inorganic compounds which favors the permeation of the compounds

through the lipid layer of the cell membrane.11–13 Importantly, inorganic drugs display a

wide range of coordination numbers, various geometries, oxidation states and

thermodynamic and kinetic characteristics of ligand substitution that offer the medicinal

chemist opportunities to explore metal-containing compounds with different strategies.14

An effective and long-standing metallo-chemotherapy agent is cisplatin (Figure 1.2(a))

that falls into the class of deoxyribonucleic acid (DNA)-damaging agents. Cisplatin is

clinically proven to combat many cancers including testicular, ovarian, cervical, head and

neck, bladder, lung and colorectal cancers.15,16 Cisplatin becomes activated once it enters

the cell, where in the cytoplasm, the two Cl atoms are displaced by water molecules

through hydrolysis. The hydrolysed product formed behaves as an electrophile that can

5

react with any nucleophile in the cell such as sulfhydryl/thiol (R-SH) groups in proteins

and nitrogen atoms in nitrogenous bases of DNA.17–19 Cisplatin binds to the N7 reactive

centre on purine of the guanine and adenine bases and drives DNA damage by blocking

cell division and ultimately inducing programmed cell death (apoptosis). The major

changes in DNA are due to the 1,2-intrastrand cross-links of purine bases with cisplatin,

which include bifunctional adducts that crosslink 90% of 1,2-intrastrand of guanine-

guanine residues in the sequences d(GpG) adducts and 10% of 1,2-intrastrand of adenine-

guanine residues in the sequences d(ApG) (Figure 1.3). The toxicity of cisplatin arises

from the presence of 1,3-intrastrand (GpXpG) adducts, interstrand cross-links and

nonfunctional adducts with cisplatin.17,20

PtClH3N

H3N ClO

Pt

O

O

O

NH3

NH3

NH2

Pt

H2N

O

OO

O

cisplatin(approved by FDA)

carboplatin(approved by FDA)

oxaliplatin(approved by FDA)

PtH3N

H3NO

OO

nedaplatin(approved for use in Japan)

heptaplatin(approved for use in South Korea)

Pt

NH2

NH2O

OO

CH3

laboplatin(approved for use in China)

(a) (b) (c)

(d) (e) (f)O

Pt

O

O

O

NH2

NH2

O

O

Figure 1.2: Structure of different platinum-based anticancer drugs.

6

Figure 1.3: Schematic diagram showing the cytotoxic pathway for cisplatin. After entering the tumour cells, cisplatin is hydrolysed, then binds to cellular DNA (Reproduced with permission from ref. 21). Platinum-based compounds remain the most effective chemotherapeutic agents for

tumour therapy, such as carboplatin, oxaliplatin, nedaplatin, heptaplatin and laboplatin

(see Figure 1.2), except for all breast and colon cancers. However, such platinum drugs

are prone to side effects such as anemia, diarrhea, alopecia, petechia, fatigue

nephrotoxicity, emetogenesis, ototoxicity, neurotoxicity and also face increasing

resistance from cancer cell lines.22 Figure 1.4 highlights the effectiveness of a

combination of new drugs and chemotherapy to prevent drug resistance in treating cancer

patients. It is therefore evident that the search for novel non-platinum-based compounds

will potentially open up new avenues in the development of chemotherapeutic drugs.

Furthermore, the discovery of new drugs with enhanced activity, selectivity and

bioavaibility, as well as fewer side effects than conventional drugs to treat cancer patients

is crucial.

7

Figure 1.4: By adding new cancer drug to an existing treatment, the cancer cells became sensitive to existing chemotherapeutics and more responsive to the treatment.23

1.3 Non-Platinum-Based Compounds in Preclinical Investigations for

Anticancer Drug Applications

From the medicinal perspective, metals are relevant because they play crucial roles

in the systems of living organisms. Many are soluble in biological systems by losing

electrons to form positively charged species. Biological molecules, such as DNA and

proteins are electron-rich, and so a metal ion’s positive charge enables it to interact and

bind with them.24 Both preclinical in vitro and in vivo studies of numerous metal

compounds have been carried out to determine their anticancer properties. Despite highly

promising results obtained with metal compounds in preclinical studies, only titanium,

ruthenium and gallium compounds have been investigated in clinical phase I and phase

II studies. All the investigations have proven that the development of non-platinum-based

drugs with different mechanism of cancer cell cytotoxicity from cisplatin was possible.25–

27

8

1.3.1 Titanium Compounds

Titanium in solution is dominated by a +4 oxidation state and has a strong tendency

to hydrolyse and form highly insoluble TiO2. However, this hydrolysis reaction can be

reduced by introducing appropriate ligands coordinated to the titanium centre, and this is

the common strategy for the development of titanium-based anticancer drugs. The

octahedral titanium β-diketone compound, budotitane (Figure 1.5, left), was the first non-

platinum-metal-based compound to enter clinical trials for the treatment of a broad

spectrum of cancers. The main target of the titanium-based compound was the inhibition

of DNA synthesis, triggering apoptosis and disrupting the topoisomerase II enzyme in

cancer cells, which led to the induction of the cell death pathway. Unfortunately, cancer

cells did not respond to budotitane after infusion twice in a week with a dose of 100

mg/m2 in an 18 patient in a Phase I clinical, which caused undesirable cardiac arrhythmia

in the patients. Unlike budotitane, titanocene dichloride (MKT4) (Figure 1.5, right)

showed good clinical outcome in Phase I trials. Testing on 40 patients with solid tumours

showed minor responses in bladder carcinoma and non-small cell lung cancer after

treatment with a lyophilised formulation of MKT4. The gastrointestinal side effects could

be reduced but were still dose-limiting because of possible accumulative nephrotoxicity.28

In clinical Phase II trials, no responses were noted in the 14 patients with metastatic renal-

cell carcinoma and 15 patients with metastatic breast cancer after administration of 270

mg/m2 of the complex every three weeks.29 Researchers also found that MKT4 inhibited

the synthesis of DNA by covalently binding to the DNA via the phosphate backbone and

as well as induced apoptosis. Binding studies suggested that the Ti(IV) ions were taken

up by the transport protein transferrin (which acts as a vehicle), where only this protein

had the capability of transporting other metal ions to the cells.29–31

9

Figure 1.5: The titanium-based compounds, budotitane (left) and titanocene dichloride (right) 27 1.3.2 Ruthenium Compounds

Ruthenium compounds have been investigated for their potential antitumour

activity since the 1970s. For example the octahedral imidazolium trans-

[tetrachloro(dimethylsulfoxide)(1H-imidazole)ruthenate(III)] (NAMI-A) and indazolium

trans-[tetrachlorobis(1H-indazole)ruthenate(III)] (KP1019) complexes (Figure 1.6) both

entered clinical trials for the treatment of tumours. Both of them were significantly

different in their structures as compared to the typically square planar Pt(II)-based drugs.

KP1019 showed a wide range of growth inhibitory effects on in vitro tumour cells

including colorectal carcinomas and a variety of primary explanted human tumours.32 In

addition, NAMI-A had marked efficacy against the formation of metastases and had the

ability to control the angiogenic potential of tumours.33,34 The safe administration of

NAMI-A was a three hour infusion dose of 300 mg/m2/day for five days, over a duration

of three weeks. Only two patients showed common toxicity criteria with grade 3 and 4

hypersensitivity reaction.35 KP1019 had excellent uptake into the cells and hence the cells

underwent apoptosis by an intrinsic mitochondrial pathway. At this point, it caused

oxidative stress and DNA damage, which suggested that KP1019 could untwist and bend

DNA.36

OTi

O

OO O

O

TiCl

Cl

10

RuClCl

Ru

N

ClCl

OMe

Me

NH

RuCl

N

ClN

ClCl

NH

HN

Figure 1.6: The ruthenium-based compounds, NAMI-A (left) and KP1019 (right).27

1.3.3 Gallium Compounds

Among the p-block metals, gallium-based compounds have entered clinical trials

for the treatment of diverse disorders, categorised as (a) accelerated bone resorption, with

or without elevated plasma calcium; (b) autoimmune disease and allograft rejection; (c)

certain cancers; and (d) infectious diseases.37 Gallium nitrate, gallium chloride and

gallium maltolate were the gallium-based compounds which entered clinical trial Phase I

and II, but not all gallium species were effective for all types of malignancies. For

example, gallium nitrate was not successful in treating various malignancies such as

melanoma and breast cancer, but was effective in other types of cancer. This was because

gallium nitrate was readily hydrolysed in biological media to give non-soluble gallium

oxides which were responsible in blocking the absorption and membrane permeation of

the gallium ions thus reducing its effectivity in cancer treatments. During Phase II clinical

trials, gallium nitrate was administered as treatment for metastatic urothelial carcinomas

with a combination of two chemotherapy medication, vinblastine and ifosfamide and

returned a very good response rate (67%) on patients. However, patients suffered from

strong toxicity effects, especially granulocytopenia. Furthermore, gallium chloride was

administered in combination of paclitaxel, gemcitabine or vinorelbine at preclinical

11

stages. To improve the bioactivities of previous gallium-based compounds, researchers

developed the orally bioavailable gallium compound, [tris(8-quinolinolato)gallium(III)]

(KP46) (Figure 1.7) which were studied in phase I clinical trials. KP46 showed better

results than the combination of platinum compounds due to its lipophilic character of

ligands after oral administration.38,39 The mechanism of action of gallium-based

compounds is mainly by the binding of Ga3+ to the transferrin followed by accumulation

in endosomes. After Ga3+ was transported into the cytosol, Ga3+ ions were able to inhibit

the function of ribonucleotide reductase (RR), where Ga3+ ions played a role in catalysing

the conversion of ribonucleotides to deoxyribonucleotides. The coordination of Ga3+ with

the RR inhibitor activated the cell cycle arrest and ultimately the cells underwent

apoptosis through a mitochondrial pathway (Figure 1.8).40

N

Ga O

NO

N

O

Figure 1.7: Structure of gallium-based compound, KP46, which was designed and used in clinical trials.

12

Figure 1.8: Schematic representation of the mechanism of action of gallium-based compounds. Tf = transferrin; NDP = nucleoside diphosphate; dNDP = deoxynucleoside diphosphate; BAX = a proapoptotic protein (Reproduced with permission from ref. 41). 1.3.4 Iron Compounds

Iron is an essential element in biological systems and also an important nutrient

involved in cancer cells’ proliferation. The first antineoplastic iron complexes, which

were ferrocenium picrate and ferrocenium trichloroacetate (Figure 1.9) showed the best

cytotoxic effects with optimum cure rates of 100% against Ehrlich ascites tumour in CF1

mice.42 These two excellent ferrocene compounds underwent an oxidation process, where

ferrocene could be converted into ferrocenium ions inside the tumour cells. The

mechanism of ferrocenium salts was not via direct interaction with DNA, but instead via

the generation of reactive oxygen species (ROS) which led to protein and DNA damage,

and ultimately cell death. Moreover, the efficacy of ferrocene derivatives as antiestrogen

drugs depended on their proton-coupled electron transfer (redox potential) including

intramolecular interactions in the molecular π system.43 Other researchers claimed that

the stable iron complexes of pentadentate pyridyl ligands showed high cytotoxicity

against HeLa (human cervical epithelioid carcinoma cells) and HepG2 (human liver

13

carcinoma cells). They were also able to induce apoptosis by cleaving the supercoiled

plasmid DNA in vitro.44

Fe

NO2

NO2

O

O2N

Fe

Cl

Cl

Cl

O

O

Figure 1.9: The iron-based compounds, ferrocenium picrate (left) and ferrocenium trichloroacetate (right). 1.3.5 Cobalt Compounds

Cobalt alkyne complexes exhibit good potential as antitumour drugs and were first

reported in studies against murine leukemic cells. Further investigations and extensive

efforts have been carried out on the most active compound, the hexacarbonyldicobalt

complex of the propargylic ester of acetylsalicylic acid (Co-ASS) (Figure 1.10) against

3677 (human melanoma), H2981 (lung adenocarcinoma),45 MCF-7 (human breast

carcinoma with positive estrogen receptor) and MDA-MB-231 (human breast carcinoma

with negative estrogen receptor) cell lines.46 Studies of inhibitory potential revealed that

Co-ASS inhibits the cyclooxygenase enzyme (COX-1 and COX-2). A synergistic effect

was produced by the combination of Co-ASS with other antitumour drugs, for instance

tyrosine kinase inhibitor in HL-60 (human acute promyelocytic leukemia) and LAMA-

84 and CML-T1 (human chronic myeloid leukemic) cell lines in vitro.47

14

O

O

O

O

Co2(CO)6

Figure 1.10: The cobalt-based compounds, hexacarbonyldicobalt complex of the propargylic ester of acetylsalicylic acid (Co-ASS).

1.3.6 Gold Compounds

Early studies reported that gold-based compounds greatly inhibited the growth of

cultured cancer cells. Auranofin (Figure 1.11), gold-based compound was proven to

inhibit DNA, RNA and protein synthesis, where these inhibitions could affect the

propagation of cancer cells. Morphological changes, for example membrane blebbing and

pitting were observed under the exposure of cells to auranofin.48 The development of

gold-based compounds continued with bis(diphenylphosphine)ethane ligand,

[Au(dppe)2]Cl (Figure 1.11). This gold containing compound demonstrated good

cytotoxicity against cultured cancer cells, where it produced DNA-protein crosslinks and

DNA strand breaks in cells after [Au(dppe)2]Cl exposure. However, cardiotoxic effects

were observed which attributed to the disruption of mitochondrial function.49,50 The

successful development of chloro(triethylphosphine)gold(I), [Au(dppe)2]Cl and

auranofin as anticancer drugs by the inhibition of mitochondrial human glutathione

reductase and thioredoxin reductase enzyme was demonstrated, where this enzyme was

commonly involved in the process of cell division.51,52 As a result, the binding of gold

compounds with catalytic residues could cause alteration in the cellular process, finally

inducing apoptosis of the cells.

15

O SO

O

H3C

O O

O

H3C

O

H3C

O

AuP

CH3 CH3

H3C

CH3

O

P

P

P

P

Au

Figure 1.11: The gold-based compounds, auranofin (left) and [Au(dppe)2]Cl (right). 1.4 Tin(IV) compounds as potential anticancer agents

Among the non-platinum compounds, tin(IV) compounds have been a promising

new lead for the development of anticancer drugs. Tin, a post-transition metal in period

5 has two 5s and two 5p electrons in its outer shell orbital. Tin can lose two electrons

from the 5p orbital to form Sn2+, or share all four outer electrons with other atoms to

achieve a stable electron configuration. Tin(IV) compounds show a diverse range of

applications, from biological to industrial uses. In recent years, the interest in tin(IV)

compounds in medicinal chemistry has increased significantly, particularly with reference

to the development of anticancer drugs,53–55 which showed their effectiveness against a

number of tumours. The National Cancer Institute (NCI) has tested the largest number of

tin compounds compared to other metals, about 2000 tin-based compounds as compared

to platinum (1600), iron (900) and cobalt (800).56,57

16

Previous studies have reported that tin(IV) compounds may bind to the

glycoproteins or cellular proteins of living organisms, and also may cause cell death via

interaction with DNA.58–61 The general interaction mechanism between small

molecules/drugs and DNA mainly occurs via three non-covalent interactions:

intercalative binding, major/minor groove binding and electrostatic binding.

Spectroscopic and molecular modelling techniques are useful in order to investigate the

DNA binding interactions.62

The earliest studies on the cytotoxicity of tin(IV) compounds against mouse cancer

were carried out in 1929.63 Some 40 years later, a number of tin(IV) compounds

R2SnX2.Ln, (R = Me, Et, Pr, Bu, c-hexyl, and Ph groups; X = halogen; L = mono- (n=2)

or bidentate (n = 1) N/O-donor ligand) were investigated for their in vivo antitumour

activity against lymphocytic leukaemia (P388) in mice.64 It was reported that many

tin(IV) compounds showed promising activity against P388, but were found to be inactive

against some solid tumours tested, e.g. B16 melanocarcinoma, CD8F1 mammary tumour,

CX-1 colon xenograft, colon 38, L1210 lymphoid leukaemia, LX-l lung xenograft, Lewis

lung carcinoma and MX-1 breast xenograft.64 This study revealed that: (a) diethyl and/or

diphenyltin compounds possessed higher activities; (b) there was no real link between the

Lewis acidity of the parent tin(IV) halide and the P-388 inhibition activity; (c) a pre-

dissociation of the bidentate ligand could be the rate determining step in vivo; and (d)

several compounds showed poor activity against human cancer cell lines. These

discoveries triggered a number of tin(IV) compounds studies against several types of

cancer cells.

Tin-based compounds have been a particular focus due to their structural features

and cytotoxic properties, both of which are required for biocompatibility and DNA

cleavage. The structure-activity relationship of tin(IV) compounds was discussed in terms

17

of various factors such as the nature of metal, coordination number and geometry and the

effect of substitution on ligands.56,64 The possible structure-activity relationship was

summarised as follows: a) Diphenyltin(IV) compounds were more potent than

dimethyltin(IV) compounds in inhibiting cancer cell lines. This is due to the planarity of

the two phenyl groups present in the tin(IV) compounds, where it is favourable to interact

to the base pair of the DNA in cancer cell lines by π-interactions, thus causing cell death

by apoptosis.53,65,66 The tin(IV) compounds were also found to bind to the phosphodiester

backbone of DNA, thus disrupting DNA repair in the presence of the phosphodiesterase

enzyme.67 Moreover, tin(IV) compounds bind to glycoproteins and cellular proteins,

which act as cancer biomarkers.59–61

In this chapter, the review of anticancer activity of tin(IV) compounds is mainly

focused on phenyl- and methyltin(IV) compounds reported over the past ten years. The

cytotoxic activity of tin(IV) compounds is compared with the reference drugs used in that

particular study.

1.4.1 Tin(IV) Compounds of Oxygen and/or Nitrogen Donor Ligands

The tin(IV) compounds (1-5, see Figure 1.12) of 1-(4-chlorophenyl)-1-

cyclopentanecarboxylic acid were screened for their in vitro antitumor activities against

HL-60, hepatocellular carcinoma (Bel-7402), gastric carcinoma (BGC-823) and

nasopharyngeal carcinoma (KB) cell lines.68 However, only compound 3 was observed

to exhibit slightly higher antitumour activity than cisplatin, with the concentration of a

drug that inhibits a biological activity by 50% (IC50) values of 5.2 and 8.1 μM (Bel-7402),

and 4.9 and 6.5 μM (BGC-823), respectively.68 In contrast, findings reported by Hadi and

Rilyati, indicated that diphenyltin(IV) compounds (7) (IC50 = 17.89 µM) showed more

potency than dibutyltin(IV) (6) (IC50 = 41.25 µM).69 The cytotoxicity of octahedral

18

tin(IV) compounds with general formulae of R2SnL2 (R = Me (8), Et (9), Bu (10), Ph (11),

Bz (12) and L = 2-phenylmonomethylglutarate) were studied against KB cell line. 70

Among the tin(IV) compounds, compound 11 (IC50 = 0.30 μM) was equipotent cytotoxic

activity with cisplatin (IC50 = 0.37 μM).70

The cytotoxicity of dimethyl- (13, see Figure 1.12), dibutyl- (14, 16, see Figure

1.12) and diphenyl- (15, 17, see Figure 1.12) tin(IV) compounds containing an o-vanillin

(oVa) Schiff base were studied against three cisplatin resistant tumour cell lines, viz.

human lung (A549), HeLa and MCF-7.66 These tin(IV) compounds showed greater

cytotoxicity than cisplatin in all cancer cell lines tested. It was suggested that activity of

the tin(IV) compounds was related to the coordination geometry of the tin(IV)

compounds, since the octahedral tin(IV) compounds (13, 14 and 15) showed higher

cytotoxic activities than the tin(IV) compounds with trigonal bipyramidal geometry (16

and 17). The coordination of the organo groups complexed to the central tin also

influenced cytotoxicity, with cytotoxicity following the order n-Bu > Ph > Me for most

of the tumour cells. The cytotoxic mechanism for these reported compounds remains

unknown. However, protein binding studies with bovine serum albumins (BSA), show

that they bind to the albumin in the blood stream, resulting in a significant impact on the

distribution, free concentration, metabolism and toxicity of the drug. The binding constant

between the tin(IV) compounds and BSA suggested that all of the compounds could be

easily stored in the protein and also could be released in the affected areas.66

Four trigonal bipyramidal tin(IV) compounds, Me2SnL1 (18), Ph2SnL1 (19),

Me2SnL2 (20) and Ph2SnL2 (21) (H2L1 = 5-chlorosalicylaldehyde isonicotinoyl hydrazone

and H2L2 = 2-hydroxy-4-methoxybenzaldehyde isonicotinoyl hydrazone), see Figure

1.12, were synthesised and tested against A549 and HeLa cell lines, showing significant

activity with IC50 values ranging between 0.7 - 15.2 µM (as compared to their free Schiff

19

bases, IC50 > 100 µM and cisplatin, > 60 µM, respectively).53 The higher potency of the

diphenyltin(IV) compounds here were attributed to the presence of the phenyl groups,

which enabled intercalation of the compounds into the DNA base pairs via π-π stacking

interactions,53 confirmed via UV-vis spectroscopy. Drug-DNA binding constant values

of 1.58 x 104 M-1 (18), 2.02 x 104 M-1 (19), 1.47 x 104 M-1 (20) and 3.36 x 104 M-1 (21)

were reported, showing that 19 and 20 had significantly higher affinity to DNA.

The efficiency of dibutyl- and diphenyltin(IV) compounds (22-29, see Figure 1.12),

synthesised from 2-hydroxy-1-naphthaldehyde or 4-substituted-2-hydroxybenzaldehyde,

benzylhydrazine and tin(IV) oxide were tested against a murine melanoma (B16F10) cell

line at different doses response ranging from 10, 5, 2.5 and 1 μg mL-1 for 24 hours. These

compounds were stable for several months, except for compound 24. Compounds 22, 25

and 28 were identified as the most cytotoxic compounds, however, all diphenyl- and

dibutyltin(IV) compounds investigated in this study showed significant cytotoxic

activities against the B16F10 cell line.71

In vitro cytotoxicity properties were evaluated for five tin(IV) compounds with o-

hydroxy-benzoic or p-hydroxy-benzoic acids against four cancer cell lines, MCF-7,

HeLa, human osteosarcoma (U2-OS) and one normal cell line (normal foetal lung

fibroblast, MRC-5). The compounds demonstrated high potency against all cancer cell

lines tested (see Table 1.1). Compound 31 showed 304 and 210 times higher potency

compared to cisplatin for MCF-7 and U2-OS cell lines, respectively. Compounds 30 and

34 (see Figure 1.9) also exhibited higher cytotoxicity than 32, but slightly lower than 31.

Toxicity against MRC-5 cells showed no inhibitory activity, except for tributyltin.72

20

Table 1.1: Cytotoxic activity of tin(IV) compounds against a panel cancer cells and survival of MRC-5 cells, compared to cisplatin.72

Compound IC50 (nM) % Survival of

MRC-5 MCF-7 HeLa U2-OS LMS

Dimethyltin(IV) (30) 142 ± 4.3 84 ± 3.2 130 ± 2.3 31 ± 1.6 97.5 ± 0.15

Dibutyltin(IV) (31) 108 ± 2.6 37 ± 1.2 97 ± 2.1 43 ± 1.5 104.1 ± 0.2

Tributyltin(IV) (32) 724 ± 5.4 295 ± 3.8 820 ± 5.7 738 ± 2.4 98.1 ± 0.1

Triphenyltin(IV) (33) 121 ± 3.7 21 ± 1.1 258 ± 3.2 45 ± 1.8 94.5 ± 0.2

Tributyltin(IV) (34) 325 ± 2.3 120 ± 2.9 535 ± 4.6 134 ± 2.8 78.4 ± 0.2

Cisplatin 32850 ± 23900

5540 ± 680

20510 ± 1250

5480 ± 530 -

Dimeric tin(IV) compounds with general formula of R2SnL2 (R = Me, Et, n-Bu, Ph

or Oct, and L = arylhydrazone of β-diketone derivatives) (35-44, Figure 1.12) were

synthesised and examined for their in vitro cytotoxic properties against HeLa, KB and

human liver hepatocellular carcinoma (HepG2) cell lines.73 These compounds were found

to be active against these three cancer cell lines, except compounds 35, 36, 37 and 44.

Dibutyltin(IV) (27-29) and diphenyltin(IV) (42, 43) compounds exhibited greater

inhibition than widely used clinical drugs, with IC50 values in the range of 0.2 to 6.3 μM

(for comparison, the IC50 for cisplatin ranges from 2.65 to 8.3 μM). The bioactivity was

shown in the order of R group, such that Me < Oct < Et < Ph < n-Bu. Cell apoptosis

against the KB cell line was observed with compounds 39, 40 and 44. It was thus

suggested that apoptosis was the principal mechanism of cell death using these

compounds.73

The in vitro cytotoxicity of dimethyltin(IV) compounds of

bis(diethylamine)phosphoric triamide (45, see Figure 1.12) and

bis(cyclopentylamine)phosphoric triamide (46, see Figure 1.12) against MCF-7, MDA-

21

MB-468 and T47-D breast cancer cell lines were studied by Gholivand et al. Both

compounds demonstrated high potency against MDA-MB-468 cells with IC50 values of

7.58 μM (45) and 11.65 μM (46) and the cells are exquisitely sensitive to cell death

compared to other cells.74

SnO

O O

OR

R

Cl

Cl

R = Me (1), Et (2), n-Bu (3), n-Oct (4), Ph (5)

OSn

O

O

OR

R

R = n-Bu (6), Ph (7)

SnO

O O

OR

R

OO O

O

R = Me (8), Et (9), n-Bu (10), Ph (11), Ph-CH2 (12)

N

NN

OO

O

N

NN

OO

O

Sn

R

R

Cl

Cl

Sn

R R

Sn N

NN

OO

O

Sn

R R

R = CH3 (13)R = C4H9 (14)R = C6H5 (15)

R = C4H9 (16)R = C6H5 (17)

R R

N Sn NCl

Cl R

R O

N N

Cl

ONN

OO

SnSn

RR

RR

Cl

R = CH3 (18) R = C6H5 (19)

OCH3

N

SnO

N

ON

R R'

R' = CH3 (20) R' = C6H5 (21)

Figure 1.12: The structure of tin(IV) compounds containing oxygen and/or nitrogen donor ligands.

22

N

O

N

OSn

R R

R = nBu, R' = NO2 (22)R = Ph, R' = NO2 (23)R = nBu, R' = H (24)R = Ph, R' = H (25)

R'

N

O

N

OSn

R R

NO2

R'

R = nBu, R' = Et2N (26)R = Ph, R' = Et2N (27)R = nBu, R' = MeO (28)R = Ph, R' = MeO (29)

O

OSn

O

OOH

HO

R

R

R = CH3 (30)R = n-C4H9 (31)

SnO

O

HO

R

R

R = n-C4H9 (32)R = Ph (33)

R SnO

OR

R

R = n-C4H9 (34)

R OH

O

NSnN

R2 O

R1

R

RO

NSn N

R2O

R1

R

R

R1

OHHN

N

O

O

R2

R1= H, R2 = CH3 (H2L1)R1 = m-NO2, R2 = CH3 (H2L2)R1 = H, R2 = OCH2CH3 (H2L3) [R2SnL]2

R = Me, L = L1 (35)R = Me, L = L2 (36)R = Et, L = L1 (37)R = Et, L = L2 (38)R = Bu, L = L1 (39)

R = Bu, L = L2 (40)R = Bu, L = L3 (41)R = Ph, L = L1 (42)R = Ph, L = L2 (43)R = Oct, L = L2 (44)

SnClO

CH3

CH3

NCl

PR1

R1NH

NH

O

PR1

R1

OSn

CH3

CH3

Cl

Cl

N

O

R1 = N(C2H5)2

(45)

Sn

CH3

CH3

O N

Cl Cl

P

R2R2

N

O

O

NH

P R2

O R2

R2 = NH(C5H9)

(46)

Figure 1.12: Continued

23

1.4.2 Tin(IV) Compounds of Sulphur Donor Ligands

A five-coordinated anionic tin(IV) compound [C5H5NH][Ph2Sn(µ2-SCH2COO)Cl]

(47, see Figure 1.13) of mercaptoacetic acid with diphenyltin, was tested against A549

and colon carcinoma (HCT-8) cell lines with IC50 values of 1.9 μΜ and 0.7 μΜ,

respectively.75 R2Sn(IV) compounds (R = Me (48) and n-Bu (49), see Figure 1.13) of N-

acetyl-L-cysteine (H2NAC) were studied against hepatocellular carcinoma (HCC) and

non-tumour Chang liver cells.76 Cytotoxic data reported that 49 was effective in inhibition

of HCC cells as compared to H2NAC and 49 considered less toxic against non-tumour

Chang liver cells. This indicated that the coordination of anionic NAC2- was able to exert

modulatory activity. Moreover, the mechanism of death in HCC cells of 49 was reported

to relate to apoptotic pathway in inducing activation of caspase-3.76 Tin(IV) compounds

with heterocyclic thioamides, 2-mercapto-benzothiazole (Hmbzt), 5-chloro-2-mercapto-

benzothiazole (Hcmbzt) and 2-mercapto-benzoxazole (Hmbzo) of formulae

[(C6H5)3Sn(mbzt)] (50), [(C6H5)3Sn(mbzo)] (51), [(C6H5)3Sn(cmbzt)] (52) and

[(C6H5)2Sn(cmbzt)2] (53), [(n-C4H9)2Sn(cmbzt)2] (54) and [(CH3)2Sn(cmbzt)2] (55) were

tested for their cytotoxicity against leiomyosarcoma cells from the Wistar rat.77,78 Among

the triphenyltin(IV) compounds (50-52), 52 exhibited higher cytotoxic activity, while

among diorganotin(IV) compounds (53-55) compound 53 showed excellent cytotoxic

potency. Furthermore, compounds (50-55) were reported to strongly inhibit the

metalloenzyme of lipoxygenase (LOX) through a free radical mechanism in the same

manner.77,78

Milaeva and co-workers (2014) studied the mode of cytotoxic action of tin(IV)

compounds Me2Sn(SR)2 (56), Et2Sn(SR)2 (57), (n-Bu)2Sn(SR)2 (58), Ph2Sn(SR)2 (59),

Me3SnSR (60), Ph3SnSR (61) and R2SnCl2 (62) (R = 3,5-di-tert-butyl-4-hydroxyphenyl),

see Figure 1.13. All compounds were significantly active against MCF-7, HeLa and

24

MRC-5 cell lines.79 The cytotoxicity of dimethyl-, dibutyl- and diphenyltin(IV)

compounds (63-65, see Figure 1.13) of N-methyl-N-phenyldithiocarbamate and N-ethyl-

N-phenyldithiocarbamate were investigated against HeLa cell lines by Adeyemi et al.

The compounds exhibited better cytoselectivity (IC50 = 5 μM (dimethyltin(IV); 25 μM

(dibutyltin(IV); 12 μM (diphenyltin(IV)) compared to the standard drug, fluorouracil

(IC50 = 40 μM). It was demonstrated that 63 showed higher potency than 64 and 65, a

finding at odds with an earlier report showing that aryl and alkyl derivatives of organo

substituents with longer chains have better bioactivities.80 Nevertheless, the mechanism

of action of 64 remains unknown.

Coordinating the binuclear diphenyltin(IV) compounds (66-70, see Figure 1.13)

with 4,4’-bis(2-cyclohexylamino)acetamido) diphenylsulfone, 4,4’-bis(2-

isopropylamino) acetamido)diphenylsulfone, 4,4’-bis(n-butylamino)acetamido)

diphenylsulfone, 4,4’-bis(2-cyclohexylamino)acetamido)phenylene and 4,4’-bis(2-

isopropylamino)acetamido)phenylene increased cytotoxic activity against HepG2 cells

22-fold for 66 (IC50 = 3.32 ± 0.14 μM), 44‐fold for 67 (IC50 = 1.77 ± 0.11 μM), 37‐fold

for 68 (IC50 = 2.02 ± 0.47 μM), 8-fold for 69 (IC50 = 8.72 ± 0.19 μM) and 29‐fold for 70

(IC50 = 2.60 ± 0.67 μM), compared with the reference drug cisplatin (IC50 = 75.67 ± 0.51

μM). The toxicity studies demonstrated that the compounds had strong selectivity on

cancer cells over normal liver (WRL-68) cells. The shrinking of HepG2 cells were

observed after the treatment suggesting that the compounds caused apoptosis of the

cells.81

Two dinuclear phenyltin(IV) compounds (71 and 72, see Figure 1.13) of N,N’-

bis(2-hydroxybenzyl)-1,2-ethanebis(dithiocarbamate) ligands were investigated for their

cytotoxicity against colon carcinoma (CoLo205) and mammary tumour Bcap37) cell

lines.82 Both compounds (IC50 = 0.042 ± 0.001 μM (CoLo205), 0.058 ± 0.004 μM

25

(Bcap37) for 71 and 9.62 ± 0.4 μM (CoLo205), 3.00 ± 0.76 μM (Bcap37) for 72) were

more active than cisplatin (13.9 ± 0.5 μM (CoLo205), 5.93 ± 0.083 μM (Bcap37).82 The

cytotoxic activity of the tin(IV) compounds (73-76, see Figure 1.13) of morpholine-1-

carbodithioate was evaluated against HeLa and human myelogenous leukemia (K562).83

The compounds exhibited higher cytotoxicity up to 15-fold compared to cisplatin.

HN

SSn

O

HN

O

O

R

R

R = Me (48) , n-Bu (49)(47)

NSn

SX

Y

PhPh

Ph

X = S, Y = H; [Ph3Sn(mbzt)] (50)X = O, Y = H; [Ph3Sn(mbzo)] (51)X = S, Y = Cl; [Ph3Sn(cmbzt)] (52)

SnN

S S

NR

R

S S

ClCl

R = Ph; [Ph2Sn(cmbzt)2] (53)R = n-Bu; [(n-Bu)2Sn(cmbzt)2] (54)R = Me; [Me2Sn(cmbzt)2] (55)

SSn

O O

Cl

PhPh

But

HO

But

S SnR'2

But

HO

But

S SnR'3

But

HO

But

SnCl2

2 2

R' = Me (56); Et (57); n-Bu (58); Ph (59) R' = Me (60); Ph (61)

(RS)2SnR'2 (RS)SnR'3 (62)

Figure 1.13: The structures of tin(IV) compounds containing sulphur donor ligands.

26

S

S

SnNS

SN

R = CH3 (63)R = C4H9 (64)R = C6H5 (65)

OH

NS

SnS Ph

Ph

XNS

Sn SX

Ph Ph HO

X = Ph (71), Cl (72)

NOS

SnS

S

SR

R

N O

R = Me (73), n-Bu (74), Ph (75), CH2-Ph (76)

SO O

NH

O

N

NH

O

NR

S S

R

S S

SOO

HN

O

NHN

O

N R

SS

R

SSSn Sn PhPhPhPh

R = Cy (66)R = iPr (67)R = nBu (68)N

H

O

NR

SSSn

NH

O

N R

S SSn

HN

O

N R

S S

HN

O

NR

SSPhPhPhPh

R = Cy (69)R = iPr (70)

Figure 1.13: continued Tin(IV) moieties are important for biological activities, in particular, cytotoxicity

and DNA binding studies, the design of the ligand attached to the tin centre also plays a

key role in order to transport and address the compounds to their specific target sites.

27

1.5 Chemistry and Bioactivities of Schiff Bases and Their Tin(IV) Compounds

Schiff bases are considered as “privileged ligands” and they are widely used as

pharmacophores for drug design due to their versatile synthesis and good solubility in

common organic solvents. The basic structure of the Schiff bases comprise stable imines

(C=N), where the basic nitrogen atom is able to stabilise metal ions in different oxidation

states. The nitrogen donor atom is also responsible for enhancing the biological activities,

by interacting via the nitrogen lone electron pair with the protein amino acid residues or

DNA nucleobases. Schiff bases are easily formed by condensation reaction of the

carbonyl group of aldehydes or ketones with primary amines, in particular dithiocarbazate

and thiosemicarbazide amine sources, which will be employed in this thesis.

Dithiocarbazate, NH2NHCS2, and thiosemicarbazide, NH2NHCSNH2, are di-

anionic chelating ligands, and can greatly modified by introducing different organic

substituents. Dithiocarbazates and thiosemicarbazides are capable of forming stable

compounds with a large number of metal elements, including tin, due to the presence of

the anionic nitrogen and sulphur moieties which enable a wide range of binding modes,

thus affording good coordination capacity. The interesting coordination modes and

stability of both dithiocarbazates and thiosemicarbazones with metals occurs where the

metal acts as a Lewis acid complex to react with the electron-rich dithiocarbazate/

thiosemicarbazone ligands according to the Hard Soft [Lewis] Acid Base principle.

Dithiocarbazate and thiosemicarbazone ligands have received significant attention

and they display diverse biological activities.66 However, from a synthetic perspective,

they can be easily derivatised with a range of organic substituents, and can coordinate to

metal ions by coordinating donors to yield complexes with a range of geometrical

structures and chemical/physical properties. They also can exist as thione and thiol

tautomers (Figure 1.14). In solid form, they usually exist in thione form, while in solution

28

they exist as an equilibrium mixture of thione-thiol tautomers (Ali and Livingstone,

1974). The tautomerism character of both these dithiocarbazate and thiosemicarbazone

ligands have been confirmed by FTIR and 1H NMR analysis.85–88,13,89–91

R1S N

H

N

S

R1S N

N

SH

R1NH

NH

N

S

R1NH

NN

SH

Thione form Thiol form

R1 indicate the presence of various different organic substituentsR2 indicate the presence of various different aldehydes/ketones

(a)

(b)

R2R2

R2 R2

Figure 1.14: Tautomerism in (a) DTC and (b) TSC Schiff bases. 1.5.1 Anticancer Activity of Dithiocarbazate Schiff Bases and Their Tin(IV)

Compounds

Dithiocarbazates and Schiff bases thereof have been shown to have promising and

wide-ranging biological activity over the past few decades. Much work has been done on

S-benzyldithiocarbazate (SBDTC) specifically, mostly focusing on SBDTC and S-

methyldithiocarabazate (SMBDTC) transition metal complexes.91–100 A number of

reports have developed dithiocarbazate analogues which specifically show selective

biological activity against certain cancer cell lines.98,101–106 For example, S-benzyl-β-N-

(benzoyl)dithiocarbazate exhibits IC50 = 3.32 µM against HL-60,107 S-2-

picolyldithiocarbazate with pyridine-2-carboxaldehyde exhibits CD50 (the concentrations

causing 50% of cell death) = 7.92 µM against CEM-SS (T-lymphoblastic leukemic cells)

29

and HT-29 (colon cancer cells),108 3-methylbenzyl-2-(6-methylpyridin-2-

ylmethylene)hydrazine carbodithioate exhibits CD50 = 0.95 and 6.36 µM against MCF-7

and MDA-MB-231, respectively,99 6-methyl-2-formylpyridine Schiff bases of SMBDTC

(IC50 = 4.43 µM), and 6-methyl-2-formylpyridine Schiff bases of SBDTC exhibit IC50 =

6.63 µM against Caov-3 (human ovarian cancer cells). The structure of these ligands can

be easily and widely modified via introducing different organic

substituents96,98,101,103,109,110 without substantially altering their coordination chemistry.

Dithiocarbazate derivatives can also be greatly modified, for instance by alkyl or aryl

substitutions at the sulphur atom (see Figure 1.15) forming S-allyl,111 isomeric S-2-/3-/4-

picolyl,108,112 isomeric S-2-/3-/4-methylbenzyl,97,99,103 S-napthylmethyl,107 S-quinolin-

2yl-methyl,107 S-4-nitrobenzyl,98 S-hexyl113,114 and S-4-chlorobenzyl115. Importantly, the

cytotoxic activity of tin(IV) compounds of dithiocarbazate Schiff base analogues remain

essentially unexplored.

The earliest tin(IV) compounds of dithiocarbazate Schiff bases were reported in

1985, where the compounds exhibited promising antibacterial activity against Gram

positive bacteria, Bacillus subtilis and Staphylococcus aureus. Since then, various tin(IV)

compounds were synthesised for various applications. However, the majority of the

studies mainly focused on structural aspects of tin(IV) compounds. The bioactivity of the

tin(IV) compounds remained less explored, particularly for anticancer

applications.85,91,93,116,117 The most recent studies focused on cytotoxicity and toxicity of

tin(IV) compounds SnL2 (73), SnLI2 (74) and SnLCl2 (75) (L = (2,2′-

(disulfanediylbis((ethylthio)methylene)) bis(hydrazin-2-yl-1-ylidene)bis(methanyl

ylidene)) diphenol (see Figure 1.13) against HeLa and MCF-7 cancer as well as Chinese

hamster ovary (CHO) normal cell lines.118 These compounds exhibited good cytotoxic

activity, with compounds L (IC50 = 0.1766 ± 0.022 μM) and 75 (IC50 = 0.05065 ± 0.0226 μM)

30

having IC50 values lower than cisplatin (IC50 = 0.324 ± 0.023 μM) against HeLa cells.

Similarly, compounds L (IC50 = 0.4762 ± 0.0261 μM) and 73 (IC50 = 0.5862 ± 0.0244 μM)

were more active than cisplatin (IC50 = 0.67430 ± 0.023 μM) against MCF-7 cells. This study

concluded that tin(IV) compounds retained cytotoxic activity against certain cell lines even

with the absence of a complexing organo group at the tin(IV) centre.118

S NH

S

NH2R

S-substituted dithiocarbazateWhere, R=

H2C

H2C

H2C

H2C

CH3

H2C

NCH2

N

H2C

H2C NO2

N

H2C

NH2C

H2C

ClH2C

CH2

methyl benzyl allyl 2-picolyl 3-picolyl

4-picolyl 2-methylbenzyl 3-methylbenzyl 4-methylbenzyl

napthylmethyl quinoline-2yl-methyl 4-nitrobenzyl hexyl

4-chlorobenzyl Figure 1.15: Different substituents at position R in dithiocarbazate derivatives. The cytotoxic activities of Sn(acpysme)I3 (76, see Figure 1.16) and the Schiff base

acpysme, derived from condensed 2-acetylpyrazine and S-methyldithiocarbazate, were

studied against MDA-MB-231 cell lines. Only acpysme was found to be active with IC50

31

values of 4.72 μM, while the tin(IV) compound 76 was found to be inactive.85 A hypothesis

that will be tested in this project is that the cytotoxicity of this compound can be improved via

manipulation of the S-substituent, aldehyde and organo groups.

S N

S

N

OSn

SN

S

N

O

S N

S

N

OSn

II

(73) (74)

(75)

S N

S

N

OSn

ClCl

S N

S

N N

NSn

I I

(76)

Figure 1.16: The structure of tin(IV) compounds derived from dithiocarbazate Schiff bases. 1.5.2 Anticancer Activity of Thiosemicarbazone Schiff Bases and Their Tin(IV)

Compounds

Thiosemicarbazones are an important class of compounds which have received

wide attention due to their remarkable biological and pharmacological properties, such as

antibacterial, antiviral, antineoplastic and antimalarial activities.119–122

Thiosemicarbazone Schiff bases are similar to dithiocarbazate and its derivatives, in that

complexation with a metal centre is achieved via the nitrogen and sulphur atoms

following deprotonation of the S-H and N-H groups.98,123–128 Thiosemicarbazones are an

interesting chemotherapeutic platform for the development of radiopharmaceuticals in the

treatment of cancer.129 Some synthetic thiosemicarbazone drugs are already available in

32

the market, such as methisazone (Figure 1.17(a)) used in the treatment of poxviruses, and

Triapine (Figure 1.17(b)), which is a promising anticancer agent following more than 30

clinical phase I/II trials in the treatment of leukaemia (principally advanced leukaemia),

lung, kidney, prostate and pancreatic cancers.130,131 The studies of Triapine in

combination with radiation or other cytotoxic agents have been promising candidates in

treating gynaecological tumours.132 It is proposed that these promising candidates interact

with intracellular iron and/or copper,133 which lead to the production of reactive oxygen

species (ROS). ROS is largely responsible for anti-proliferative activity in cancer cells,

which in turn effects ribonucleotide reductase (RR), the enzyme responsible for the

synthesis of DNA building blocks,134 by destroying the tyrosyl radical of the R2

subunit.135 The rapid growth of tumour cells are frequently characterised by RR

overexpression, thus this enzyme is a promising target for cancer therapy.136 In addition,

thiosemicarbazone compounds also exhibit other mechanisms of action, such as

promoting redox cycling reactions that lead to lysosomal membrane permeabilisation

which demonstrate the potency and selectivity against a range of tumours.137

N

H2N

NNH

S

H2NH2N

S

NH

N

NO

(a) (b)

Figure 1.17: Chemical structure of (a) Methisazone and (b) Triapine. Concerning the bioactivity of thiosemicarbazone Schiff base, in many cases

complexation with a metal ion enables higher activity than the parent thiosemicarbazone

Schiff bases, by lowering the concentration of drugs and essentially overcome some side

33

effects. The tin(IV) compounds (77, 78, 79, see Figure 1.18) of 3-methoxysalicylaldehyde

thiosemicarbazone (H2mstsc) were evaluated for their in vitro cytotoxicity against

immortalised line of human T lymphocyte cells, Jurkat cells. 138 It was observed that the

antitumour activity for dialkyltin(IV) compound increased with the length of the carbon

chain of the alkyl coordinated to the tin atom, while the most potent compound exhibited

an IC50 value of 110 µM. The activity explained the cytotoxicity order as dibutyl > dipehyl

> dimethyl = H2mstsc.138 The ability of 2-acetylpyridine N4-cyclohexyl

thiosemicarbazone Schiff base (L) and the distorted pentagonal bipyramidal tin(IV)

compound, [Ph2Sn(L)(OAc)].EtOH (80, see Figure 1.18) to inhibit tumour cell growth

against HepG2 cells were studied by Liu et al.139 Compound 80 and its ligand showed 3-

fold higher cytotoxic potency compared to the reference drug Mitoxantone (IC50 = 5.3 ±

2.38 µM), with IC50 values of 3.32 ± 0.52 µM IC50 = 10.10 ± 2.07 µM, respectively. The

better activity of 80 was attributed to chelation effects; the reduction in the polarity of the

central metal atom is due to partial sharing of its positive charge with the partial

negatively charged of Schiff base, as well as the nature R group attached directly to the

tin atom.139

In vitro cytotoxic activity of tin(IV) compounds (81-84, see Figure 1.18) of 2-

hydroxy-5-methoxybenzaldehyde and 4-methylthiosemicarbazide were studied against

human colorectal carcinoma (HCT116) cell line by Salam et al.,140 and significant activity

was observed in comparison to the standard drug 5-fluorouracil (IC50 = 7.6 µM). The

coordination of the Schiff base to the tin(IV) centre thus enhanced cytotoxic activity,

particularly for the compounds 83 and 84, where the reported IC50 values were 4.4 and

3.9 µM, which was better than 5-fluorouracil. The enhancement of cytotoxicity of both

compounds was attributed to the presence of phenyl groups coordinating the tin centre

and this phenyl group, which was suggested to interact with biomacromolecules in the

34

cells. A similar conclusion was reached by Li and co-workers,141 who studied

diphenyltin(IV) (86 and 88, see Figure 1.18) compounds of 2-benzoylpyridine N(4)-

phenylthiosemicarbazone and 2-acetylpyrazine N(4)-phenylthiosemicarbazone. These

compounds exhibited higher antitumour activities (against K562) with IC50 values of 5.8

μM and 14.8 μM, respectively, compared to the dimethyltin(IV) compounds (85 and 87)

and their free Schiff bases. Antiproliferative efficiencies of these compounds were found

to follow the order Ph2 > Ph > Me, implicating the structure of the organo group

coordinating the tin metal centre. An independent biological study suggested that the

diffusion, lipophilic character and steric effects associated with the ligand could also be

factors in determining cytotoxic activity in these compounds.140

Tin(IV) compounds of 2,3-dihydroxybenzaldehyde-N(4)-methyl

thiosemicarbazone (89-91, see Figure 1.18) and 2-hydroxy-5-methylbenzaldehyde-N(4)-

methylthiosemicarbazone (92-94, see Figure 1.18) showed significant antitumour activity

against MCF-7 cell line compared to 5-fluorouracil.65 The lowest IC50 values were 4.65

µM and 3.99 µM for diphenyltin(IV) compounds 91 and 94, due to the presence of the

bulky phenyl group chelating the centre tin, which affords π-π interactions with the cell

components of MCF-7 cells. The improvement of cytotoxic activity was also suggested

to be due to the presence of OH/NH groups, which enabled hydrogen bonding with DNA

base pairs.65 Similar studies of tin(IV) compounds of benzoylformic acid 3-hydroxy-2-

naphthoyl hydrazone142 reported that the diphenyltin(IV) compound (96, see Figure 1.18)

was more effective than its dimethyltin(IV) compound (95, see Figure 1.18) in inhibiting

the growth of A549, human hepatocellular carcinoma (SMMC-7721), P388 and HCT116

cell lines. Interestingly however, this dimethyltin(IV) compound exhibited promising

cytotoxicity against human colon carcinoma (WiDr) cell lines, which contradicted a

previous report53,128,143 that dimethyltin(IV) compounds have lower cytotoxic activities

35

compared to diphenyltin(IV) compounds. It was proposed that different cell lines

exhibited different selective resistances to these tin(IV) compounds.144 For instance,

WiDr, SMMC-7721, HCT116 and P388 cell lines demonstrated similar moderate

sensitivity towards compounds 95 and 96. However, the A549 cell line exhibited low

sensitivity to both of the compounds. This study also found that 95 was less toxic than 96

against the non-cancerous human umbilical vein endothelial cell line.145

N

NN

HN

SnS

O OPh Ph

(80)

ON

NHN

SO Sn

R X

R = Me, X = Cl (81)R = n-Bu, X = Cl (82)R = Ph, X = Cl (83)R = Ph, X = Ph (84)

N

NN

HN

SSn

O ORR

R = Me (85)R = Ph (86)

N

N

NN

HN

SSn

O OPhPh

(88)

N

N

NN

HN

SSn MeMeCl

(87)

R2

R1

O

NN

HN

SSn

R3Cl

R1 R2 R3OH H Me (89)OH H n-Bu (90)OH H Ph (91)H Me Me (92)H Me n-Bu (93)H Me Ph (94)

O

O

NN

O

OHSn

R Y R

O

O

NN

O

OHSn

RYR

R = CH3, Y = H2O (95)R = Ph, Y = CH3CH2OH (96)

NN NH2

SO

O

Sn

R R

R = Ph (77), Me (78), n-Bu (79)

Figure 1.18: The structure of tin(IV) compounds derived from thiosemicarbazone Schiff bases.

36

1.6 Mechanism of Action

1.6.1 DNA Binding

Tin(IV) compounds potentially affect the production of energy by inhibiting

multimeric protein complexes (F1F0 ATP synthases), membrane-associated functions,

macromolecular synthesis, protein synthesis and DNA replication.146,147 Determining the

major site of inhibition is crucial to understanding the mechanism of a compound’s

activity. The binding ability of tin(IV) compounds towards DNA depends on the

coordination number and nature of groups bound to the tin centre. In solution, some

triorganotin(IV) compounds undergo spontaneous disproportionation into their

corresponding tetraorganotin(IV) and diorganotin(IV) compounds. While in vivo, the

spontaneous reaction into diorganotin(IV) compounds can occur by loss of the alkyl or

aryl group, via intervention of enzymes such as aromatase.148 Triorganotin(IV)

compounds are highly toxic, particularly those for which the tin centre is coordinated to

propyl, butyl, pentyl, phenyl and cyclohexyl groups.149 Therefore, diorganotin(IV)

compounds could be considered to be the ultimate cytotoxic agents; indeed, they have

been widely observed to display higher cytotoxicity than standard drugs tested against

several types of cancer cells,142,150,151 possibly related to pharmacokinetic effects. In

contrast, monoorganotin(IV) compounds are less potent.152

Tin(IV) compounds induce DNA damage by interaction with the phosphate

backbone, which leads to the contraction of the DNA and thus change in DNA

conformation. Phosphate oxygen and nitrogen atoms of DNA bases act as anchoring sites

for effective binding to tin(IV) compounds. In vitro DNA binding studies of dimethyltin

compounds of (R-/S-)2-amino-2-phenyl-ethanol and dibromoethane153 revealed strong

Sn-OPO3-2 bonding at the 5’-GMP phosphate group via outer-sphere coordination, and

negligible interaction between the tin(IV) compounds and purine nitrogen. UV-vis

37

absorption analysis showed the presence of hyperchromism with increasing DNA

concentrations, indicating electrostatic binding mode of tin(IV) compounds to DNA

phosphate group (which carries a dianion).153

Small aromatic molecules can bind to DNA by intercalating between two adjacent

DNA base pairs at either the major or minor groove sites on the DNA helix via non-

covalent π-π interactions between the compounds and DNA base-pairs. Figure 1.19 shows

an example of this interaction for a chromone Schiff base. The compound slightly bends

the DNA base pairs that leads to van der Waals interaction and hydrophobic contacts with

DNA functional groups.154

Figure 1.19: Molecular docked model of diphenyltin(IV) complex with DNA dodecamer duplex of sequence d(CGCGAATTCGCG)2 (PDB ID: 1BNA). The image provides the side view of the docked model of complexes (Reproduced with permission from ref. 154). Nevertheless, such interactions are not necessarily correlated with high cytotoxic

activity. For instance, the diethyltin(IV) compound of 1-(2-[(2-Hydroxy-benzylidene)-

amino]-phenylimino-methyl)-7-methoxy-naphthalen-2-ol showed higher cytotoxic

activities against human oral epidermoid carcinoma (KB) cell lines, compared with

analogous diphenyltin(IV), dibutyltin(IV) and dimethyltin(IV) compounds155. Rehman et

al. showed that this diethyltin(IV) compound intercalates between individual DNA

38

strands and also binds with Topoisomerase II via hydrogen bonding at guanine (DG9)

and thymine (DT8) threonine (Thr911) and glutamic (Glu914, Glu831) residues (Figure

1.20), resulting in DNA damage and cell death.

Figure 1.20: a) Molecular docking simulation of diethyltin(IV) complex-DNA complex (binding site of topoisomerase II) b) detailed molecular interactions of diethyltin(IV) complex with amino acid residues (Reproduced with permission from ref. 155). 1.6.2 Apoptosis

Due to their lipophilicity, tin(IV) compounds can interact with the cellular

membrane and activate apoptosis. Apoptosis is a form of programmed cell death which

involves biochemical events, leading to morphological changes such as cell shrinkage,

membrane blebbing, chromatin cleavage, nuclear condensation and formation of

apoptotic bodies of condensed chromatin. Ultimately, the apoptotic bodies are rapidly

engulfed by the neighbouring cells. By contrast, necrosis is unprogrammed cell death in

39

living tissue by autolysis; necrotic cells exhibit loss of membrane integrity, cellular and

nuclear swelling and an associated inflammatory response.156 Unlike necrosis, apoptosis

produces cell fragments with two main death signals - extrinsic pathway via cell surface

receptors and intrinsic pathway though mitochondrial disturbance (Figure 2.14). Both of

these pathways involve p53 tumour suppressor proteins, TRAIL receptor, caspases and

Bcl-2 proteins.157 The p53 tumour suppressor gene is the most mutated gene in human

cancer cells, and the activation of p53 can lead to either cell cycle arrest, apoptosis or

DNA repair. Moreover, the increasing level of nuclear p53 gene in human neuroblastoma

cell line, SY5Y was activated by the trinuclear tin(IV) compound containing L-

tryptophan.57,158 The changes of p53 and histone phosphorylation, DNA damage markers,

which can be activated in response to DNA damage were studied by Esmail et al.159

Tin(IV) compound containing propyl gallate and 1,10‐phenanthroline reported to induce

MCF-7 cell p53-activated apoptosis through triggering DNA damage.159

Apoptosis seems to be the main type of cell death caused by tin(IV) compounds,

even though the path that leads to apoptosis may differ between them.160–162 Gennari and

co-workers61 reported that dibutyltin(IV) compounds initiate an increase of intracellular

Ca2+, which generates reactive oxygen species (ROS) and release of cytochrome C by

mitochondria. ROS include peroxides (R-O-O-R), superoxides (O2-), hydroxyl radicals

(•OH), singlet oxygen (1O2) and alpha-oxygen (α-O),163 and play key roles in cell

signalling and homeostasis.164 However, uncontrollable ROS generation results in

significant damage of cell structures. Similarly, cytochrome C has an intermediate role in

apoptosis.165 The release of uncontrollable ROS and cytochrome C results in activating

caspases, and thus cleaving defined target proteins and irreversible apoptotic damage of

the cells (Figure 1.21). The increase in intracellular Ca2+ may also be due to the disruption

of the cell membrane and cytoskeletal functioning.166 Tin(IV) compounds also could

40

accumulate in the Golgi apparatus and endoplasmic reticulum but not in the plasma

membrane and nucleus of the cells.167–169

Figure 1.21: Mechanism of extrinsic and intrinsic apoptotic pathway (Reproduced with permission from ref. 170). Apoptosis (22.4 ± 1.8%) and necrosis (2.1 ± 0.2%) of leiomysarcoma (LMS) cells

occurred following treatment with 50 nM of dimethyltin(IV) compound 30.72 However,

the percentage of cells that underwent apoptosis decreased to 21.1 ± 2.8% at higher

concentration (100 nM), while the rate of necrosis was essentially the same (1.6 ± 0.6%).

On the other hand, the rate of cell apoptosis increased with increasing concentration of

dibutyltin(IV) (31, 32 and 34, see Figure 1.12) and triphenyltin(IV) (33) compounds. The

characteristic DNA laddering pattern observed by electrophoresis gel indicated that the

LMS cells underwent apoptosis trigged by DNA damage and fragmentation.72

1.7 Knowledge Gap

Tin-based compounds were not considered as serious potential anticancer

candidates in the 1990s. They are now enjoying a resurgence and appear very promising

41

as anticancer drugs, particularly active against cisplatin-resistant cancers. Exciting

progress in medicinal tin(IV) chemistry has been reported in the past few years, allowing

the rational design of novel tin(IV) compounds with various types of substituted ligands.

The preceding literature review has established the potential of tin(IV) compounds in

inhibiting cancer cells over the common organic-based drugs via several mechanisms of

action. However, current literature demonstrates that the cytotoxicity of tin(IV) Schiff

base compounds derived from dithiocarbazate and thiosemicarbazone has not been

considered as good cytotoxic agents. Therefore, this project addresses this knowledge gap

by investigating cytotoxic activity of organotin(IV) and tin(IV) compounds of

dithiocarbazate and thiosemicarbazone Schiff bases. Dithiocarbazate and

thiosemicarbazone Schiff bases were selected as precursors for the development of

tin(IV) compounds and small modifications in their structure, by reacting o-vanillin (oVa)

or 2,3-dihydroxybenzaldehyde (catechol) with dithiocarbazate or thiosemicarbazide

ligands were investigated for their effect on cytotoxic activities. This project also

investigated DNA binding ability of the compounds as well as molecular modelling to

support the experimental data from the theoretical perspective.

1.8 Project Aims

Specific objectives of this project include:

1. Synthesis and structural characterisation of O, N, S tridentate Schiff bases and

their tin(IV) compounds.

2. Investigation of the cytotoxic activity of compounds formed in Aim 1 against a

panel of cancer cell lines

3. Experimental and theoretical investigation of the DNA binding interactions of

bioactive tin(IV) compounds obtained from Aim 2.

42

1.9 Thesis Outline

The synthesis and structural evaluation of oVa dithiocarbazate Schiff bases (1-3)

are presented and discussed in Chapter 2. Chapter 3 details the synthesis of diorganotin,

diphenyl- (4-6) and dimethyltin(IV) (7-9) compounds derived from oVa dithiocarbazate

Schiff bases (1-3), as well as their cytotoxicity. Chapter 3 also presents DNA binding

studies and molecular docking and molecular dynamic simulations for highly active

compounds that provide further insight regarding their mechanism of action. The

synthesis, characterisation and cytotoxicity of diphenyl- (13-15) and dimethyltin(IV) (16-

18) compounds of catechol dithiocarbazate Schiff bases (10-12) are presented in Chapter

4. The synthesis, structural study and biological evaluation of homoleptic tin(IV)

compounds (19-24) containing of two molecules of Schiff bases (1, 2, 3, 10, 11 and 12)

are reported in Chapter 5. Chapter 6 describes the synthesis, characterisation and

cytotoxicity of tin(IV) compounds (29-40) derived from thiosemicarbazone Schiff bases

(25-28) as well as their in-silico studies. Finally, the overall findings of this study are

summarised in Chapter 7 and recommendations for future study are suggested.

43

CHAPTER 2

SYNTHESIS, STRUCTURAL EVALUATION AND CYTOTOXICITY STUDIES OF O-VANILLIN DERIVED DITHIOCARBAZATE SCHIFF BASES

2.1 Introduction

As described in the Aims and Objectives section of Chapter 1, dithiocarbazate was

selected as precursor for the development of anticancer drug candidates. In

pharmaceuticals, vanillin (4-hydroxy-3-methoxybenzaldehyde) (Figure 2.1a) has been

used to synthesise trimethoprim (antibacterial drug), and papaverine, a drug that is

primarily used in the treatment of visceral spasms, vasospasms, and erectile

dysfunction.171 Ortho-vanillin (2-hydroxy-3-methoxybenzaldehyde) (oVa) (Figure 2.1b)

is a yellow crystalline material present in the extracts of essential oils of many plants. It

has distinctly different properties from its vanillin analogues. oVa-containing Schiff bases

are of immense interest in medicinal chemistry due to their potential biological properties

and a wide range of coordination modes. The azomethine connection between oVa and

the amine moieties (for example, dithiocarbazate) in Schiff bases, has been known to be

responsible for antitumour, antibacterial and antifungal activity.85,87,94,96,108,109,172–174

OH

OO

OO

OH(a) (b)

Figure 2.1: The structures of (a) vanillin and (b) o-vanillin In vitro cytotoxicity screening of two oVa Schiff bases derived from (R)-(+)-2-

amino-3-phenyl-1-propanol and 2-amino-2-ethyl-1,3-propanediol have recently been

evaluated against A549, HeLa, HL-60 and K-562 (chronic myelogenous leukemic) cell

lines.175 However, the concentrations needed to inhibit the cell lines were more than 50

44

μM, so they were considered as inactive compounds. Generally, most biological functions

of therapeutic agents are strongly dependent on the compound structure. Dithiocarbazate,

NH2NHC(=S)S-, and more specifically dithiocarbazate-substituted derivatives, have

attracted the attention of researchers for decades,84 since transition metal complexes of

these anions relate to potential biological activity.89,101,110,176 It is hypothesised, in this

project that the cytotoxic activity of dithiocarbazate Schiff base can be improved via

incorporation of oVa. This chapter investigates this hypothesis, and presents the

synthesis, structural characterisation and cytotoxic activities of three dithiocarbazate

Schiff bases derived from oVa.

This chapter contains material from published articles, adapted from reference177

with permission from the Elsevier and from reference178 with permission from MDPI.

Copyright permission attached in Appendices, page 222 and 223.

2.2 Experimental and Computational Methods

2.2.1 Materials

All solvents and reagents were of analytical reagent grade and used without further

purification. Chemicals: Hydrazine hydrate, 80% (Fluka, Buchs, Switzerland),

benzylchloride, ≥ 99% (Merck, New York, NY, USA), 2-methybenzyl chloride, 99%

(ACROS, Carson, CA, USA), 4-methylbenzyl chloride (ACROS, Carson, CA, USA),

potassium hydroxide (HmbG, Hamburg, Germany), carbon disulfide (BDH, Radnor, PA,

USA), 2-hydroxy-3-methoxybenzaldehyde (Merck, New York, NY, USA). Solvents:

Acetonitrile (Baker, Sanford, ME, USA), absolute ethanol, 99.8% (Scharlau, Barcelona,

Spain), ethanol, 95% (John Kollin Corporation, Midlothian, UK), methanol (Fisher,

Pittsburgh, PA, USA) and dimethylsulfoxide (Scharlau, Barcelona, Spain).

45

Chemicals used for biological assay were purchased as follows: RPMI-1640

medium with L-glutamine (without sodium bicarbonate) (Sigma Aldrich, St. Louis, MO,

USA), Penicillin/Streptomycin solution (Biowest, Riverside, MO, USA), Fetal Bovine

Serum (sterile filtered) (Biowest, Riverside, MO, USA), Trypsin 0.25%, EDTA in HSSS

without calcium, magnesium, with phenol red (Biowest, Riverside, MO, USA),

Phosphate Buffered Saline (PBS) 10× concentrate (Sigma Aldrich, St. Louis, MO, USA),

dimethyl sulfoxide (DMSO) (Fisher Scientific, Pittsburgh, PA, USA).

2.2.2 Synthesis

2.2.2.1 S-N-R-Benzyldithiocarbazates (N = 2, 3, R = Methyl)

The synthetic procedure used here was adapted from Tarafder and co-

workers.97,103,179 Potassium hydroxide (11.4 g, 0.2 mol) was dissolved in ethanol (70 cm3,

90%). Then, hydrazine hydrate (10 g, 0.2 mol) was added and the mixture was maintained

at 0 °C in an ice-salt bath. Carbon disulfide (15.2 g, 0.2 mol) was added dropwise with

vigorous stirring (750 rpm) over a period of 1 h. The two layers that formed were

separated and the light-brown lower layer was dissolved in 40% ethanol (60 cm3) below

5 °C. The mixture was kept in an ice-bath and 2-methylbenzyl chloride/4-methylbenzyl

chloride/benzyl chloride (0.2 mol) was added dropwise with vigorous stirring. The sticky

white product, which formed, S-2-methylbenzyldithiocarbazate (S1), S-4-

methylbenzyldithiocarbazate (S2) or S-benzyldithiocarbazate (S3), was filtered and left

to dry overnight in a desiccator over anhydrous silica gel. The products were then kept in

a freezer. The melting points of S-substituted dithiocarbazate derivatives are detailed in

Table 2.1.

46

Table 2.1: Physical data of synthesised S-substituted dithiocarbazate derivatives

Dithiocarbazates Yield (%) Melting point (°C) Related literature

S1 52 171.0-171.8 Lit. m.p: 171.2°C;103

S2 67 160.0-161.2 Lit. m.p: 86.0-89.0°C;97

S3 80 121.3-122.0 Lit. m.p: 125.0°C;179 2.2.2.2 S-2-Methybenzyl-β-N-(2-hydroxy-3-methoxybenzylmethylene)

dithiocarbazate (1)

The synthesis of 1 was adapted from literature.101,110 S1 (2.12 g, 10 mmol) was

dissolved in hot acetonitrile (100 cm3) and added to an equimolar amount of 2-hydroxy-

3-methoxybenzaldehyde (1.52 g, 10 mmol) in absolute ethanol (20 cm3). The mixture was

heated (80 °C) with continuous stirring for about 30 min and then allowed to stand

overnight at room temperature. The resultant product was recrystallised from

CH3CN/EtOH (1:1) to produce a light yellow crystalline solid that was filtered and

washed with cold absolute ethanol. Yield: 65%. M.p.: 180–183 °C. Elemental analysis:

calculated for C17H18N2O2S2: C, 58.93; H, 5.24; N, 8.09. Found: C, 59.75; H, 5.25; N,

7.81. FT-IR (ATR, cm−1): 3084, (N–H); 1600, (C=N); 1117, (N–N); 1026, (C=S). 1H

NMR (DMSO-d6) δ (ppm.): 13.34 (s,1H, NH), 9.57 (s, 1H, OH), 8.51 (s, 1H, CH), 6.75–

7.34 (multiplet, 7H, Ar–H), 4.40 (s, 2H, CH2), 3.76 (s, 3H, O–CH3), 2.30 (s,3H, Bz–CH3);

13C NMR (DMSO-d6) δ (ppm.): 196.1 (C=S), 148.6 (C=N); 148.6, 147.4, 144.9, 137.4,

134.4, 130.8, 130.7, 128.2, 126.7, 120.0, 118.8, 114.4 (aromatic-C), 56.4 (O–CH3), 36.9

(CH2), 19.4 (CH3). m/z calculated for C17H18N2O2S2: 346.47, found: 346.15.

47

2.2.2.3 S-4-Methybenzyl-β-N-(2-hydroxy-3-methoxybenzylmethylene)

dithiocarbazate (2)

The synthesis of 2 was similar to 1, detailed above. S2 (2.12 g, 10 mmol) was

dissolved in hot acetonitrile (100 cm3) and added to an equimolar amount of 2-hydroxy-

3-methoxybenzaldehyde (1.52 g, 10 mmol) in absolute ethanol (20 cm3). The mixture was

heated (80 °C) with continuous stirring for about 30 min and then allowed to stand

overnight at room temperature. The resultant product was recrystallised from

CH3CN/EtOH (1:1) to produce a light yellow crystalline solid that was filtered and

washed with cold absolute ethanol. Yield: 68%. M.p.: 177–178 °C. Elemental analysis:

Calculated for C17H18N2O2S2: C, 58.93; H, 5.24; N, 8.09. Found: C, 58.48; H, 5.18; N,

8.76. FT-IR (ATR, cm−1): 3092, (N–H); 1598, (C=N); 1118, (N–N); 1030, (C=S). 1H

NMR (DMSO-d6) δ (ppm.): 13.32 (s,1H, NH), 9.61 (s, 1H, OH), 8.51 (s, 1H, CH), 6.97–

7.24 (multiplet, 7H, Ar–H), 4.39 (s, 2H, CH2), 3.76 (s, 3H, O-CH3), 2.23 (s,3H, Bz–CH3);

13C NMR (DMSO-d6) δ (ppm.): 196.2 (C=S), 148.6 (C=N); 148.6, 147.4, 145.0, 137.0,

134.0, 129.7, 129.6, 120.0, 118.8, 114.4 (aromatic-C), 56.4 (O–CH3), 38.0 (CH2), 21.2

(CH3). m/z calculated for C17H18N2O2S2: 346.47, found: 346.10.

2.2.2.4 S-Benzyl-β-N-(2-hydroxy-3-methoxybenzylmethylene) dithiocarbazate (3)

The synthesis of 3 was similar to 1, detailed above. S3 (1.98 g, 10 mmol) was

dissolved in hot absolute ethanol (100 cm3) and added to an equimolar amount of 2-

hydroxy-3-methoxybenzaldehyde (1.52 g, 10 mmol) in absolute ethanol (20 cm3). The

mixture was heated (78 °C) with continuous stirring for about 30 min and then allowed

to stand overnight at room temperature. The resultant product was recrystallised from

ethanol to yield light yellow crystals that were filtered and washed with cold absolute

ethanol. Yield: 70%. M.p.: 173–174 °C. Elemental analysis: calculated for C16H16N2O2S2:

48

C, 57.81; H, 4.85; N, 8.43. Found: C, 58.21; H, 4.87; N, 8.28. FT-IR (ATR, cm−1): 3089,

(N–H); 1598, (C=N); 1125, (N–N); 1030, (C=S). 1H NMR (DMSO-d6) δ (ppm.): 13.34

(s,1H, NH), 9.58 (s, 1H, OH), 8.52 (s, 1H, CH), 6.77-7.37 (multiplet, 8H, Ar–H), 4.45 (s,

2H, CH2), 3.77 (s, 3H, O–CH3); 13C NMR (DMSO-d6) δ (ppm.): 196.1 (C=S), 148.6

(C=N); 148.6, 147.4, 144.9, 137.3, 129.8, 129.0, 127.8, 120.0, 118.8, 114.4 (aromatic-C),

56.4 (CH3), 38.1 (CH2). m/z calculated for C16H16N2O2S2: 332.44, found: 332.10.

2.2.3 Physical Measurements

Melting points were determined using an Electrothermal digital melting point

apparatus (Cole-Parmer, Staffordshire, UK). IR spectra were recorded using the Perkin

Elmer Spectrum 100 with Universal ATR Polarisation (PerkinElmer, Boston, MA, USA)

in the range 4000-280 cm−1. C, H and N elemental analyses were carried out using a

LECO CHNS-932 instrument (LECO, Saint Joseph, MI, USA). Electronic spectra were

recorded on a Shimadzu UV-1650 PC recording spectrophotometer (1000-200 nm)

(Shimadzu, Tokyo, Japan). 1H and 13C NMR spectra were recorded using an NMR JNM

ECA400 spectrometer (JEOL, Peabody, MA, USA) with tetramethylsilane (TMS) was

used as an internal standard. The mass spectra were recorded using a Shimadzu GC-MS

QP2010Plus mass spectrometer (Shimadzu, Tokyo, Japan).

2.2.4 Single Crystal X-ray Structure Determination

Crystals suitable for single-crystal X-ray diffraction studies were obtained for 3.

Unfortunately, suitable crystals for 1 and 2 were not able to be obtained despite repeated

crystal-growth attempts using various techniques and solvent systems. The intensity data

for 3 were measured at T = 100 K on an Oxford Diffraction Gemini E CCD diffractometer

(Oxford Diffraction Ltd., Agilent Technologies, Santa Clara, CA, USA) fitted with Mo

49

Kα radiation (λ = 0.71073 Å). Data reduction, including analytical absorption correction,

was accomplished with CrysAlisPro (Oxford Diffraction Ltd., UK).180 The structures

were solved by direct methods181 and refined (anisotropic displacement parameters, C-

bound H atoms in the riding model approximation) on F2. A weighting scheme w =

1/[σ2(Fo2) + (aP)2 + bP] where P = (Fo2 + 2Fc2)/3).182 The molecular structure diagram

was generated with ORTEP for Windows183 with 50% displacement ellipsoids and the

packing diagrams were drawn with DIAMOND.184 Crystal data, data collection and

structure refinement details are summarized in Table 3.2. The carbon-bound H-atoms

were placed in calculated positions (C-H = 0.95–0.99 Å) and were included in the

refinement in the riding-model approximation, with Uiso(H) set to 1.2Ueq(C). The oxygen-

and nitrogen- bound H-atoms were located in a difference Fourier map but were refined

with distance restraints of O-H = 0.84±0.01 Å and N-H = 0.88±0.01 Å, and with Uiso(H)

set to 1.5Ueq(O) and 1.2Ueq(N).

2.2.5 Density Functional Theory (DFT) Calculations

All DFT calculations were performed using Gaussian09185 and Gaussview5186

software. The molecular structures and geometries of the Schiff bases were fully

optimised from the X-ray crystallographic structures using DFT method with the

B3LYP187,188 hybrid exchange correlation functional with a 6-311G(d,p) Pople basis set

for all atoms. Vibrational frequencies were calculated via the harmonic approximation to

ensure that all optimised structures corresponded to local minima of the potential energy

surface. All DFT vibrational frequencies were scaled using a scaling factor of 0.9682.189

The electronic stabilities of the optimised geometries were computed using the time-

dependent density functional theory (TD-DFT) formalism190,191 and included solvation

50

effects (DMSO) via the polarisable continuum method (PCM),192–194 using the same basis

set.

Table 2.2: Crystal data and refinement details for Schiff base 3 Formula C16H16N2O2S2

Molecular weight 332.43 Crystal system, space group Monoclinic, C2/c Temperature (K) 100 a, b, c (Å) 20.7896 (10), 4.6965 (2), 32.5217 (13) β (°) 95.004 (4) V (Å3) 3163.3 (2) Z 8 Radiation type Mo Kα μ (mm-1) 0.35 Crystal size (mm) 0.25x0.15x 0.07 Data collection Diffractometer Agilent Xcalibur Eos Gemini Absorption correction Multi-scan (CrysAlis PRO; Agilent, 2011) Tmin, Tmax 0.87, 0.98 No. of measured, independent and observed [I > 2σ(I)] reflection 6979, 3270, 2595 Rint 0.033 (sin θ/λ)max (Å-1) 0.628 Refinement, R[F2 > 2σ(F2)], wR(F2), S 0.041, 0.097, 1.06 No. of reflections 3270 No. of parameters 206 No. of restraints 2 H-atom treatment H atoms treated by a mixture of independent and constrained refinement ∆ ρmax, ∆ρmin (e Å-3) 0.33, -0.24

2.2.6 MTT Assays

The inhibitory effect of the synthesised compounds on the growth of bladder (EJ-

28 and RT-112) cancer cells was evaluated by MTT assay. The Schiff bases were

dissolved in DMSO and diluted in culture medium, where the final concentration of

DMSO is 0.05% (v/v) and was not toxic to the cancer cells. RT-112 and EJ-28, human

bladder cancer cell lines used in this study (ATCC, Virginia, USA) were cultured in

51

RPMI-1640 (High glucose) medium supplemented with 10% fetal bovine serum

containing 1% penicillin. The cells were cultured at 37 °C in a humidified atmosphere of

5% CO2 in air. Cells were seeded in 96-well plates at 6000 cells per well for 24 h, then

they were treated with different concentrations of compounds for 72 h. The medium was

subsequently removed and the wells were washed with 200 µL of phosphate buffer saline.

Aliquots of 20 µL of 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide

(MTT) was added to each well and incubated for 4 h at 37 °C. Aliquots of 200 µL of

media was then removed from each well and 200 µL of DMSO was added. Optical density

(OD) was measured using an ELISA plate reader at 570 nm. Control wells (100%

viability), containing media and DMSO, were included in all the experiments. All data

points represented an average of triplicate assays. Cytotoxicity was expressed as GI50,

i.e., the concentration that reduced the absorbance of treated cells by 50% with reference

to the control (untreated cells).195

For HT29 (colon), U87 and SJ-G2 (glioblastoma), MCF-7 (breast), A2780

(ovarian), H460 (lung), A431 (skin), Du145 (prostate), BE2-C (neuroblastoma), MIA

(pancreas) cell lines and one normal breast cell line, MCF-10A (normal breast), the MTT

assays were performed using the following methods:

Cell Culture and Stock Solutions. Stock solutions were prepared as follows and stored

at −20 °C: Trial compounds were stored as 10 mM solutions in DMSO. All cell lines were

cultured in a humidified atmosphere 5% CO2 at 37 °C. The cancer cell lines were

maintained in Dulbecco’s modified Eagle’s medium (DMEM) (Trace Biosciences)

supplemented with 10% foetal bovine serum, 10 mM sodium bicarbonate, penicillin (100

IU/mL), streptomycin (100 μg/mL) and glutamine (4 mM). The normal cell line, MCF-

10A was cultured in DMEM:F12 (1:1) cell culture media, 5% heat inactivated horse

serum, supplemented with penicillin (50 IU/mL), streptomycin (50 μg/mL), 20mM

52

Hepes, L-glutamine (2 mM), epidermal growth factor (20 ng/mL), hydrocortisone (500

ng/mL), cholera toxin (100 ng/mL) and insulin (10 μg/mL).

In Vitro Growth Inhibition Assay. Cells in logarithmic growth were transferred to 96-

well plates. Cytotoxicity was determined by plating cells in duplicate in 100 μL medium

at a density of 2500–4000 cells/well. On day 0 (24 h after plating), when the cells were

in logarithmic growth, 100 μL medium, with or without the test agent, was added to each

well. After 72 h drug exposure, growth inhibitory effects were evaluated using the MTT

(3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide) assay and absorbance

read at 540 nm. Percentage growth inhibition was determined at a fixed drug

concentration of 25 μM. A value of 100% was indicative of complete cell growth

inhibition. Those analogues showing appreciable percentage growth inhibition underwent

further dose-response analysis allowing for the calculation of a GI50 value. This value is

the drug concentration at which cell growth is 50% inhibited based on the difference

between the optical density values on day 0 and those at the end of drug exposure.196,197

2.3 Results and Discussion

2.3.1 Synthesis

The Schiff bases, 1, 2 and 3, were prepared by condensation reaction of S-N-R-

benzyldithiocarbazates (N = 2, 3, R = methyl) and 2-hydroxy-3-methoxyenzaldehyde

(oVa) in equimolar ratios (Scheme 2.1). The Schiff bases were obtained in fairly good

yields of above 60 %. All the Schiff bases are non-hygroscopic, stable at room

temperature and soluble in common organic solvents especially dimethylsulfoxide

(DMSO) and dimethylformamide (DMF). The melting points of Schiff bases were sharp

(below 5 °C difference) indicating that the Schiff bases were free of impurities. All the

analytical data obtained were in good agreement with the proposed molecular structures.

53

2.3.2 IR Spectral Analysis

The IR spectra indicated successful formation of the dithiocarbazate Schiff bases,

1, 2 and 3. Infrared spectra of Schiff bases were measured in the range 4000 to 280 cm−1

and verified using frequencies predicted using DFT. All DFT vibrational frequencies

were scaled using a scaling factor of 0.9682.189 The Schiff bases have a thioamide (–NH–

CS–) functional group and thus, can exist either in the thione (Figure 2.2a) or thiol (Figure

2.2b) forms or as a mixture of both thione and thiol tautomeric forms. In the solid-state,

the Schiff bases are primarily in the thione form as shown by the presence of the v(NH)

band at 3089 cm−1 and absence of the v(SH) band at c.a. 2600 cm−1.95 Complete IR data

are provided in Table A2.1.

Scheme 2.1: Synthetic pathway for the formation of 1-3.

C SS+ +KOH

HN

NH2

CSS

K

HO

HH2N

NH2

ClCl Cl

S

S

NH

N

HOO

S

S

NH

N

HOO

S

S

NH

N

HOO

HO

O

O

S

S

NH

NH2S

S

NH

NH2 S

S

NH

NH2

S3S1 S2

HO

O

OHO

O

O

1

2

3

54

NH

NS

SHO

O

(a)

NN

S

SHHO

O

(b)

R R

R = o-CH3 (1)R = p-CH3 (2)R = H (3)

Figure 2.2: The (a) thione and (b) thiol tautomeric forms of the Schiff bases 1-3.

2.3.3 NMR Spectroscopic Analysis

While IR spectra indicate that the dithiocarbazate Schiff bases 1, 2 and 3 exist as

thione tautomers, a previous study reported that the thione tautomers of similar Schiff

bases containing dithiocarbazate are relatively unstable in solution, and have a high

tendency to convert into thiol tautomers by enethiolisation.198 The solution-phase

structure of 1, 2 and 3 was therefore probed here via 1H and 13C NMR spectroscopy.

The absence of the exchangeable SH signal in the 1H NMR spectra of the Schiff

bases indicated the predominance of the thione tautomer in DMSO-d6.94 The spectra of

the Schiff bases also exhibited a downfield signal at ~13.32–13.34 ppm due to the NH

group. A singlet appeared in the downfield region of the 1H NMR spectra of the Schiff

bases, which corresponded to a sp2-C–OH proton. The presence of a methylene (-CH2)

proton between the sulphur atom and benzene ring was clearly observed at ~4 ppm,

slightly deshielded due to the close proximity to the electronegative sulphur atom. The

singlet peak was observed at ~8 ppm which corresponded to the –CH proton. The –CH

proton appeared at the deshielded region due to the electron repulsion from the

electronegative nitrogen atom. The singlet signal appearing at 2.30 and 2.23 ppm for 1

and 2, respectively, corresponding to a sp3 CH3 protons located at ortho and para position

55

of the benzene ring in the dithiocarbazate backbone. The methylene proton of the

methoxy group (O-CH3) was clearly observed at 3.76 (1 and 2) and 3.77 ppm (1). The

13C NMR spectra of the dithiocarbazate Schiff bases showed a downfield chemical shift

at ~196 ppm, indicating the existence of the C=S group and thus the predominance of the

thione tautomer in solution. The NMR details for 1, 2 and 3 are tabulated in Tables A2.2

and A2.3.

2.3.4 Mass Spectral Analysis

Mass spectral data for the Schiff bases 1-3 were recorded in DMSO and were found

to be consistent with the proposed formulation of the Schiff bases. Mass spectra displayed

prominent peaks at m/z 346.15 and 346.10 for 1 and 2, respectively, corresponding to

[C17H18N2O2S2]+, and m/z 332.10 for 3 corresponding to [C16H16N2O2S2]+. The mass

spectra for 1, 2 and 3 are supplied in Figure A2.1.

2.3.5 UV-vis Absorption Spectroscopy

The experimental UV-vis spectra of Schiff bases 1, 2 and 3 measured in DMSO

exhibited prominent absorption bands at 340–348 and 371–389 nm. The corresponding

peaks were predicted at 334 and 377–378 nm using B3LYP/6-311G(d,p). The highest

occupied molecular orbital and lowest unoccupied molecular orbital (HOMO-LUMO)

electron density distribution of the Schiff bases (see Figure 2.3) showed that the HOMO

was predominantly centred on the thiolate group and azomethine moieties of the

dithiocarbazate backbone in each of the compounds investigated here. Conversely, the

LUMO was consistently centred on the 2-hydroxy-3-methoxyphenyl ring, as well as on

the backbone of the dithiocarbazate moieties. Thus, the 334 nm absorption peak was

attributed to the n→π* transitions associated with the non-bonding electron pair of the

56

azomethine nitrogen and sulphur atoms. The π→π* transition was also observed related

to electron delocalisation in the aromatic rings. Complete experimental and calculated

UV-vis data are listed in Table A2.4.

Figure 2.3: HOMO-LUMO of (a) 1, (b) 2 and (c) 3.

2.3.6 X-ray Crystallography

Compound 3, Figure 2.4, comprises two almost planar regions, one being the

phenyl ring, the other being the remaining 14 heavy (non-hydrogen) atoms. The

maximum deviations from the least-squares plane through the latter plane, with a r.m.s.

deviation (RMSD) = 0.0410 Å, are 0.0715(15) Å for the O1 atom and -0.0796(18) for

atom C16. To a first approximation, the molecule can be described as having mirror

symmetry with the 1,4-atoms of the terminal ring being bisected by the plane. The

dihedral angle between the planes (82.72(5))°, substantiates this description, indicating a

very close to perpendicular relationship. The observed planarity in the larger fragment

may be ascribed, in part, to the presence of an intramolecular hydroxy-O-H…N(imine)

hydrogen bond (Table 2.3), which leads to the formation of an S(6) loop. The thione

C1=S2 bond length, 1.670(2) Å, is considerably shorter than the thiol C1-S1 and,

especially, C2-S1 bonds of 1.749(2) and 1.817(2) Å, respectively. The conformation

about the C N bond is E, and the amine-N-H atom is flanked on either side by the thione-

57

S and imine-H atoms. The detailed geometry for 3 is tabulated in the Appendix (Table

A2.5, A2.6, A2.7 and A2.8).

Figure 2.4: The molecular structure of 3, showing the atom- labelling scheme and displacement ellipsoids at the 50% probability level. The most prominent feature of the packing of compound 3 is the formation of

centrosymmetric, eight-membered …HNCS2 synthons through the agency of thio-

amide-N-H…S(thione) hydrogen bonds, Table 2.3. The dimeric aggregates are connected

by phenyl-C-H…O O(hydroxy) interactions to form a supramolecular layer in the bc-

plane, Figure 2.5a. The layers stack along the a axis with no directional interactions

between them, Figure 2.5b.

Table 2.3: Hydrogen-bond geometry (Å, °) of 3 obtained from single crystal X-ray diffraction analysis. The structure is shown in Figure 2.2. D-H…A D-H H…A D…A D-H…A O1-H1O…N2 0.84 (2) 1.90 (2) 2.639 (2) 146 (2) N1-H1N…S2i 0.88 (2) 2.47 (2) 3.3351 (18) 168 (2) C6-H6…O1ii 0.95 2.51 3.453 (3) 170

58

Figure 2.5: Molecular packing in 3: (a) a perspective view of the supramolecular layer sustained by thioamide-N-H…S(thione) and phenyl-C-H…O(hydroxy) interactions and, (b) a view of the unit-cell contents shown in projection down the b axis, highlighting one layer in space-filling mode. The N-H…S and C-H…O interactions are shown as blue and orange dashed lines, respectively. For (a), non-interacting H atoms have been omitted.

2.3.7 In Vitro Cytotoxicity

Compounds 1-3 were screened for in vitro cytotoxicity against a panel of twelve

cancer cell lines: EJ-28 (muscle invasive bladder), RT-112 (minimum-invasive bladder),

HT29 (colon), U87 and SJ-G2 (glioblastoma), MCF-7 (breast), A2780 (ovarian), H460

(lung), A431 (skin), Du145 (prostate), BE2-C (neuroblastoma), MIA (pancreatic) and one

normal breast cell line (MCF-10A) using the MTT metabolic assay. After 72 h of

incubation, a dose-dependent proliferative effect towards the cancer cell lines was

measured at very low micromolar (µM) concentrations of the compounds. The growth

inhibition concentration of the compounds required to inhibit 50% cell proliferation

relative to the cells that were treated with DMSO (control) are given in Table 2.4. The

59

stability of these compounds was assessed in DMSO as well as in a mixture of DMSO-

H2O. The absorbance obtained from UV-vis spectroscopy analysis was monitored for 72

h, and an unchanged pattern in the spectra was indicative that the compounds were stable

in both the solvent systems tested.

The GI50 values for the three Schiff bases showed a slight difference; the presence

of methyl group at ortho (1) and para (2) positions resulted in a slight reduction in

potency as compared to 1. This cytotoxicity difference was potentially due to the

hydrophobicity of the methyl group.199 2 exhibited the lowest potency against HT29 and

MIA with GI50 values of 8.0 ± 3.5 and 11 ± 2.3 µM, respectively. Taking cisplatin into

consideration, 1, 2 and 3 demonstrated equipotent cytotoxicity against EJ-28, RT-112 and

U87 cell lines. On both selective cell lines, HT29 and MIA, 1 and 3 displayed a twofold

increase in activity compared to cisplatin, indicating that the ortho methyl group yielded

better activity than the para methyl group. A twofold increase in activity was also

displayed in MCF-7 cell line treated with 1, 2 and 3. However, 1, 2 and 3 showed lower

activity against A2780, H460, A431, Du145, BE2-C and SJ-G2 cell lines compared to

cisplatin. A previous in vitro study on dithiocarbazate-containing compounds proposed

that these compounds induce apoptosis by DNA fragmentation and suppression of two

key genes, Epidermal Growth Factor Receptor (EGFR) and Mouse Double Minute 2

(MDM2) expression.200

60

Table 2.4: Summary of the in vitro cytotoxicity of compounds 1-3 in several cell lines, determined by MTT assay and expressed as GI50 values with standard errors. (GI50 is the concentration at which cell growth is inhibited by 50% over 72 h).

Compounds Growth Inhibition Concentration, GI50 (µM)

EJ-28 RT-112 HT29 U87 MCF-7 A2780 H460 A431 Du145 BE2-C SJ-G2 MIA MCF10A

1 3.4 ± 0.5

4.6 ± 0.5

3.9 ± 0.7

4.3 ± 0.3

2.7 ± 0.2

3.2 ± 0.2

4.3 ± 0.1

3.8 ± 0.0

4.1 ± 0.1

3.1 ± 0.0

3.0 ± 0.1

4.9 ± 0.4

3.1 ± 0.1

2 3.6 ± 0.4

3.2 ± 0.5

8.0 ± 3.5

3.9 ± 0.0

3.5 ± 0.0

3.1 ± 0.4

5.5 ± 0.7

4.2 ± 0.7

5.4 ± 0.2

3.6 ± 0.3

3.3 ± 0.9

11 ± 2.3

3.5 ± 0.2

3 3.1 ± 0.2

3.7 ± 0.6

2.2 ± 0.0

3.1 ± 0.3

2.5 ± 0.1

3.0 ± 0.1

3.3 ± 0.2

3.0 ± 0.2

3.4 ± 0.3

2.8 ± 0.2

2.4 ± 0.3

4.5 ± 0.3

3.0 ± 0.1

Cisplatin 3.97 ± 0.5

3.56 ± 0.9

11.0 ± 2.0

4.0 ± 1.0

6.5 ± 0.8

1.0 ± 0.1

0.9 ± 0.2

2.4 ± 0.3

1.2 ± 0.1

1.9 ± 0.2

0.4 ± 0.1

8.0 ± 1.0 nd

GI50 (μM): (the colors indicate) =0.1–0.99 (strong); =1.0–9.9 (moderate); =10-100 (weak); =nd (not determined).

61

2.4 Conclusions

This chapter presented three dithiocarbazate Schiff bases derived from oVa which

were synthesised via condensation. The structures of synthesised Schiff bases were

confirmed by various spectroscopic techniques and DFT calculations. The Schiff bases

exist as thione tautomers even in solution which was proven by FT-IR, NMR and mass

spectroscopy analysis. The molecular structure of 3 was authenticated by single crystal

X-ray crystallography. The in vitro anticancer activity against a panel of cancer cell lines,

viz., EJ-28, RT-112, HT29, U87, SJ-G2, MCF-7, A2780, H460, A431, Du145, BE2-C,

MIA and one normal breast cell line (MCF10A) revealed that the oVa containing

dithiocarbazate Schiff bases afforded a modest activities, with slight enhancements

achieved by the introduction of the methyl groups at the ortho or para positions of the

Schiff base phenyl ring. The Schiff bases had no obvious cytotoxic cell line selectivity. It

is suggested that the presence of nitrogen, sulphur and oxygen donor atoms enabled

interaction with the donor atoms of nitrogenous bases of DNA, as well as proteins and

organelles of the cancer cells.

While this chapter demonstrated the synthesis and characterisation of three Schiff

bases with moderate cytotoxic activity, it is hypothesised that this activity could be

improved via their complexation with a metal such as tin. This hypothesis will be tested

in Chapter 3, which reports structural and cytotoxic studies of Schiff bases 1, 2 and 3 with

diphenyl- and dimethyltin, was further improved by complexation with diphenyl- and

dimethyltin. Since DNA is the primary pharmacological target of many chemotherapeutic

compounds, DNA-tin(IV) compounds interactions will also be discussed towards

establishing a mechanism of cancer cell inhibition in these compounds.

62

CHAPTER 3

ORGANOTIN(IV) COMPOUNDS OF O-VANILLIN SCHIFF BASES: SYNTHESIS, STRUCTURAL CHARACTERISATION, IN-SILICO STUDIES

AND CYTOTOXICITY

3.1 Introduction

In recent years, metal compounds with a stable d10 electronic configuration have

received considerable attention due to the wide variety of structures formed that depend

on the coordinating ligands and synthetic procedures.86,139,140,201–207 Tin, a group V metal,

can be found in various synthesis of inorganic and organometallic compounds with a

diverse range of applications. Organotin compounds were first discovered in 1894 by

Edward Frankland, with the first such compound being diethyltindiiodide ((C2H5)2SnI2).

The presence of one or more covalent C-Sn bonds in organotin compounds affects their

activity. In particular, the cytotoxicity of these compounds depends on the nature of alkyl

or aryl substituents coordinated to the tin center. Variation in the number and nature of

substituents in organotin(IV) compound has shown notable effects on their biological

activities, due to the unique characteristics of the compounds, such as structural diversity,

redox capacity and catalytic activity. Furthermore, such ligands play important roles in

transporting and addressing the organotin(IV) compound to specific sites in

biomacromolecules.

Organotin(IV) compounds derived from dithiocarbazate Schiff bases have received

attention due to their ability to stabilise specific stereochemistry. Many mechanistic

studies on organotin(IV) compounds have indicated that these compounds are able to

interact with DNA,151,155 one of the main targets in designing cytotoxic drugs. The

biological activity of organotin(IV) compounds is believed to arise from their tendency

in binding to or cleaving DNA.208,209 Wang et al. reported that organotin(IV) compounds

63

containing Schiff bases could interact with CT-DNA through intercalation of cytotoxic

drugs with serum albumin (which is involved in the transportation of compounds through

the blood stream).142 The nature and magnitude of these interactions could significantly

influence the pharmacokinetics.210,211 Organotin(IV) compounds have been found to

induce DNA damage by specifically binding to both nitrogenous DNA bases and the

phosphate backbone of DNA, thus leading to changes in DNA conformation.57,128,212

Organotin(IV) compounds have also been reported to bind to glycoprotein or cellular

proteins, hence interacting directly with DNA, causing cell death.

The activity of organotin(IV) compounds can be readily modified by the choice of

ligand coordinated to the central tin(IV) ion. Organotin(IV) centres coordinated to hetero-

donor atoms, especially oxygen, nitrogen and sulphur, have been investigated with a

particular focus on their structure-activity correlations.213–215 However, the cytotoxicity

of organotin(IV) compounds containing dithiocarbazate Schiff bases are essentially

unexplored in the literature. It is hypothesised here that complexing dithiocarbazate Schiff

bases with tin enhances the cytotoxicity of the compounds. This chapter investigates this

hypothesis, by reporting the synthesis, structural characterisation and cytotoxic activity

of six organotin(IV) compounds (4-9) of Schiff bases (1-3), see Figure 3.1.

64

SHN

NS

HOO

S NN

SO

OSn

S NN

SO

OSn

R

R R

R = o-CH3 (4)R = p-CH3 (5)R = H (6)

R = o-CH3 (1) R = p-CH3 (2)R = H (3)

R = o-CH3 (7)R = p-CH3 (8)R = H (9)

Figure 3.1: The structures of oVa Schiff bases (1-3) and their organotin(IV) compounds (4-9).

This chapter contains material from published articles, adapted from reference178

with permission from MDPI. Copyright permission attached in Appendices, page 223.

3.2 Experimental

3.2.1 Materials

All solvents and reagents were of analytical reagent grade and used without further

purification. Chemicals: potassium hydroxide (HmbG, Hamburg, Germany),

dichlorodiphenyltin(IV) and dichlorodimethyltin(IV) (Sigma Aldrich, St. Louis, MO,

USA). Solvents: Acetonitrile (Baker, Sanford, ME, USA), absolute ethanol, 99.8%

(Scharlau, Barcelona, Spain), ethanol, 95% (John Kollin Corporation, Midlothian, UK),

methanol (Fisher, Pittsburgh, PA, USA) and dimethylsulfoxide (Scharlau, Barcelona,

Spain).

65

3.2.2 Synthesis

3.2.2.1 Diphenyltin(IV) Compounds

Each Schiff base (0.35 g (1, 2); 0.33 g (3), 1 mmol) from Chapter 2 was dissolved

in absolute ethanol (50 cm3) and mixed with an ethanolic solution (10 cm3) of Ph2SnCl2

(0.34 g, 1 mmol). The resultant yellow solution was refluxed for ca. 6 h. The mixture was

left overnight at room temperature and the reduction in the volume of the reaction mixture

resulted in the deposition of a yellow precipitate which was filtered off and recrystallised

from methanol.

3.2.2.1.1 Diphenyltin(IV) [S-2-Methybenzyl-β-N-(2-hydroxy-3-

methoxybenzylmethylene)dithiocarbazate] (4)

Yellow solid. Yield: 44%. M.p.: 149–150 °C. Elemental analysis: Calculated for

C29H34N2O6S2Sn: C, 56.42; H, 4.24; N, 4.54. Found: C, 56.52; H, 4.97; N, 4.06. FT-IR

(ATR, cm−1): 1588, (C=N); 1076, (N–N); 963, (C=S). 1H NMR (CDCl3) δ (ppm.): 8.77

(s, 1H, CH), 4.47 (s, 2H, CH2), 2.43 (s, 3H, CH3), 3.97 (s, 3H, O–CH3), 6.69-7.94

(multiplet, 17H, Ar–H); 13C NMR (CDCl3) δ (ppm.): 171.9 (C–S), 166.1 (C=N); 159.3,

152.0, 142.0, 136.1, 130.6, 130.4, 130.2, 128.9, 127.9, 126.3, 126.1, 117.2, 116.2

(aromatic-C), 56.6 (O-CH3), 34.4 (CH2), 19.4 (Bz–CH3).119Sn NMR (CDCl3) δ (ppm.):

−236.5.

3.2.2.1.2 Diphenyltin(IV) [S-4-Methybenzyl-β-N-(2-hydroxy-3-methoxy

benzylmethylene)dithiocarbazate] (5)

Yellow solid. Yield: 72%. M.p.: 130–132 °C. Elemental analysis calculated for

C17H17N2O2S2: C, 54.82; H, 4.44; N, 4.41. Found: C, 54.42; H, 4.05; N, 4.18. FT-IR

(ATR, cm−1): 1589, (C=N); 1068, (N–N); 958, (C=S). 1H NMR (CDCl3) δ (ppm.): 8.74

66

(s, 1H, CH), 4.41 (s, 2H, CH2), 2.32 (s, 3H, CH3), 3.94 (s,3H, O–CH3), 6.69-7.93

(multiplet, 17H, Ar–H); 13C NMR (CDCl3) δ (ppm.): 171.8 (C–S), 166.1 (C=N), 159.3,

152.0, 142.0, 137.2, 136.0, 135.8, 133.5, 130.2, 129.4, 129.2, 128.9, 126.1, 117.2, 116.3

(aromatic-C), 56.6 (O–CH3), 36.0 (CH2), 21.2 (Bz–CH3).119Sn NMR (CDCl3) δ (ppm.):

−236.3.

3.2.2.1.3 Diphenyltin(IV) [S-Benzyl-β-N-(2-hydroxy-3-methoxybenzyl methylene)

dithiocarbazate] (6)

Yellow crystals. Yield: 51%. M.p.: 131–133 °C. Elemental analysis calculated for

C28H24N2O2S2Sn: C, 55.74; H, 4.01; N, 4.64. Found: C, 56.35; H, 3.91; N, 4.98. FT-IR

(ATR, cm−1): 1579, (C=N); 1019, (N–N); 958, (C=S). 1H NMR (CDCl3) δ (ppm.): 8.74

(s, 1H, CH), 4.45 (s, 2H, CH2), 3.96 (s, 3H, O-CH3), 6.69-7.93 (multiplet, 18H, Ar–H);

13C NMR (CDCl3) δ (ppm.): 171.6 (C–S), 166.2 (C=N), 152.0, 141.9, 136.7, 136.1, 130.3,

129.3, 128.9, 128.7, 127.5, 126.1, 117.2, 117.1, 116.2, 115.2, (aromatic-C), 56.6 (O–

CH3), 36.1 (CH2). 119Sn NMR (CDCl3) δ (ppm.): −236.3.

3.2.2.2 Dimethyltin(IV) Compounds

Each Schiff base (0.35 g (1, 2); 0.33 g (3), 1 mmol) from Chapter 2 was dissolved

separately in absolute ethanol (50 cm3) and dichloromethane (DCM) (20 cm3) and then

mixed with an ethanolic solution (10 cm3) of Me2SnCl2 (0.22 g, 1 mmol). The resultant

yellow solution was refluxed for ca. 6 h. The mixture was left overnight at room

temperature and kept for a few days yielding brown crystals that were suitable for single

crystal X-ray diffraction analysis.

67

3.2.2.2.1 Dimethyltin(IV) [S-2-Methybenzyl-β-N-(2-hydroxy-3-methoxy benzyl

methylene)dithiocarbazate] (7)

Yellow crystals. Yield: 78%. M.p.: 128–130 °C. Elemental analysis calculated for

C19H22N2O2S2Sn: C, 46.27; H, 4.50; N, 5.68. Found: C, 46.52; H, 4.51; N, 5.82. FT-IR

(ATR, cm−1): 1580, (C=N); 1076, (N–N); 959, (C=S). 1H NMR (CDCl3) δ (ppm.): 8.76

(s, 1H, CH), 4.42 (s, 2H, CH2), 2.42 (s, 3H, CH3), 3.85 (s,3H, O–CH3), 6.69–7.34

(multiplet, 7H, Ar–H), 0.97 (Sn–CH3); 13C NMR (CDCl3) δ (ppm.): 173.9 (C–S), 166.3

(C=N), 158.5, 151.5, 137.2, 134.1, 130.6, 130.4, 127.9, 126.3, 126.1, 116.9, 116.5, 115.9

(aromatic-C), 56.3 (CH3), 34.6 (CH2), 19.4 (Bz–CH3), 7.1 (Sn–CH3).119Sn NMR (CDCl3)

δ (ppm.): −109.4.

3.2.2.2.2 Dimethyltin(IV) [S-4-Methybenzyl-β-N-(2-hydroxy-3-methoxy

benzylmethylene)dithiocarbazate] (8)

Yellow crystals. Yield: 55%. M.p.: 98–102 °C. Elemental analysis calculated for

C19H22N2O2S2Sn: C, 46.27; H, 4.50; N, 5.68. Found: C, 47.07; H, 4.66; N, 6.11. FT-IR

(ATR, cm−1): 1589, (C=N); 1071, (N–N); 958, (C=S). 1H NMR (CDCl3) δ (ppm.): 8.73

(s, 1H, CH), 4.36 (s, 2H, CH2), 2.32 (s, 3H, CH3), 3.84 (s, 3H, O-CH3), 6.69–7.27

(multiplet, 7H, Ar–H), 0.95 (Sn–CH3); 13C NMR (CDCl3) δ (ppm.): 173.7 (C–S), 166.2

(C=N), 158.5, 151.5, 137.1, 133.6, 129.2, 126.1, 119.5, 116.8, 116.4, 115.9 (aromatic-C),

56.3 (CH3), 36.2 (CH2), 21.2 (Bz–CH3), 7.1 (Sn–CH3). 119Sn NMR (CDCl3) δ (ppm.):

−109.2.

68

3.2.2.2.3 Dimethyltin(IV) [S-Benzyl-β-N-(2-hydroxy-3-methoxybenzyl

methylene)dithiocarbazate] (9)

Yellow crystals. Yield: 48%. M.p.: 104–106 °C. Analysis calculated for

C18H20N2O2S2Sn: C, 45.11; H, 4.21; N, 5.85. Found: C, 45.36; H, 4.16; N, 6.02. FT-IR

(ATR, cm−1): 1581, (C=N); 1026, (N–N); 959, (C=S). 1H NMR (CDCl3) δ (ppm.): 8.73

(s, 1H, CH), 4.40 (s, 2H, CH2), 3.85 (s, 3H, O–CH3), 6.69-7.40 (multiplet, 8H, Ar–H),

0.95 (Sn–CH3); 13C NMR (CDCl3) δ (ppm.): 173.6 (C–S), 166.3 (C=N), 158.5, 151.5,

136.8, 129.3, 128.7, 127.4, 126.1, 116.9, 116.5, 115.9 (aromatic-C), 56.3 (O–CH3), 36.4

(CH2), 7.1 (Sn–CH3); 119Sn NMR (CDCl3) δ (ppm.): −109.0.

3.2.3 Physical Measurements

The melting point, CHNS analyser, FTIR, UV-vis and NMR (1H and 13C NMR)

instruments used in this chapter as the same as mentioned in Chapter 2 (§2.2.3). Molar

conductivities of 10−3 M solutions of the organotin(IV) compounds in DMSO were

measured at 27 °C using a Jenway 4310 conductivity meter fitted with a dip-type cell with

a platinised electrode (Cole-Parmer, Staffordshire, UK). 119Sn NMR were recorded using

a JOEL NMR spectrophotometer (JEOL, Peabody, MA, USA).

3.2.4 Single Crystal X-ray Structure Determination

Crystals suitable for single-crystal X-ray diffraction studies were obtained for 7, 8

and 9. Unfortunately, suitable crystals for the diphenyltin(IV) compounds were not

obtained despite repeated crystal-growth attempts using various techniques and solvent

systems. The intensity data for 7, 8 and 9 were measured at T = 150 K on an Oxford

Diffraction Gemini E CCD diffractometer (Oxford Diffraction Ltd., Agilent

Technologies, Santa Clara, CA, USA) fitted with Mo Kα radiation source (λ = 0.71073

69

Å) so that θmax was 29.4°. Data reduction, including analytical absorption correction, was

accomplished with CrysAlisPro.180 The structures were solved by direct methods181 and

refined (anisotropic displacement parameters, C-bound H atoms in the riding model

approximation) on F2. A weighting scheme w = 1/[σ2(Fo2) + (aP)2 + bP] where P = (Fo2

+ 2Fc2)/3)182 was introduced in each case. The crystal of 7 was refined as a twin with the

fraction of the minor component being 0.1664(20); details are given in the deposited CIF.

In the final cycles of the refinement of 9, four reflections, i.e., (−3 4 0), (−1 −1 5), (0 6 0)

and (−3 4 4), were omitted owing to poor agreement. The molecular structure diagrams

was generated with ORTEP for Windows183 with 35, 50 and 70% displacement ellipsoids

for 7, 8 and 9, respectively, and the packing diagrams were drawn with DIAMOND

(Crystal Impact, GbR, Bonn, Germany).184 Additional data analysis was carried out with

PLATON.216 Crystal data and refinement details are given in Table 3.1.

3.2.5 Computational Methods

3.2.5.1 DFT Calculations

DFT calculations were performed as the same software and parameters mentioned

in Chapter 2. The molecular structures and geometries of 4, 5 and 6 were fully optimised

with the B3LYP187,188 hybrid exchange correlation functional in conjunction with the

LanL2DZ pseudopotential on Sn217–219 and 6-311G(d,p) Pople basis set for all other

atoms. The initial single crystal X-ray molecular structures and geometries of 7, 8 and 9

were used for DFT calculations using the same functional and basis set.

70

Table 3.1: Crystal data and refinement details for compounds 7, 8 and 9. Compound 7 8 9 Formula C19H22N2O2S2Sn C19H22N2O2S2Sn C18H20N2O2S2Sn Formula weight 493.19 493.19 479.19 Crystal colour light-yellow light-yellow light-yellow

Crystal size/mm3 0.03 × 0.05 × 0.10

0.05 × 0.13 × 0.20

0.03 × 0.05 × 0.10

Crystal system Monoclinic Triclinic Triclinic Space group Pn P1 P1 a/Å 11.3716(9) 10.5908(4) 8.9408(4) b/Å 12.0262(8) 13.9243(5) 10.4182(4) c/Å 16.3644(15) 14.3974(5) 22.6034(12) α/° 90 95.596(3) 97.713(4) β/° 109.285(9) 93.693(3) 94.600(4) γ/° 90 102.116(2) 111.634(4) V/Å3 2112.4(3) 2058.11(13) 1920.11(16) Z 4 4 4 Dc/g cm−3 1.551 1.592 1.658 F(000) 992 992 960 μ(MoKα)/mm−1 1.422 1.46 1.562 Measured data 6348 18090 15229 θ range/° 3.6–29.4 3.5–29.4 3.4–29.4 Unique data 6348 9402 8720 Observed data (I ≥ 2.0σ(I)) 5090 7605 6671

No. parameters 478 477 457 R, obs. data; all data 0.037; 0.064 0.034; 0.070 0.041; 0.083 a; b in weighting scheme 0.022; 0 0.026; 0.240 0.032; 0.781 Rw, obs. data; all data 0.055; 0.074 0.049; 0.078 0.062; 0.092 GoF 0.99 1.01 1.01 Range of residual electron

density peaks/eÅ−3 −0.56–0.66 −0.46–0.75 −1.03–0.63

3.2.5.2 Molecular Docking Studies

The coordinates of B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2

were obtained from the Protein Data Bank (http://www.rcsb.org/pdb). The coordinates of

the dimethyltin(IV) compounds, 7, 8 and 9 were taken from crystal structures, whereas

the coordinates for 4, 5 and 6 were obtained after DFT geometry optimisation. Molecular

docking studies were carried out using the AutoDock version 4.2.5.1 program (Molecular

Graphics Laboratory, La Jolla, CA, USA).220 Water molecules were removed, polar

71

hydrogen atoms and Kollman charges221 were added to the receptor (DNA sequences). In

the docking analysis, the binding site was assigned across all of the minor and major

grooves of the DNA molecule, which was enclosed in a 64 × 64 × 122 grid box with a

grid spacing of 0.375 Å. AutoDock was ran using the following parameters: 30 docking

trials, population size of 300, maximum number of energy evaluation ranges of 2,500,000,

maximum number of generations of 27,000, mutation rate of 0.02, cross-over rate of 0.8

and the default values were used for other parameters. The parameter file of Autodock

was modified to incorporate van der Waals interactions and other required parameters for

Sn obtained from the Autodock website.222 The lowest binding energy of the

conformation from the highest cluster was selected as the best binding conformation

between the ligand and receptor.

3.2.6 Biological Assays

3.2.6.1 MTT Assays

The protocol used for MTT assay followed that outlined in Chapter 2 (§2.2.6).

3.2.6.2 Quantification of Apoptosis by Annexin V

The apoptotic death of RT-112 cells was measured by fluorescence microscopy

using Annexin V/ PI double-staining. To observe the effect of the diphenyltin(IV)

compounds (4, 5 and 6) on cell apoptosis, RT-112 cells, which are minimum invasive

bladder cancer cells were seeded in 6-well plates at 1 × 106 cells per well. After growth

overnight, the inhibition concentration obtained from the GI50 analysis was used to treat

the cells. After 24 h, the cells were washed with phosphate buffer saline (PBS) and 100

µL of Annexin-V-FLUOS labelling solution containing the incubation buffer and then,

propidium iodide (PI) solution was added. The plate was then incubated for 10–15 min at

72

15–25 °C and analysed by fluorescence microscopy with a Fluorescence Inverted

Microscope (Olympus IX51, Waltham, USA). Green staining of the plasma membrane

indicated that the cells had bound to Annexin-V. The cells that showed red staining were

considered to have lost their membrane integrity. The cells stained green were concluded

to be apoptotic cells; the cells with both green and red staining were concluded as late

apoptotic cells whereas the cells that showed only red staining were considered as necrotic

cells.223

3.2.6.3 Measurement of Reactive Oxygen Species (ROS)

Measurements were performed using a fluorescence plate reader following the

procedure outlined in the ROS detection kit (OxiSelect™). RT-112 cells were cultured in

100 µL growth medium in a 96-well cell culture plate and incubated in an atmosphere of

5% CO2 at 37 °C for a period of 24 h. The growth medium was then removed, the cells

were loaded with an oxidation sensitive cell-permeable fluorescent probe, dichloro-

dihydro-fluorescein diacetate (DCFH-DA), and subsequently stabilised in the highly

reactive DCFH form. In this reactive state, ROS species in living cells were reacted with

DCFH, which is rapidly oxidised to the highly fluorescent 2′,7′-

dichlorodihydrofluorescein (DCF) The plates were then incubated for 1 h and the growth

medium was removed. The cells in each well were washed with PBS and the growth

medium containing 3.05 µM of 5, negative control (DMSO), or positive control (H2O2)

was added. The plates were incubated in an atmosphere of 5% CO2 at 37 °C for a period

of 24 h. After incubation, the cells were transferred to black 96-well plates with the

addition of 2× Cell Lysis Buffer. The cells were then read on a standard fluorescence

microplate reader (Infinite M200). The ROS or antioxidant content in unknown samples

was determined by comparison with the predetermined DCF standard curve.224

73

3.2.6.4 DNA Binding Studies

DNA binding experiments were carried out at 25°C. DNA concentration per

nucleotide was determined using the molar absorption coefficient (6600 M−1 cm−1) at 260

nm.53,225 Solutions of calf thymus (CT) DNA in Tris-HCl buffer containing (5 mM Tris,

pH 7.2, 50 mM NaCl)53 showed A260/A280 of 1.9, indicating that the DNA was sufficiently

free of proteins and contaminants.226 Absorption titration experiments were performed

maintaining the concentration of the organotin(IV) compounds solutions at 50 µM and

gradually increasing the concentration of CT-DNA. The organotin(IV) compounds were

dissolved in DMSO and diluted with Tris-HCI buffer at room temperature (25 °C). The

solutions were scanned over 230–600 nm. Absorbance values were recorded 10 min after

the addition of DNA solution. The binding constant, Kb was determined using equation

(1):

[DNA]

(εa − εf)=

[DNA]

(εb − εf)+

1

Kb(εb − εf) (1)

where [DNA] is the concentration of DNA in base pairs, εa corresponds to the apparent

molar extinction coefficient Aabs/[M], εf corresponds to the extinction coefficient for the

free organotin [M] and εb corresponds to the extinction coefficient for the fully bound

organotin compound.

3.3 Results and Discussion

3.3.1 Synthesis

Organotin(IV) compounds (4-9) were synthesised using the corresponding

diorganotin(IV) dichloride salts under reflux as shown in Scheme 3.1. The organotin(IV)

compounds are yellow, stable at room temperature and soluble in organic solvents

74

(especially chloroform (CHCl3), DMSO and DMF. The room temperature molar

conductance values of the organotin(IV) compounds in DMSO at a concentration of 10−3

M were in the range of 2.30–8.21 Ω−1·cm2·mol−1, i.e., values well below 25

Ω−1·cm2·mol−1, indicating that these organotin(IV) compounds are non-electrolytes

proving the absence of counter-ions, thus indicating that the dithiocarbazate Schiff bases

(1-3) were covalently bound to the central tin ion.106

SHN

NS

HOO

S NN

SO

OSn

S NN

SO

OSn

+ Ph2SnCl2EtOH

reflux 6 hours

+ Me2SnCl2EtOH and DCM

reflux 6 hours

R

R

R

R = o-CH3 (4)R = p-CH3 (5)R = H (6)

R = o-CH3 (1) R = p-CH3 (2)R = H (3)

R = o-CH3 (7)R = p-CH3 (8)R = H (9)

Scheme 3.1: Synthesis of organotin(IV) compounds (4-9).

3.3.2 IR Spectral Analysis

Infrared spectra of organotin(IV) compounds, 4-9 were measured in the range 4000

to 280 cm−1 and verified using frequencies predicted using DFT. All DFT vibrational

frequencies were scaled using a scaling factor of 0.9682.189 The v(NH) stretching

vibration at ca 3084–3092 cm−1 of the respective Schiff bases, was absent in the

organotin(IV) compounds, indicating that the Schiff bases were deprotonated upon

75

complexation. Sharp v(C=N) bands corresponding to the Schiff base azomethine group

were shifted upon complexation, indicating coordination via the azomethine nitrogen.227

The v(CSS) band splitting provides strong evidence of coordination of the thiolate sulphur

atom to the tin ion. The red shifted of the hydrazinic v(N–N) band in the organotin(IV)

compounds shows evidence of coordination via the azomethine nitrogen due to reduction

in the repulsion between lone pairs of electrons on the nitrogen atom.92,116,228 The

calculated frequencies correlate well with the experimental frequencies observed in the

FT-IR spectra, considering that DFT frequency calculations were performed in the gas-

phase. Only slight differences were observed in the experimental and calculated data for

v(C=N), v(N–N) and v(C=S). Complete IR data are provided in Appendix, Table A3.1.

3.3.3 NMR Spectroscopic Analysis

Complete 119Sn NMR data are provided in Experimental Section 3.2.3. The NMR

details for the organotin(IV) compounds are tabulated in Appendix, Tables A3.2 and

A3.3. NMR spectra of the compounds 4-9 were recorded in CDCl3. The data obtained

were in good agreement with the expected structures in solvents. For all compounds, the

1H NMR signal for the azomethine NH proton of Schiff bases (discussed in Chapter 2)

was absent due to the coordination with the metal ion, corroborating the absence of the

v(NH) FTIR band noted in §3.3.2. A singlet appeared in the downfield region of the 1H

NMR spectra of the Schiff bases, which corresponded to a sp2-C–OH proton. Proton

chemical shifts for both NH and OH disappeared in the spectra of the organotin(IV)

compounds suggesting double deprotonation of the Schiff bases and supporting

coordination of both the phenolic-oxygen and -nitrogen atoms of the Schiff base to the

tin ion.

76

The 13C NMR spectra of the dithiocarbazate Schiff bases showed a downfield

chemical shift at ~196 ppm, indicating the existence of the C=S group and thus the

predominance of the thione tautomer in solution. This signal was shifted upfield in the

organotin(IV) compounds (170–180 ppm), indicating a decrease in electron density at the

C=S carbon atom when the sulphur atom was complexed to the tin ion, consistent with

the structures proposed for other organotin(IV) compounds.229

The 119Sn NMR spectra of the organotin(IV) compounds were recorded at room

temperature in CDCl3. Theoretically, as the tin coordination number increases, chemical

shifts in 119Sn NMR move towards lower frequencies.230 119Sn NMR spectra are also very

sensitive to the type of donor atom (e.g., alkyl or aryl groups in the case of a C donor)

and the coordination environment (e.g., the bond angles around the tin center).231

However, shielding or deshielding of the tin nucleus does not have a significant effect on

the chemical shifts.232 Dimethyltin(IV) compounds showed less negative 119Sn chemical

shifts compared to diphenyltin(IV) compounds. The chemical shifts obtained were within

the range reported for other diphenyl- or dimethyl(IV) compounds with heterocyclic

ligands having penta-coordinated trigonal bipyramidal geometry.230,233

3.3.4 X-ray Crystallography

The crystal structures of 7, 8 and 9 each featured two independent molecules in the

crystallographic asymmetric unit cell, i.e., molecules a and b. Figure 3.2a–c shows the

molecular structures for the first independent molecule in each crystal and images for the

second independent molecules are shown in Figures 3.3a, 3.4a and 3.5a; selected

geometric parameters are collated in Table 3.2. The major conformational difference

between the molecules in each of 7, 8 and 9 is highlighted in the overlay diagram shown

in Figure 3.2d. Whereas in the tolyl species, the tolyl substituent is orientated away from

77

the coordinated S1 atom, i.e., can be considered anti to S1, in each molecule of 8, the

phenyl ring is syn to the S1 atom.

The tin atom in 7 is coordinated by the thiolate-S1, phenoxide-O1 and imine-N

atoms derived from the dinegative, [(3-methoxy-2-oxyphenyl)methyl] dithiocarbazate

ligand. The five-coordinate geometry is completed by a carbon atom from each of the two

methyl groups. The assignment of thiolate-S1 is readily made by comparing the length of

the C1‒S1 bond in 7, i.e., 1.726(9) and 1.717(8) Å, for the two independent molecules

comprising the asymmetric unit, with that in the uncoordinated Schiff base (which lacks

the 2-methyl group in the ester group) of 1.670(2) Å,234 i.e., the bond lengths in 7 are

significantly longer. The C1‒S2 bond lengths are experimentally equivalent at 1.744(8)

and 1.747(8) Å.234 However, the two C1=N1 imine bond lengths in 7 are 1.304(11) and

1.273(10) Å, which are significantly longer than the equivalent C1‒N1(H) bond of

1.338(2) Å in the respective free Schiff base. The coordination mode of the

dithiocarbazate ligand leads to the formation of five- and six-membered chelate rings,

with the chelate angle in the former being approximately 5° more acute than that in the

latter, see Table 3.2. The Sn, S1, N1, N2, C1 chelate ring in molecule is non-planar and

is twisted about the Sn‒S1 bond. On the other hand, the Sn, O1, N1, C10–C12 chelate

ring is best described as having an envelope conformation whereby the Sn atom lies

0.528(11) Å above the plane. However, the structure of the second independent molecule

in the asymmetric unit, molecule b, is slightly different, whereby the five-membered ring

is an envelope with the Sn atom lying 0.372(12) Å above the plane defined by the

remaining four atoms in this ring [RMSD = 0.008 Å]. The six-membered ring is also an

envelope with the Sn atom lying 0.528(11) Å out of the plane defined by the five

remaining atoms which present a RMSD = 0.022 Å. The dihedral angle between the best

planes through the chelate rings in the molecule a is 11.12(17)° while for molecule b, the

78

equivalent angle is 12.4(3)°. The benzyl ring is almost orthogonal to the five-membered

chelate ring for both independent molecules in the unit cell, forming dihedral angles of

88.3(3) and 86.8(3)°, respectively. Similarly, the dihedral angles between the outer rings

are 72.3(3) and 75.5(3)°, respectively, also indicating an approximate orthogonal

relationship. The similarity between the conformations found for the two independent

molecules in the asymmetric unit is highlighted in the overlay diagram shown in Figure

3.3. The five-coordinate geometry defined by the C2NOS donor set for molecule a, with

five-coordinate geometry descriptor τ = 0.48, is intermediate between ideal square-

pyramidal (τ = 0.0) and trigonal-bipyramidal geometries (τ = 1.0);235 the value of τ is 0.50

for molecule b. For each independent molecule, the widest angle corresponds to the S1‒

Sn‒O1 angle with the second widest angle being subtended by the tin-bound methyl

groups. The angles about the quaternary-C1 atom span nearly 20° with the widest angles

involving the doubly bonded N1 atom. Of these, the angle involving the S1 atom, being

the widest, is consistent with some double bond character in the C1‒S1 bond.

Concomitantly, the C1‒S1 bond lengths are systematically shorter than the C1‒S2 bonds

(see Table 3.2).

The isomeric compound, 8, having an ester 4-tolyl residue rather than 2-tolyl,

presents very similar characteristics to those exhibited in the structure of 7 (Figure 3.2b,

Figure 3.4 and Table 3.3). The conformations of the two chelate rings in the independent

molecules are approximately the same, and are based on an envelope conformation with

the tin atom being the flap in all cases. The envelope is more pronounced however, cf. 7,

in the five-membered rings with the tin atom lying 0.486(4) Å [0.490(4) Å for molecule

b] above the plane of the four remaining atoms which have a RMSD = 0.009 Å [0.012 Å]

but, is somewhat flattened in the six-membered rings with the tin atom lying 0.187(4) Å

[0.132(4) Å] out of the plane defined by the five remaining atoms with a RMSD = 0.013

79

Å [0.017 Å]. The dihedral angle between the best planes through the chelate rings is

10.24(8)° [8.53(11)°]. Near orthogonal relationships between the five-membered chelate

rings and the pendant tolyl rings persist, and the dihedral angles between the outer phenyl

rings is 65.75(10)° [78.30(10)°]. The value of τ is 0.54 [0.41]. A difference is noted in the

coordination geometries in that the second widest angles at the tin atom, after S1‒Sn‒O1,

are subtended by the N2 and C19 atoms, i.e., 125.83(12)° [131.09(12)°], rather than the

C18‒Sn‒C19 angle, Table 3.2.

Figure 3.2: Molecular structures of the first independent molecules of (a) 7, (b) 8 and (c) 9 showing atom labelling schemes. (d) Overlay diagram of the two independent molecules of each of 7, a (red image) and b (green), 8, a (blue) and b (pink), and 9, a (yellow) and b (aqua) with the SnMe2 planes superimposed.

80

Figure 3.3: Crystallographic diagrams for 7: (a) Molecular structure of the second independent molecule, molecule b, (b) overlay diagram of molecules a (red image) and inverted-b (green) drawn so the SnC2 atoms of the Me2Sn moiety are overlapped and (c) a view of the supramolecular layer in the ab-plane (left image) and a view of the unit cell contents in projection down the a-axis with one layer highlighted in space-filling mode (right image). The C‒H…O and C‒H…π interactions are shown as orange and purple dashed lines, respectively.

Figure 3.4: Crystallographic diagrams for 8: (a) Molecular structure of the second independent molecule, molecule b, (b) overlay diagram of molecules a (blue image) and inverted-b (pink) drawn so the SnC2 atoms of the Me2Sn moiety are overlapped and (c) a view of the supramolecular dimer sustained by C‒H…π (chelate) interactions (left image; non-participating hydrogen atoms have been omitted) and a view of the unit cell contents in projection down the a-axis. The C‒H…O and C‒H…π interactions are shown as orange and purple dashed lines, respectively.

81

The unsubstituted analogue of the aforementioned isomeric structures, i.e., 9, is

similar to those just described (Figures 3.2c and 3.5, Table 3.2). A likely trend in the Sn‒

S1 bond lengths may be discerned, Table 3.2, in that these are longest in 9 compared to

the tolyl derivatives, 7 and 8, pointing to an electronic influence, i.e., activation of the

aromatic ring owing to the presence of electron-donating methyl groups; while there is

evidence for shorter Sn‒S1 bond lengths in 8 cf. 7, the standard uncertainty values in the

latter preclude a definitive statement on this although the trend for shorter Sn‒S1 bonds

in 8 is consistent with expectation. The conformational flexibility in the chelate ligands

formed by these tridentate ligands is reflected in the flattened envelope conformations for

the five-membered rings with the Sn atom lying 0.326(6) Å above the plane defined by

the four remaining atoms [RMSD = 0.021 Å; comparable parameters for molecule b are

0.287(6) and 0.024 Å, respectively]. By contrast, pronounced envelope conformations are

found for the six-membered rings [0.675(5) and 0.043 Å for molecule a; 0.706(5) and

0.048 Å for molecule b]. The dihedral angle between the best planes through the chelate

rings is 13.66(16)° [13.92(17)°]. The values of τ for the independent molecules compute

to 0.47 and 0.44, respectively. As for 8, after the S1‒Sn‒O1 angles, the widest angle

subtended at the tin atom in 9 for the second independent molecule involves the

coordinated imine-N2 and a tin-bound methyl-C17 atoms, i.e., 127.66(17)°. Of the three

new structures reported herein, the greatest agreement between the two independent

molecules comprising the asymmetric unit is found for 8, see Figure 3.5 and Table 3.2.

82

Table 3.2: Selected geometric parameters (Å, °) for 7, 8 and 9.

Parameter 7 8 9

Bond lengths molecule a molecule b B3LYP

molecule a molecule b B3LYP

molecule a molecule b B3LYP

LanL2DZ/6-311G(d,p)

LanL2DZ/6-311G(d,p)

LanL2DZ/6-311G(d,p)

Sn‒S1 2.545(2) 2.542(2) 2.576 2.5391(8) 2.5309(7) 2.566 2.5789(11) 2.5607(10) 2.590 Sn‒O1 2.092(6) 2.102(6) 2.085 2.090(2) 2.214(2) 2.078 2.103(3) 2.093(3) 2.078 Sn‒N2 2.188(6) 2.192(6) 2.285 2.223(2) 2.088(2) 2.232 2.194(3) 2.205(3) 2.226 C1‒S1 1.726(9) 1.717(8) 1.751 1.736(3) 1.734(3) 1.753 1.725(4) 1.726(4) 1.748 C1‒S2 1.744(8) 1.747(8) 1.772 1.752(3) 1.757(3) 1.771 1.771(4) 1.763(4) 1.783 N1‒N2 1.394(9) 1.406(9) 1.389 1.402(3) 1.409(3) 1.388 1.397(5) 1.395(4) 1.383 C1‒N1 1.304(11) 1.273(10) 1.293 1.289(4) 1.287(4) 1.293 1.298(5) 1.291(5) 1.296

C(x)‒N2 1.308(9) 1.306(9) 1.313 1.309(4) 1.307(4) 1.312 1.311(5) 1.304(5) 1.307

Bond angles

S1‒Sn‒O1 157.89(17) 158.25(17) 159.5 158.51(7) 155.88(7) 159.7 155.82(9) 154.24(10) 159.8 S1‒Sn‒N2 77.35(18) 76.94(18) 77.6 77.33(7) 77.78(6) 77.8 76.70(9) 77.13(8) 77.3 O1‒Sn‒N2 82.2(2) 82.3(2) 82.0 82.13(9) 82.15(8) 82.2 82.76(11) 82.09(12) 82.4

C18‒Sn‒C(y) 128.9(5) 128.2(5) 123.9 122.58(14) 123.05(14) 123.9 127.70(18) 124.8(2) 123.7 S1‒C1‒S2 112.1(5) 111.9(5) 113.4 111.46(17) 111.81(17) 113.4 120.2(2) 120.1(2) 120.8 S1‒C1‒N1 129.1(6) 128.9(6) 128.2 129.1(2) 129.6(2) 128.2 128.7(3) 128.3(3) 127.7 S2‒C1‒N1 118.9(7) 119.3(6) 118.4 119.5(2) 118.6(2) 118.4 111.1(3) 111.6(3) 111.5

83

In the absence of conventional hydrogen bonding interactions, the crystals of 7, 8

and 9 feature a range of weak non-covalent interactions. Packing diagrams are given in

Figures 3.3c, 3.4c and 3.5c and geometric details characterising the specified

intermolecular interactions are given in Appendix, Tables A3.4, A3.5 and A3.6. In the

molecular packing of 7, imine-C‒H…O(phenoxide), methylene-C‒

H…π(methoxybenzene), involving both methylene groups, and tin-bound-methyl-C‒H…π

(tolyl) interactions combine to stabilise supramolecular layers in the ab-plane; layers

stack without directional interactions between them. The crystal of 8 features additional

recognisable points of contact between molecules largely due to the participation of the

methoxy substituents. Thus, tolyl-C‒H…O (methoxy) and methoxy-methyl-C‒H…O

(methoxy, phenoxide) interactions are apparent as are tolyl-C5‒H…π(chelate) and tin-

bound-methyl-C‒H…π(tolyl) points of contact. These consolidate the three-dimensional

architecture, see Figure 3.3c. The presence of C‒H…π (chelate) interactions is of interest

and have only relatively recently been appreciated as being important in contributing to

the stability of coordination compounds.236–238 These interactions usually stabilise

(centrosymmetric) dimeric aggregates237 and that is the case in 8, see Figure 3.4c. In the

crystal of 9, both phenyl-C-H…π (methoxybenzene) and less common

π(chelate)…π(methoxybenzene) interactions are formed. The latter interactions have been

the subject of a recent comprehensive review.239 In the present case, Sn1-containing

molecules interact in this mode via association between the five-membered

(Sn/S1/N1/N2/C1) chelate ring and a methoxybenzene ring. However, these interactions

are best described as edge-to-edge interactions with the closest contact occurring between

a nitrogen atom of the chelate ring and the benzene-carbon, see Figure 3.5c. Similar edge-

to-edge contacts are found for the second independent molecule with the resulting dimeric

aggregates being connected into a supramolecular layer in the ab-plane by phenyl-C-H…π

84

(methoxybenzene) interactions. Globally, the packing comprises alternating layers of

molecules of a (with no obvious points of contact between the dimeric units) and

molecules of b along the c-axis with no directional interactions between successive layers.

Figure 3.5: Crystallographic diagrams for 9: (a) Molecular structure of the second independent molecule, molecule b, (b) overlay diagram of molecules a (yellow image) and b (aqua) drawn so the five-membered rings are overlapped and (c) supramolecular dimer sustained by edge-to-edge chelate ring…benzene interactions (upper image) between Sn1-containing molecules, supramolecular layer sustained by edge-to-edge π (chelate ring)…π (ethoxybenzene) and phenyl-C‒H…π (methoxybenzene) interactions occurring between Sn2-containing molecules and a view of the unit cell contents in projection down the b-axis. The edge-to-edge π (chelate ring)…π (ethoxybenzene) and C‒H…π interactions are shown as blue and purple dashed lines, respectively.

B3LYP-derived bond lengths and angles of 7, 8 and 9 are in excellent agreement

with the values obtained from the crystallographic analysis above (see Table 3.2),

consistent with the absence of strong directional intermolecular interactions in their

85

crystal. For instance, the deviations in the selected bond lengths/angles of 7, 8 and 9

ranged from 0.002–0.144 Å/0–5.6°, whereby the maximum bond length deviation

corresponded to the Sn‒N2 bond for 8 and the maximum bond angle deviation

corresponded to the S1‒Sn‒O1 angle for 9.

3.3.5 UV-vis Absorption Spectral Analysis

The experimental UV-vis spectra of the organotin(IV) measured in DMSO

exhibited prominent absorption bands at 306-315, 364-373 and 433-447 nm. The

corresponding peaks were predicted at 304-320, 357-366 and 423-426 nm using B3LYP.

The highest occupied molecular orbital and lowest unoccupied molecular orbital

(HOMO-LUMO) electron density distribution of the organotin(IV) compounds (see

Figure 3.6) were comparable to those of the corresponding Schiff bases (1-3), i.e., the

HOMO was largely centered on the backbone of the Schiff bases and the LUMO mostly

covered the 2-hydroxy-3-methoxyphenyl ring as well as the backbone of the Schiff base

thus explaining the observed n→π* and π→π* transitions at 306–315 and 364–373 nm,

respectively. The additional, strong S→SnIV intra-ligand band at 433-447 nm indicated

coordination. The presence of the S→SnIV LMCT band was corroborated by IR evidence

which showed that the tin center was coordinated to the sulphur atom in the organotin(IV)

compounds.93 The HOMO and LUMO transitions were the same for all the organotin(IV)

compounds due to their similar structures, the difference being only the presence and

position of the methyl group attached to the phenyl ring. Complete experimental and

calculated UV-vis data are listed in Table A3.7.

86

Figure 3.6: HOMO-LUMO of (a) 4, (b) 5, (c) 6, (d) 7, (e) 8 and (f) 3.3.6 Biological Assays

3.3.6.1 Cytotoxicity

Compounds 4-9 were screened for in vitro cytotoxicity against a panel of 12 cancer

cell lines: EJ-28, RT-112, HT29, U87, SJ-G2, MCF-7, A2780, H460, A431, Du145, BE2-

C, MIA and one normal breast cell line (MCF-10A) using the MTT metabolic assay. After

72 h of incubation, a dose-dependent proliferative effect towards the cancer cell lines was

measured at very low micromolar (µM) concentrations of the compounds. The growth

inhibition concentration of the organotin(IV) compounds required to inhibit 50% cell

proliferation relative to the cells that were treated with DMSO (control) are given in Table

3.3. The stability of these compounds was assessed in DMSO as well as in a mixture of

87

DMSO-H2O. The absorbance obtained from UV-vis spectroscopy analysis was monitored

for 72 h, and an unchanged pattern in the spectra indicated that the compounds were stable

in both the solvent systems tested.

It was also observed that the coordination of the Schiff bases with diphenyltin(IV)

induced potent selective growth inhibition in A2780, BE2-C and MIA cells, in which the

compounds exhibited potency higher than their respective Schiff bases, discussed in

Chapter 2. Cytotoxic selectivity was also observed for 5 and 6 against MCF-7 with GI50

values of 0.82 ± 0.14 and 0.90 ± 0.16 µM, respectively. This constituted a potency that

was almost three times higher than that of the corresponding Schiff bases, 2 and 3 which

had values of 2.5 ± 0.12 and 3.5 ± 0.00 µM, respectively. In addition, 5 exhibited a much

higher potency (GI50 value of 0.87 ± 0.22 µM) than 2 (GI50 value of 8.0 ± 3.5 µM) against

HT-29 cancer cells. A 3-fold greater activity against SJ-G2 was observed for 4 and 6 with

GI50 values of 0.84 ± 0.09 µM and 0.74 ± 0.15, respectively, than their respective Schiff

bases. It is noted that the diphenyltin(IV) compounds showed high selectivity against EJ-

28, RT-112, MCF-7, A2780, BE2-C, SJ-G2 and MIA cell lines in the range 1.4-5.0-fold

as compared to MCF10A.

In contrast, dimethyltin(IV) compounds showed almost similar cytotoxicity values

to their Schiff bases when assayed against the panel of cancer cell lines suggesting that

the presence of two phenyl groups coordinated directly to the tin ion improves the potency

of the Schiff bases. Electron delocalisation over the chelate rings, which increased the

lipophilicity of the molecules and favoured their permeation through the lipid layer of the

cancer cell membranes could also be a contributing factor to the enhancement of

cytotoxicity.240 A third contributing factor could be the ease of ability of the phenyl planes

to intercalate into the base pairs of DNA in the cancer cell lines.53 To better understand

the mode of biological action of the compounds that showed promising activity, apoptosis

88

assays using the minimally invasive bladder cancer RT-112 cells and DNA binding

studies were carried out to evaluate the compounds’ mechanisms of action.

3.3.6.2 Annexin V Assays of 4, 5 and 6-Treated RT-112 Cells

Cell death usually occurs by apoptosis or necrosis, the latter induces severe

inflammation. The apoptotic cells containing apoptotic bodies (small membrane-bound

vesicles), are engulfed by macrophages, and hence no inflammatory response is triggered.

On this account, as apoptotic cell induction is considered crucial in cytotoxic drug

development, investigations using the Annexin V assay were performed to study the

apoptosis effect on cancer cells of the organotin(IV) compounds. The three most highly

cytotoxic diphenyltin(IV) compounds at their specific GI50 concentrations were chosen

for this assay. The analysis was carried out using RT-112 minimum-invasive bladder

cancer cells with high levels of epidermal growth factor receptor expression. In the early

stages of apoptosis, changes occurred at the cell surface.241 The plasma membrane

alteration occurs due to the translocation of phosphatidyl-serine (PS) from the inner part

of the plasma membrane where it is normally located on the cytoplasmic surface, to the

outer layer. The PS exposed on the cell surface has high binding affinity to the Annexin

V. In the later stages of apoptosis, as well as necrosis, the cell membranes lose their

membrane integrity and allow propidium iodide (PI) to enter into the nucleus of the cells.

Annexin V/PI double staining was used to differentiate viable cells (Annexin V- and PI-)

in early apoptosis (Annexin V+ and PI−) or late apoptosis (Annexin V+ and PI+) and

necrosis (annexin V− and PI+) stages.242 Under fluorescence microscopy, the cells stained

by annexin V appear green, whereas the cells stained by PI appear red. After 24 h

exposure to the diphenyltin(IV) compounds, Figure 3.7 show that the cells undergoing

89

late apoptosis and necrosis features. Thus, this assay proved that the diphenyltin(IV)

compounds triggered the activation of apoptosis.243

3.3.6.3 Measurement of Reactive Oxygen Species (ROS) in RT-112 Cells Treated

with 4 and 5

Two compounds with the most promising bioactivity, 4 and 5 , were selected for

further cell apoptosis studies, in order to evaluate the primary damage to biological

macromolecules through redox reactions, particularly via formation of ROS.244 ROS are

highly reactive molecules/radicals (O2, O2•, H2O2, HO•) that are continuously generated

during aerobic metabolism, but when present in excessive amounts they can lead to

protein and DNA oxidation, protein cross-linking and cell death. Estimation of the levels

of ROS in cell culture is an important step towards better understanding the mechanisms

that lead to cell death or disease processes. Figure 3.8 shows that both 4 and 5 stimulated

ROS generation. ROS production was measured using the ROS-detecting fluorescent dye

DCFH-DA and DMSO was used as a control. Both compounds were able to induce ROS

production in RT-112 cell lines after 24 h. The significant increase in ROS generation

supports the findings that the compounds have high potential cytotoxicity and suggested

that the mechanism of cell death was via the simulation of mitochondrial-initiated

events.245

90

Table 3.3: Summary of the in vitro cytotoxicity of the Schiff bases and their organotin(IV) compounds in several cell lines, determined by MTT assay and expressed as GI50 values with standard errors. (GI50 is the concentration at which cell growth is inhibited by 50% over 72 h).

Compounds Growth Inhibition Concentration, GI50 (µM)

EJ-28 RT-112 HT29 U87 MCF-7 A2780 H460 A431 Du145 BE2-C SJ-G2 MIA MCF10A

4 0.35 ± 0.05

0.31 ± 0.05

1.8 ± 0.1

4.4 ± 1.7

1.4 ± 0.12

0.23 ± 0.01

2.8 ± 0.00

2.1 ± 0.1

1.7 ± 0.07

0.47 ± 0.05

0.84 ± 0.09

0.34 ± 0.02

1.2 ± 0.3

5 3.1 ± 0.4

1.7 ± 0.5

0.87 ± 0.22

1.3 ± 0.4

0.82 ± 0.14

0.35 ± 0.07

1.9 ± 0.23

1.4 ± 0.4

1.5 ± 0.31

0.53 ± 0.01

1.0 ± 0.2

0.39 ± 0.03

1.1 ± 0.4

6 0.32 ± 0.05

0.58 ± 0.29

1.2 ± 0.2

1.1 ± 0.4

0.90 ± 0.16

0.23 ± 0.03

2.2 ± 0.00

1.6 ± 0.1

1.4 ± 0.33

0.50 ± 0.05

0.74 ± 0.15

0.43 ± 0.04

1.3 ± 0.2

7 2.5 ± 0.8

2.4 ± 0.8

2.7 ± 0.1

3.6 ± 0.2

3.1 ± 0.12

3.1 ± 0.1

3.3 ± 0.13

3.0 ± 0.0

3.9 ± 0.03

2.9 ± 0.0

2.5 ± 0.3

3.9 ± 0.1

3.0 ± 0.1

8 >5.0 >5.0 6.9 ± 3.5

3.3 ± 0.2

3.3 ± 0.32

3.1 ± 0.6

4.7 ± 0.46

4.2 ± 0.3

4.2 ± 0.10

3.3 ± 0.2

3.1 ± 0.6

8.9 ± 1.3

3.3 ± 0.2

9 3.4 ± 0.3

2.4 ± 0.3

2.8 ± 0.2

4.5 ± 0.2

2.5 ± 0.21

3.2 ± 0.0

3.8 ± 0.10

3.2 ± 0.0

3.2 ± 0.13

2.7 ± 0.1

2.1 ± 0.1

5.0 ± 0.1

2.8 ± 0.4

Cisplatin 3.97 ± 0.48

3.56 ± 0.88

11.0 ± 2.0

4.0 ± 1.0

6.5 ± 0.8

1.0 ± 0.1

0.9 ± 0.2

2.4 ± 0.3

1.2 ± 0.1

1.9 ± 0.2

0.4 ± 0.1

8.0 ± 1.0 nd

GI50 (μM): (the colors indicate) = 0.1–0.99 (strong); =1.0–9.9 (moderate); =10-100 (weak); = nd (not determined).

91

Figure 3.7: Apoptosis detection through fluorescence microscopy. Cells were treated for 24 h with 4 (0.31 µM), 5 (1.66 µM) and 6 (0.58 µM), and the negative control (DMSO) in complete media. After staining with Annexin V and PI, necrotic and apoptotic cells were detected by fluorescence microscopy (20×).

92

Figure 3.8: Percent reactive oxygen species (ROS) production in RT-112 cells treated with (a) 4 (0.31 μM) (b) 5 (1.66 μM) for 24 h and stained with 1 mM DCFH-DA for 60 minutes at 37 °C. DMSO and H2O2 acted as negative and positive controls, respectively.

3.3.6.4 DNA Interaction Studies

Binding of the synthesised organotin(IV) compounds with DNA were investigated

both theoretically and experimentally. The interaction of organotin(IV) compounds with

CT-DNA was elucidated via UV-vis absorption spectroscopy. The electronic spectra of

the organotin(IV) compounds exhibited three absorption bands around 306–315 nm, 350–

377 nm and 433–450 nm, which corresponded to intra-ligand transitions and ligand metal

charge transfer (LMCT) transitions. The absorption spectra of diphenyltin(IV)

compounds at a constant concentration (50 μM) in the presence of different

concentrations of CT-DNA are shown in Figure 3.9. The DNA binding spectra for

dimethyltin(IV) compounds are presented in Appendix Figure A3.1. Upon increasing

concentration of CT-DNA, hypochromism and bathochromism occurred at around 350-

377 nm, indicating strong π–π stacking interactions between the aromatic chromophore

of the organotin(IV) compounds and the nitrogeneous pair of the DNA strand.206,246 The

same hypochromic and bathochromic shifts were also observed in other transitions. In

93

order to compare the binding strength of the organotin(IV) compounds to CT-DNA, the

intrinsic binding constants (Kb) were quantitatively determined and are presented in Table

3.4. The Kb values show that these compounds have strong binding affinities with DNA,

and potentially prevent enzyme binding to the nitrogenous bases of DNA.

The Gibbs free energies of interaction between the organotin(IV) compounds and

DNA ranged from −31.0 to −37.6 kJ mol−1, meaning that the interaction of the

organotin(IV) compounds with DNA was spontaneous.247 To investigate and understand

the interactions between DNA and these compounds, molecular docking simulations were

carried out.

3.3.7 Molecular Docking Studies

Molecular docking analyses of the organotin(IV) compounds with the DNA duplex

of sequence d(CGCGAATTCGCG)2 dodecamer (PDB ID: 1BNA) were performed and

results are detailed in Table 3.5. These simulations demonstrated that the organotin(IV)

compounds considered here were commensurate with the dimensions of the grooves of

the DNA duplex sequence. Figure 3.10(a) revealed that these small organotin(IV)

compounds fitted well into the A-T rich region of the minor grooves of DNA. Generally,

the electronegativity of A-T sequences provided better van der Waals interactions due to

their narrower sequences compared to G-C regions. In groove regions, van der Waals and

hydrophobic interactions play important roles in the stabilisation of compounds.248 Their

interactions therefore, involved hydrogen bonding, hydrophobic and electrostatic

interactions yielding binding energies between −9.10 to −10.07 kcal mol−1, which is better

than cisplatin (−8.65 kcal mol−1).155 The interactions occurred either through the

phosphate-oxygen atom of the DNA or via DNA bases. 4 exhibited two hydrogen bonding

interactions through methoxy-oxygen atoms and methyl groups with phosphate-oxygen

94

atoms of DT20 and DT8, respectively. Similar interactions were observed for 7 through

the oxygen atom of methoxy group with the phosphate backbone atoms of DT19; 5, 8 and

9 interacted through the carbon atoms of the azomethine group with oxygen atoms of

thymine base, DT8, DT7 and DT20, respectively. Detailed interactions between the

organotin(IV) compounds and DNA residues are presented in Figure 3.10(b). Thus, on

the basis of docking studies, it could be inferred that the cytotoxic activities of

organotin(IV) compounds could have arisen due to effective interactions of the

compounds with DNA.

Table 3.4: Binding constants (Kb), hypochromism (%), bathochromic shifts (nm) and Gibbs free energy (kJ mol−1) values for the interaction of organotin(IV) compounds with calf thymus DNA (CT-DNA).

Compounds Binding Constant

(Kb)

Hypochromism, (%)

Bathochromism Shift (nm)

Gibbs Free

Energy (kJ mol−1)

4 6.21 × 105 41.06 6 −33.1

5 2.67 × 105 28.25 4 −31.0

6 1.21 × 106 70.72 12 −34.7

7 3.46 × 105 37.97 10 −31.6

8 4.15 × 105 43.72 7 −32.1

9 3.91 × 106 53.26 1 −37.6

95

Figure 3.9: (a) Electronic absorption spectra of (i) 4, (ii) 5 and (iii) 6; (b) Plot of [DNA]/εa − εf vs. [DNA] for absorption titration of DNA with (i) 4, (ii) 5 and (iii) 6. The arrow indicates the change in absorbance in tandem with increasing DNA concentration.

96

Table 3.5: Molecular docking data for the organotin(IV) compounds with B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2.

Compound

Final Intermolecular Energy, kcal mol−1 Final total Internal

Energy (2), kcal mol−1

Torsional Free Energy

(3), kcal mol−1

Unbound System’s Energy

[=(2)] (4), kcal mol−1

Estimated Free Energy of

Binding [(1) + (2) + (3) − (4)], kcal

mol−1

Estimated Free

Energy of Binding, kJ

mol−1

vdW + Hbond + desolv. Energy

Electrostatic Energy

Total (1)

4 −11.46 0.00 −11.47 −2.19 1.79 −2.19 −9.68 −40.50 5 −11.87 0.01 −11.86 −1.82 1.79 −1.82 −10.07 −42.13 6 −11.07 0.03 −11.04 −1.86 1.79 −1.86 −9.25 −38.70 7 −10.27 −0.03 −10.30 0.57 1.19 −0.57 −9.10 −38.07 8 −10.84 0.00 −10.83 −0.53 1.19 −0.53 −9.64 −40.33 9 −10.58 −0.01 −10.59 −0.54 1.19 −0.54 −9.40 −39.33

97

Figure 3.10: (a) Schematic representation of organotin(IV) compounds that fit well in the grooves of the DNA strand obtained by docking simulations. The two double-stranded DNA comprise of the phosphate deoxyribose backbone with guanine (DG, red), cytosine (DC, blue), adenine (DA, pink) and thymine (DT, orange). (b) Molecular interactions of organotin(IV) compounds within the grooves of double stranded DNA residues.

98

3.4 Conclusions

This chapter presented six organotin(IV) compounds derived from o-vanillin

dithiocarbazate Schiff bases (discussed in Chapter 2) characterised by elemental analysis

and various spectroscopic techniques (UV-visible, FTIR, 1H, 13C and 119Sn NMR). The

molecular structures of 7, 8 and 9 were elucidated by single crystal X-ray structure

analysis which indicated that the compounds had five-coordinate C2NOS geometries,

derived from thiolate-S1, phenoxide-O1 and imine-N atoms of a dinegative, tridentate

dithiocarbazate ligand along with two methyl-carbon atoms of the methyl groups, which

were intermediate between ideal trigonal-bipyramidal and square-pyramidal geometries.

The in vitro cytotoxicity against a panel of cancer cell lines viz., EJ-28, RT-112, HT29,

U87, SJ-G2, MCF-7, A2780, H460, A431, Du145, BE2-C and MIA cancer cell lines

revealed the remarkable cytotoxicity of the compounds 4, 5 and 6 against most of these

cell lines, which were more potent than cisplatin. In order to evaluate the mechanism of

death of the organotin(IV) compounds, Annexin-V and ROS assays were performed using

the RT-112 cell line, where the diphenyltin(IV) compounds were able to induce apoptosis

and ROS generation in the cells. The DNA binding of the organotin(IV) compounds were

investigated using UV-vis absorption and the diphenyltin(IV) compounds bound

reasonably well into the grooves of DNA with binding constants in the range 105–106

M−1. The mode of interaction of the organotin(IV) compounds with DNA was further

validated by molecular docking studies which were in accordance with the experimental

results, suggesting the strong binding of diphenyltin(IV) at the AT-rich region of the

minor groove of DNA. The observed trend in the cytotoxicity and binding interactions of

organotin(IV) compounds were attributed to the nature of the group bound to the tin metal

center where the compounds 4, 5 and 6 exhibited potent efficacy against cells tested.

99

The biologically active species formed from the cleavage of the tin-ligand bonds is

the major effect of the overall activity. The compounds with stronger Sn-O, Sn-N and Sn-

S bonds hence have difficulty forming could not easily form their active species, such as

[R2Sn(IV)]2+ (R = Ph and Me) species through the dissociation of these bonds. Such active

species reportedly cross the cell membrane by deeply affecting the structural arrangement

of membranes of biological systems.249,250 The moderate stability of the bonds results in

slow hydrolysis and gradual dissociation of the Sn-O, Sn-N and Sn-S bonds. It is

conceivable therefore that the gradual dissociation of these bonds could facilitate delivery

of these active species to their specific active site.251 For instance, the Sn-O (2.092-2.214

Å), Sn-N (2.088-2.223 Å) and Sn-S (2.542-2.5789 Å) bonds in the compounds discussed

in this chapter were shorter than that reported by Saxena and Huber,251 but still possessed

very good activity that is higher than the respective Schiff bases, particularly for the

diphenyltin(IV) compounds. It is concluded therefore that the bioactivity of tin(IV)

compounds cannot solely be attributed to the stability of the ligand-metal bonds.

While excellent cytotoxic activity of three oVa dithiocarbazate Schiff bases and

their diphenyl- and dimethyltin(IV) compounds were discussed in Chapter 2 and 3, it is

hypothesised that this activity could be improved via structural modification such as

introduction of hydroxyl group in compounds. The addition of hydroxyl group in

compounds were proven to enhance binding and activity due to their great ability of

forming hydrogen bond interactions with target macrobiomolecules such as DNA and

enzymes.252–254

This hypothesis will be tested in Chapter 4, which reports structural evaluation and

cytotoxic studies of catechol dithiocarbazate Schiff bases (10, 11, 12) with diphenyl- (13,

14, 15) and dimethyltin(IV) (16, 17, 18). The small modification of the structure was

100

explored by substituting a hydroxyl group at the position of the methoxy group of oVa

dithiocarbazate Schiff bases and their organotin(IV) compounds.

101

CHAPTER 4

SELECTIVE CYTOTOXICITY OF ORGANOTIN(IV) COMPOUNDS CONTAINING CATECHOL DITHIOCARBAZATE SCHIFF BASES

4.1 Introduction

A number of dihydroxybenzyl (catechol) containing drugs are currently available

on the market, such as L-dopa, which is used for treating Parkinson's disease, and

Protocatechualdehyde, which is used to treat human breast cancer255 and inhibit casein

kinase II activity in leukemic cells.256 While the discovery of catechol dithiocarbazate

Schiff base derived from S-methyldithiocarbazate was reported by Fayed et al.,257 the

cytotoxic activity of dithiocarbazate Schiff bases of catechol and their tin(IV) compounds

remains unexplored.

Building on results reported in Chapters 2 and 3, this chapter focuses on the

synthesis, structural characterisation and cytotoxic activity of catechol dithiocarbazate

Schiff bases (10-12) and their organotin(IV) compounds (13-18), see Figure 4.1.

Interactions of the synthesised organotin(IV) compounds with DNA as a function of the

coordination complex structure will also be discussed in the context of their structure

activity relationships.

This chapter contains material from submitted article for publication in Research

on Chemical Intermediates on 18 October 2019.

102

R = o-CH3 (13)R = p-CH3 (14)R = H (15)

R = o-CH3 (10) R = p-CH3 (11)R = H (12)

R = o-CH3 (16)R = p-CH3 (17)R = H (18)

SHN

SN

HOOH

R

S N

SN

OOH

R

Sn

S N

SN

OOH

R

Sn

Figure 4.1: The structures of catechol dithiocarbazate Schiff bases (10-12) and their organotin(IV) compounds (13-18).

4.2 Experimental

4.2.1 Materials and Physical Measurements

All solvents and reagents were of analytical reagent grade and used without further

purification as described in Chapter 2 (§2.2.1) and Chapter 3 (§3.2.1). Chemical: 2,3-

dihydroxybenzaldehyde (Merck, New York, NY, USA). The instrumentations used as

outlined in Chapter 2 (§2.2.3) and Chapter 3 (§3.2.3).

4.2.2 Synthesis

4.2.2.1 S-N-R-benzyldithiocarbazate (N = 2, 3; R = methyl)

The details of the synthesis procedure of S1, S2 and S3 are described in Chapter 2

(§2.2.2.1).

4.2.2.2 Schiff Bases

S1/ S2 (2.12 g, 10 mmol) was dissolved in hot acetonitrile (100 cm3) and S3 (1.98

g, 10 mmol) was dissolved in hot absolute ethanol (100 cm3). 2,3-

103

Dihydroxybenzaldehyde (1.38 g, 10 mmol) in ethanol (50 cm3) was then added to the

solution of S-substituted dithiocarbazate. The mixture was stirred and heated (78 °C) for

about 1 hour and then allowed to cool to room temperature. The pale-yellow Schiff bases

that formed were filtered and washed with absolute ethanol. The products were

recrystallised using absolute ethanol and dried over silica gel.101,110

4.2.2.2.1 S-2-methylbenzyl-β-N-(2,3-dihydroxybenzylmethylene)

dithiocarbazate (10)

Pale yellow solid. Yield: 36%. M.p.: 150-152 °C. Analysis calculated for

C16H16N2O2S2: C, 57.81; H, 4.85; N, 8.43 %. Found: C, 56.98; H, 4.65; N, 8.21. FT-IR

(ATR, cm-1): 3496 (O-H), 3097 (N-H), 1608 (C=N), 1119 (N-N), 1013 (C=S). 1H NMR

(DMSO-d6) δ (ppm): 4.45 (s, 2H, CH2), 6.69-7.39 (multiplet, 7H, Ar-H), 8.51 (s, 1H,

CH), 9.51, 9.55 (s, 2H, OH), 13.41 (s,1H, NH). 13C NMR (DMSO-d6) δ (ppm): 19.3 (Ar-

CH3), 36.9 (CH2), 118.1, 118.5, 119.7, 120.0, 126.6, 128.2, 130.7, 130.8, 134.3, 137.3,

146.2, 146.2 (12 x aromatic-C), 146.5 (C=N), 195.6 (S-C=S). m/z calculated for

C16H16N2O2S2 332.44 g/mol, found 332.05.

4.2.2.2.2 S-4-methylbenzyl-β-N-(2,3-dihydroxybenzylmethylene)

dithiocarbazate (11)

Pale yellow solid. Yield: 68%. M.p.: 203-205 °C. Analysis calculated for

C18H19N3O2S2: C, 57.81; H, 4.85; N, 8.43 %. Found: C, 57.55; H, 4.87; N, 8.60. FT-IR

(ATR, cm-1): 3501 (O-H), 3106 (N-H), 1607 (C=N), 1114 (N-N), 1012 (C=S). 1H NMR

(DMSO-d6) δ (ppm): 4.44 (s, 2H, CH2), 6.69-7.30 (multiplet, 7H, Ar-H), 8.51 (s, 1H,

CH), 9.53, 9.54 (s, 2H, OH), 13.39 (s,1H, NH). 13C NMR (DMSO-d6) δ (ppm): 21.2 (Ar-

CH3), 38.0 (CH2), 118.1, 118.4, 119.8, 119.9, 129.6, 129.7, 133.8, 137.0, 146.1, 146.2

104

(10 x aromatic-C), 146.5 (C=N), 195.8 (S-C=S). m/z calculated for C16H16N2O2S2 332.44

g/mol, found 332.05.

4.2.2.2.3 S-benzyl-β-N-(2,3-dihydroxybenzylmethylene)dithiocarbazate (12)

Pale yellow crystals. Yield: 77%. M.p.: 201-203 °C. Analysis calculated for

C15H14N2O2S2: C, 56.58; H, 4.43; N, 8.80 %. Found: C, 56.56; H, 4.67; N, 8.67. FT-IR

(ATR, cm-1): 3488 (O-H), 3097 (N-H), 1608 (C=N), 1119 (N-N), 1018 (C=S). 1H NMR

(DMSO-d6) δ (ppm): 4.47 (s, 2H, CH2), 6.67-7.39 (multiplet, 8H, Ar-H), 8.50 (s, 1H,

CH), 9.53, 9.56 (s, 2H, OH), 13.39 (s, 1H, NH). 13C NMR (DMSO-d6) δ (ppm): 38.06

(CH2), 118.1, 118.4, 119.7, 119.9, 127.7, 129.0, 129.7, 137.1, 146.1 (9 x aromatic-C);

146.5 (C=N); 195.6 (S-C=S). m/z calculated for C15H14N2O2S2 318.41 g/mol, found

318.05.

4.2.2.3 Diphenyltin(IV) Compounds

Each Schiff base (0.33 g (10 and 11); 0.32 g (12), 1 mmol) was dissolved separately

in absolute ethanol (50 cm3) and mixed with an ethanolic solution (10 cm3) of Ph2SnCl2

(0.34 g, 1 mmol). The yellow solution was refluxed for ca. 6 hours. The mixture was left

stirring overnight and the reduction in the volume (15 cm3) of the reaction mixture

resulted in the deposition of a yellow precipitate that was filtered off and recrystallised

from methanol.

105

4.2.2.3.1 Diphenyltin(IV)[S-2-methylbenzyl-β-N-(2,3-dihydroxybenzyl methylene)

dithiocarbazate] (13)

Yellow solid. Yield: 42%. M.p.: 112-113 °C. FT-IR (ATR, cm-1): 1610 (C=N),

1021 (N-N), 959 (C=S). 1H NMR (CDCl3) δ (ppm): 2.40 (s, 3H, CH3), 4.44 (s, 2H, CH2),

6.40-7.80 (multiplet, 17H, Ar-H), 8.81 (s, 1H, CH). 13C NMR (CDCl3) δ (ppm): 19.3

(CH3), 34.4 (CH2), 114.7, 118.3, 118.5, 125.1, 127.5, 128.6, 129.0, 129.2, 130.4, 135.6,

136.4, 141.8, 147.8, 154.9 (aromatic-C), 166.1 (C=N), 172.3 (S-C-S). 119Sn NMR

(CDCl3) δ (ppm.): -233.7.

4.2.2.3.2. Diphenyltin(IV)[S-4-methylbenzyl-β-N-(2,3-dihydroxybenzyl

methylene)dithiocarbazate] (14)

Light-orange crystals. Yield: 71%. M.p.: 137-138 °C. FT-IR (ATR, cm-1): 1611

(C=N), 1006 (N-N), 956 (C=S). 1H NMR (CDCl3) δ (ppm): 2.34 (s, 3H, CH3), 4.41 (s,

2H, CH2), 6.43-7.81 (multiplet, 17H, Ar-H), 8.80 (s, 1H, CH). 13C NMR (CDCl3) δ (ppm):

21.1 (CH3), 35.9 (CH2), 114.7, 118.2, 118.4, 125.1, 129.0, 129.1, 129.3, 130.4, 133.2,

135.6, 137.2, 141.8, 147.8, 154.9 (aromatic-C), 166.1 (C=N), 172.1 (S-C-S). 119Sn NMR

(CDCl3) δ (ppm.): -233.4.

4.2.2.3.3 Diphenyltin(IV)[S-benzyl-β-N-(2,3-dihydroxybenzylmethylene)

dithiocarbazate] (15)

Light-yellow crystals. Yield: 41%. M.p.: 150-151 °C. FT-IR (ATR, cm-1): 1610

(C=N), 1016 (N-N), 960 (C=S). 1H NMR (CDCl3) δ (ppm): 4.44 (s, 2H, CH2), 6.43-7.80

(multiplet, 18H, Ar-H), 8.76 (s, 1H, CH). 13C NMR (CDCl3) δ (ppm): 36.0 (CH2), 114.7,

118.3, 118.5, 125.1, 127.5, 128.6, 129.0, 129.2, 130.4, 135.6, 136.4, 141.8, 147.8, 154.9

(aromatic-C), 166.2 (C=N), 172.0 (S-C-S). 119Sn NMR (CDCl3) δ (ppm.): -233.4.

106

4.2.2.4 Dimethyltin(IV) Compounds

Each Schiff base (0.33 g (10 and 11); 0.32 g (12), 1 mmol) was dissolved separately

in absolute ethanol (50 cm3) and dichloromethane (20 cm3) and was then mixed with an

ethanolic solution (10 mL) of Me2SnCl2 (0.22 g, 1 mmol). The yellow solution was

refluxed for ca. 6 hours. The mixture was left overnight and kept at room temperature for

a few days, yielding brown crystals that were suitable for single crystal X-ray diffraction

analysis.

4.2.2.4.1 Dimethyltin(IV)[S-2-methylbenzyl-β-N-(2,3-dihydroxybenzyl

methylene)dithiocarbazate] (16)

Yellow solid. Yield: 31%. M.p.: 184-185 °C. FT-IR (ATR, cm-1): 1597 (C=N),

1081 (N-N), 984 (C=S). 1H NMR (CDCl3) δ (ppm): 0.93 (Sn-CH3), 2.42 (s, 3H, CH3),

4.43 (s, 2H, CH2), 6.27-7.34 (multiplet, 7H, Ar-H), 8.79 (s, 1H, CH). 13C NMR (CDCl3)

δ (ppm): 6.9 (Sn-CH3), 19.5 (CH3), 34.6 (CH2), 114.5, 117.9, 119.0, 124.8, 126.1, 126.4,

128.0, 129.9, 130.4, 130.7, 137.2, 148.0 (aromatic-C), 166.1 (C=N), 173.3 (S-C-S). 119Sn

NMR (CDCl3) δ (ppm.): -99.9.

4.2.2.4.2 Dimethyltin(IV)[S-4-methylbenzyl-β-N-(2,3-dihydroxybenzyl

methylene)dithiocarbazate] (17)

Light-yellow crystals. Yield: 66%. M.p.: 180-182 °C. FT-IR (ATR, cm-1): 1613

(C=N), 1006 (N-N), 956 (C=S). 1H NMR (CDCl3) δ (ppm): 0.92 (Sn-CH3), 2.34 (s, 3H,

CH3), 4.37 (s, 2H, CH2), 6.24-7.27 (multiplet, 7H, Ar-H), 8.77 (s, 1H, CH). 13C NMR

(CDCl3) δ (ppm): 6.8 (Sn-CH3), 21.2 (CH3), 36.0 (CH2), 114.4, 117.7, 124.6, 129.1,

129.3, 133.4, 137.1, 147.8, 154.5 (aromatic-C), 166.1 (C=N), 173.3 (S-C-S). 119Sn NMR

(CDCl3) δ (ppm.): -99.8.

107

4.2.2.4.1 Dimethyltin(IV)[S-benzyl-β-N-(2,3-dihydroxybenzylmethylene)

dithiocarbazate] (18)

Light-yellow crystals. Yield: 44%. M.p.: 110-112 °C. FT-IR (ATR, cm-1): 1612

(C=N), 1016 (N-N), 960 (C=S). 1H NMR (CDCl3) δ (ppm): 0.92 (s, 6H, Sn-CH3), 4.41

(s, 2H, CH2), 6.24-7.38 (multiplet, 8H, Ar-H), 8.76 (s, 1H, CH). 13C NMR (CDCl3) δ

(ppm): 6.8 (Sn-CH3), 36.2 (CH2), 114.3, 117.7, 124.6, 127.4, 128.6, 129.2, 136.6, 147.8,

154.2, 154.7 (aromatic-C), 166.2 (C=N), 173.1 (S-C-S). 119Sn NMR (CDCl3) δ (ppm.): -

99.5.

4.2.3 X-ray Crystallographic Analysis

X-ray intensity data for 12, 14, 15, 17 and 18 were measured at T = 150 K on an

Oxford Diffraction Gemini Eos CCD diffractometer equipped with Mo Kα radiation

source (λ = 0.71073 Å). Data reduction, including absorption correction, was

accomplished with CrysAlisPro.258 The structures were solved by direct-methods181 and

refined (anisotropic displacement parameters and C-bound H atoms in the riding model

approximation) on F2.182 When applicable, the O- and N-bound H atoms were refined

with O‒H and N‒H constrained to 0.84 ± 0.01 and 0.88 ± 0.01 Å, respectively, and with

Uiso(H) = 1.5Ueq(O) or 1.2Ueq(N). A weighting scheme of the form w = 1/[σ2(Fo2) + (aP)2

+ bP] where P = (Fo2 + 2Fc2)/3) was introduced in each case. The absolute structures of

17 and 18 were determined based on differences in Friedel pairs included in the respective

data sets. In 14, the maximum and minimum residual electron density peaks of 2.36 and

0.92 eÅ-3 were located 0.96 and 0.62 Å from the Sn1 atom. In 15, the residual peaks of

1.39 and -0.49 eÅ-3 were located 0.87 and 0.79 Å from the C9 and Sn1a atoms,

respectively. The molecular structure diagrams were generated with ORTEP for

Windows183 with 50% displacement ellipsoids, and the packing diagrams were drawn

108

with DIAMOND.184 Additional data analysis was made with PLATON.216 Crystal data

and refinement details are given in Table 4.1.

4.2.4 DFT Calculations and Molecular Docking Studies

DFT were performed using parameters detailed in Chapter 2 (§2.2.6) and Chapter

3 (§3.2.5.1), respectively. The coordinates of 14, 15, 17 and 18 were taken from their

crystal structures as a Crystallographic Information File (CIF) and converted to the

Protein Data Bank (PDB) format, whereas the coordinates for 13 and 16 were obtained

after minimisation of energy using DFT method. The parameters for molecular docking

studies described in Chapter 3 (§3.2.5.2).

4.2.5 MTT Assays

The protocol used for MTT assay followed that outlined in Chapter 2 (§2.2.6).

4.3 Results and Discussion

4.3.1 Synthesis

The general synthesis of the S-substituted Schiff bases (10-12) is outlined in

Scheme 4.1. The desired Schiff bases formed in good yields, except for 10, due to the

presence of an unexpected by-product (3-[(1Z)-2-[bis([(2-methylphenyl)methyl]

sulfanyl)methylidene]hydrazin-1-ylidenemethyl]benzene-1,2-diol) which was

manually isolated from the filtrate of 10.259 The Schiff bases were then synthesised via

condensation, by reacting the three S-substituted dithiocarbazates with 2,3-

dihydroxybenzaldehyde in an alcoholic medium.

109

RS N

H

SN

RS N

H

SNH2

HOOH

R = o-CH3 (10) R = p-CH3 (11) R = H (12)

O

HOOH

Scheme 4.1: Synthetic pathway for the formation of Schiff bases 10, 11 and 12.

Schiff bases containing the thioamide functional group [-NH-C(=S)] are capable of

undergoing thione-thiol tautomerism. They can exist either as thione (Figure 4.2a) or thiol

(Fig. 4.2b) tautomeric forms, or a mixture of both in solution. The three Schiff bases were

then each reacted with diphenyltin(IV) dichloride and dimethyltin(IV) dichloride

separately to produce six new organotin(IV) compounds (Scheme 4.2). The products were

isolated as yellow precipitates. Compounds 12, 14, 15, 17 and 18 produced crystalline

materials which were suitable for single crystal X-ray diffraction analysis. The Schiff

bases and their organotin(IV) compounds were non-hygroscopic and soluble in common

organic solvents, especially DMSO, CHCl3 and DMF. The organotin(IV) compounds

behaved as non-electrolytes in DMSO where the molar conductivity values were in the

range of 2.49-6.07 Ω−1cm2mol−1.106

RS N

H

SN

HOOH

RS N

SHN

HOOH

Where: R=o-CH3 (10) R=p-CH3 (11) R=H (12)

Figure 4.2: (a) Thione and (b) thiol tautomerism of Schiff bases 10-12.

110

Table 4.1: Crystallographic data and refinement details for 12, 14, 15, 17 and 18. Compound 12 14 15 17 18

Formula C15H14N2O2S2 C28H24N2O2S2Sn C27H22N2O2S2Sn C18H20N2O2S2Sn C17H18N2O2S2Sn Molecular weight 318.40 603.30 589.27 479.17 465.14 Crystal size/mm3 0.06 × 0.10 × 0.13 0.10 × 0.12 × 0.15 0.09 × 0.14 × 0.18 0.10 × 0.15 × 0.19 0.18 × 0.21 × 0.25 Colour pale-yellow light-orange light-yellow light-yellow light-yellow Crystal system monoclinic monoclinic triclinic monoclinic monoclinic Space group P21/n P21/c P1 Ia Ia a/Å 5.6635(8) 17.9855(4) 9.8664(5) 11.5124(4) 11.6385(5) b/Å 20.454(2) 27.5260(7) 14.0199(6) 13.0512(5) 12.4681(5) c/Å 12.7577(15) 10.4376(2) 19.7735(8) 12.6896(5) 12.9079(5) α/° 90 90 69.369(4) 90 90 β/° 101.843(12) 97.613(2) 89.747(3) 94.724(5) 98.229(4) γ/° 90 90 71.234(4) 90 90 V/Å3 1446.4(3) 5121.8(2) 2405.3(2) 1900.14(12) 1853.78(13) Z 4 8 4 4 4 Dc/g cm-3 1.462 1.565 1.627 1.675 1.667 μ/mm-1 0.373 1.189 1.264 1.578 1.615 Measured data 7533 27084 20871 4585 4287 θ range/° 3.4 – 25.0 3.5 – 29.5 3.3 – 29.4 3.6 – 29.2 3.5 – 29.3 Unique data 2530 11954 10961 3039 2914 Observed data (I 2.0σ(I)) 1199 9102 8908 2894 2789 No. parameters 199 639 619 235 222 R, obs. data; all data 0.073; 0.170 0.046; 0.069 0.034; 0.047 0.027; 0.029 0.026; 0.027 a; b in weighting scheme 0.034; 0 0.046; 3.970 0.030; 0 0.033; 0 0.028 Rw, obs. data; all data 0.120; 0.170 0.098; 0.110 0.071; 0.078 0.065; 0.067 0.062; 0.063 Range of residual electron density peaks/eÅ-3 -0.40 – 0.43 -0.92 – 2.36 -0.49 – 1.39 -0.43 – 0.38 -0.53 – 0.36

111

SHN

NS

HOOH

S NN

SO

OHSn

S NN

SO

OHSn

EtOH(reflux 6 hour)

EtOH and DCM(reflux 6 hour)

R

R

R

Me2SnCl2

Ph2SnCl2

Where: R = o-CH3 (10) R = p-CH3 (11) R = H (12)

R = o-CH3 (13)R = p-CH3 (14)R = H (15)

R = o-CH3 (16)R = p-CH3 (17)R = H (18)

Scheme 4.2: Synthetic pathway for the synthesis of organotin compounds 13-18.

4.3.2 IR Spectral Analysis

The experimental and calculated frequencies for the Schiff bases (10-12) and their

organotin(IV) compounds (13-18) were recorded in the range 4000 to 280 cm-1 and 4000

to 0 cm-1, respectively. All DFT vibrational frequencies were scaled using a scaling factor

of 0.9682.189 The presence of the Schiff base azomethine C=N group was indicated by

the sharp peak at 1607-1608 cm-1, which shifted to higher frequencies upon complexation.

In a similar vein, the red-shift of the hydrazinic v(N-N) frequencies in the organotin(IV)

compounds provided further evidence of this coordination mode, and was related to the

reduction in the repulsion between lone pairs of electrons on the nitrogen atom.92,116,228

The v(C=S) splitting of the Schiff bases in the range of 1012-1018 cm-1 were shifted to

lower frequencies (956-966 cm-1) in the spectra of organotin(IV) compounds, thus

showing strong evidence to prove that the tin ion coordinated the thiolate sulphur.227 The

112

spectra of the Schiff bases also exhibited v(N-H) at 3097-3196 cm-1 which was not present

in the spectra of the organotin(IV) compounds due to the protonation of the N-H group

upon complexation.95 Furthermore, the hydroxyl group attached at the ortho position of

the benzene ring participated in the coordination to the tin metal centre, which was

consistent with the disappearance of v(OH) upon complexation. The calculated

frequencies correlated well with the experimental frequencies, considering that the DFT

frequencies calculations were performed in the gas-phase. The N-H band in experimental

FT-IR spectra were at slightly lower frequencies due to intermolecular hydrogen

bonding260 interactions between the molecules. A similar pattern was observed for

hydroxyl stretching, where only one red-shifted hydroxyl band appeared in the spectra

which could be due to intermolecular and intramolecular hydrogen bonding as was

observed in similar reported organotin(IV) compounds.234 Complete IR data is provided

in Appendix, Table A4.1.

4.3.3 NMR Spectroscopic Analysis

The NMR spectra of the Schiff bases were recorded in DMSO-d6 and their

organotin(IV) compounds were recorded in CDCl3. The Schiff bases were present in the

thione tautomeric form in DMSO-d6 as indicated by the absence of an exchangeable SH

signal in the 1H NMR spectra.94 The spectra of the Schiff bases also exhibited a downfield

signal in the ~13.39-13.41 ppm range, which indicated the presence of the NH group. For

all the organotin(IV) compounds, the 1H NMR signal for the azomethine NH proton was

absent due to the coordination of the metal ion, in agreement with the absence of the

v(NH) FTIR peak mentioned above. Furthermore, a singlet signal appeared in the

downfield region of the spectrum, corresponding to a sp2-type OH proton. Both proton

chemical shifts for NH and OH disappeared in the spectra of the organotin(IV)

113

compounds, suggesting double deprotonation of the Schiff bases and hence confirming

the coordination to the tin(IV) ion through both phenolic oxygen and nitrogen atoms. The

complete 1H NMR data are collated in Appendix Table 4A.2.

The 13C NMR spectra of the Schiff bases showed a downfield chemical shift at ~196

ppm, indicating the presence of the C=S thione tautomer in solution. This signal was

shifted upfield in the organotin(IV) compounds in the range of 160-175 ppm, indicating

a decrease in electron density at the C-S carbon atom when the sulphur atom was

complexed to the tin atom.229 All other signals are tabulated in Appendix Table A4.3.

The tin chemical shift (119Sn) values of organotin(IV) compounds indicated the

metal’s coordination number, the substituent (phenyl, methyl, ethyl) group attached

directly to tin and the geometry of the organotin(IV) compounds: δ(119Sn) values of four-

coordinate complexes fall in the range +200 to −60 ppm; five-coordinate complexes fall

between −90 and −190 ppm, and six- and seven-coordinate complexes fall between −210

and −400 ppm.230 The high displacement of the 119Sn chemical shifts of the

diphenyltin(IV) compounds reported here (-233.4 to -233.7 ppm) is caused by the

electronegativity of the Schiff bases as well as the organo group attached to the tin

center.205 Dimethyltin(IV) compounds exhibited a higher chemical shift than

diphenyltin(IV) compounds in the range of -99.5 to -99.8 ppm, which was consistent with

their lower electron density and hence nuclear shielding. The 119Sn NMR values obtained

for both diphenyl- and dimethyltin(IV) compounds indicated that the compounds were in

a penta-coordinated trigonal bipyramidal geometry. This was corroborated by single

crystal XRD analysis.

114

4.3.4 Mass Spectral Analysis

The mass spectra were recorded for 10-12. The mass spectroscopic data revealed

that the molecular skeleton of 10-12 were obtained as predicted. The prominent molecular

ion peaks were observed at m/z 332.05 for 10 and 11, and m/z 318.05 for 12 respectively.

The mass spectra for 10, 11 and 12 are supplied in Appendix Figure A4.1.

4.3.5 UV-vis Absorption Spectroscopy

UV-vis spectra of compounds 10-18 were investigated experimentally and

theoretically in DMSO. B3LYP frontier orbitals of 10-12 are presented in Figure 4.2,

which shows that the HOMO was predominantly centred on the dithiocarbazate backbone

(S-(C=S)-NH-N), and the LUMO was largely centred on the 2,3-dihydroxyphenyl ring

and as well as dithiocarbazate backbone. Thus, the excitation of electrons observed in the

range 326-343 nm (experimental) and 335-345 nm (theoretical) was assigned as an n→π*

excitation, where nonbonding electrons of the azomethine nitrogen and sulphur atom in

the ground state were excited to the π* LUMO. The other excitations observed in the

range of 381-392 nm (experiment) and 372-386 nm (theoretical) indicated a π→π*

excitation involving the aromatic rings. In the spectra of the organotin(IV) compounds,

the HOMO was largely centred on the dithiocarbazate backbone and the 2,3-

dihydroxyphenyl ring moiety. Conversely, the LUMO was centred on the dithiocarbazate

backbone and 2,3-dihydroxyphenyl ring. The n→π* and π→π* transitions were observed

at 312-331 and 359-363 nm respectively, similar to the Schiff bases. The additional

intraligand band of the S→SnIV LMCT transition at 418-424 nm suggested coordination

of the Sn metal centre with sulphur atoms which was corroborated by evidence from

NMR and IR spectral analysis.93 All the transitions are summarised in Appendix Table

A4.4 and the HOMO-LUMO orbital diagrams are presented in Figure 4.3.

115

4.3.6 Structure Descriptions for 12, 14, 15, 17 and 18

The crystal and molecular structures of the hydrazinecarbodithioate, 12, a precursor

molecule for the structures of 15 and 18, and those of four diorganotin compounds, 14,

15, 17 and 18, were determined by X-ray crystallography. In 12, Figure 4.4 and Table

4.2, the central CN2S2 residue is planar with the RMSD of the fitted atoms being 0.011

Å; the maximum deviation from the least-squares plane is 0.018(4) Å for the N1 atom.

The dihedral angles formed between this central plane and the hydroxyl- and benzyl-rings

are 2.1(3) and 63.55(14)°, respectively, indicating co-planar and twisted relationships,

respectively; the dihedral angle between the rings is 61.63(17)°. The co-planarity in the

(phenylethylidene)hydrazine region of the molecule is due in part to the formation of

intramolecular hydroxyl-O‒H…O(hydroxyl) and hydroxyl-O‒H…N(imine) hydrogen

bonds to close S(5) and S(6) loops, respectively, details of which are given in Appendix

Figure A4.2.

116

Figure 4.3: Frontier molecular orbitals LUMO and HOMO of the optimised (a) 10, (b) 11, (c) 12, (d) 13, (e) 14, (f) 15, (g) 16, (h) 17 and (i) 18 using B3LYP functional.

117

Figure 4.4: Molecular structure of 12 and atom labelling scheme.

The crystallographic asymmetric unit of 14 comprises two independent molecules,

as shown in Figure 4.5a. Selected geometric data for all diorganotin structures described

herein are collated in Table 4.2. The tin atom is coordinated by thiolate-sulphur, imine-

nitrogen and phenoxide-oxygen atoms derived from the di-negative tridentate ligand. The

five-coordinate geometry about the tin atom is completed by two ipso-carbon atoms

derived from two phenyl substituents. A measure of the distortion away from ideal

trigonal-bipyramidal and square-pyramidal coordination geometries is given by the

descriptor τ 235 for which these ideal geometries correspond to τ = 1.0 and 0.0,

respectively. In 14, τ is 0.58 and 0.52 for the two independent molecules, indicating

environments almost half-way between these extremes. The structure of 15 presents very

similar features, again with two molecules in the asymmetric unit, Figure 4.5b. In this

case, the values of τ for the two molecules are 0.41 and 0.46, indicating intermediate five-

coordinate geometries but, a greater tendency towards a square-pyramidal coordination

structure. The availability of the crystal structure of hydrazinecarbodithioate 12 allows

comparisons of bond length and angle data with the carbonodithiohydrazonate dianion of

15. From Table 4.2, obvious differences exist in key geometric parameters such as the

elongation of the C1‒S1 bonds for the molecules in 15 cf. that in 12 as well as the

elongation of the N1‒N2 and C2‒N2 bonds with a concomitant decrease in the C1‒N1

118

bond length; no significant change is noted for the C1‒S2 bond. These systematic

variations indicate the formation of a C1‒S1 thiolate ligand as well as a second imine C2‒

N2 bond in 15. Given the comparable geometric parameters are equivalent within

experimental error to those in 14, that same conclusion pertains as, indeed, for the

structures of 17 and 18, Table 4.2. The reorganisation in π-electron density upon

deprotonation of the ligand and coordination to tin results in systematic reductions in the

angles subtended at the N1 and N2 atoms. The decrease and increase of π-electron density

in the C1‒S1 and C1‒N1 bonds upon coordination results in a decrease in the S1‒C1‒S2

angles and accompanying increases in the S1‒C1‒N1 and S2‒C1‒N1 bond angles. The

structures of 17 and 18 are isostructural and present τ values of 0.55 and 0.50,

respectively, i.e. in the range of values established for 14 and 15. In the same way, the

C‒Sn‒C angles in 17 and 18 lie within the range of C‒Sn‒C angles seen in 14 and 15,

Table 4.2.

119

Figure 4.5: ORTEP structures showing atom labelling scheme and overlay diagrams for (a) the two independent molecules of 14, (b) the two independent molecules of 15, (c) 17, (d) 18 and (e) overlay diagram for 17 and 18. For the overlay diagrams, molecules have been overlapped so that the CS2 residues are coincident.

120

Table 4.2: Selected geometric data (Å, °) characterising 12, 14, 15, 17 and 18

Parameter 12 14 14a 15 15a 17 18 Sn‒S1 ‒ 2.5319(12) 2.5434(9) 2.5262(8) 2.5312(7) 2.5477(15) 2.5676(14) Sn‒O1 ‒ 2.080(3) 2.100(2) 2.0974(18) 2.1004(17) 2.105(4) 2.120(3) Sn‒N2 ‒ 2.213(3) 2.191(3) 2.225(2) 2.223(2) 2.208(4) 2.201(4) N1‒N2 1.377(6) 1.404(4) 1.402(4) 1.398(3) 1.403(3) 1.405(5) 1.408(5) C1‒S1 1.660(6) 1.717(5) 1.729(4) 1.734(3) 1.731(3) 1.734(5) 1.743(5) C1‒S2 1.751(6) 1.768(4) 1.746(4) 1.754(3) 1.758(3) 1.758(4) 1.762(4) C1‒N1 1.330(7) 1.283(5) 1.292(4) 1.290(4) 1.292(3) 1.290(6) 1.277(6) C2‒N2 1.280(7) 1.295(5) 1.305(4) 1.301(3) 1.294(3) 1.293(6) 1.294(6) S1‒Sn‒O1 ‒ 159.77(9) 158.22(7) 155.87(6) 156.11(5) 159.17(11) 157.52(10) S1‒Sn‒N2 ‒ 77.84(9) 77.05(7) 77.53(6) 77.10(6) 77.39(10) 76.84(11) O1‒Sn‒N2 ‒ 82.99(12) 81.38(10) 81.36(8) 81.24(7) 81.78(14) 81.20(13) C‒Sn‒C ‒ 122.88(16) 126.93(15) 131.44(10) 128.57(10) 126.4(2) 127.8(2) C1‒N1‒N2 120.4(5) 115.4(4) 115.1(3) 115.4(2) 114.5(2) 115.7(4) 115.5(4) C2‒N2‒N1 117.5(5) 111.7(3) 112.7(3) 113.4(2) 113.0(2) 112.6(4) 113.0(4) S1‒C1‒S2 125.4(4) 112.1(2) 112.68(19) 111.86(17) 112.25(15) 112.3(3) 112.7(3) S1‒C1‒N1 121.3(5) 130.1(3) 128.3(3) 129.0(2) 129.3(2) 128.7(3) 128.8(4) S2‒C1‒N1 113.3(4) 117.7(3) 119.0(2) 119.1(2) 118.5(2) 119.0(4) 118.4(4)

121

The tridentate mode of coordination of the carbonodithiohydrazonate leads to the

formation of five-membered SnSCN2 and six-membered SnOC3N chelate rings. In the

first independent molecule of 14, the five-membered ring is planar [RMSD = 0.007 Å]

and to a first approximation, so is the six-membered ring which presents a RMSD = 0.042

Å. However, a more apt description is one based on an envelope as the tin atom lies

0.162(5) Å above the plane of the five remaining atoms of the chelate ring which exhibit

a RMSD of 0.023 Å. The angle between chelate rings is 3.74(16)°, and the angle between

the six-membered ring and appended phenyl ring is 3.2(2)°, both consistent with a planar

ligand. Quite distinct conformations are adopted by the chelate rings in the second

independent molecule of 14. Thus, each is best described as having an envelope

conformation with the tin ion lying 0.544(5) Å above the plane defined by the four

remaining atoms of the five-membered chelate ring [RMSD = 0.008 Å], and with tin

being 0.761(4) Å above the plane of the five remaining atoms for the six-membered

chelate [RMSD = 0.038 Å]. The dihedral angle between the planar portions of the chelate

rings is 34.23(15)°; the angle between the five co-planar atoms of the larger ring and

appended phenyl ring is 2.3(2)°. This stark difference in coordination mode is clearly

evident from the overlay diagram in Figure 4.5a. As seen in the overlay diagram of Figure

4.5b, there is higher degree of concordance between the two independent molecules of

15. However, a key difference exists between 14 and 15 in that in the latter, each chelate

ring has the form of an envelope. Thus, for the five-membered ring, the tin atom lies

0.320(4) Å above the plane of the remaining atoms [RMSD = 0.015 Å; second molecule:

0.472(4) Å/0.020 Å]. The equivalent parameters for the six-membered ring are 0.750(3)

Å/0.047 Å [0.788(3) Å/0.052 Å]. The dihedral angle between the planar regions of the

chelate rings is 26.47(11) [31.20(11)°]. The overlay diagram in Figure 4.5e confirms the

similarity of the molecular structures of 17 and 18. The chelate rings in 17 and 18 adopt

122

conformations as for the first independent molecule of 14 and both molecules of 15, with

tin ion deviations from the planes through the lighter atoms being 0.452(7) Å/0.007Å and

0.620(6) Å/0.046 Å, respectively, for the five- and six-membered rings of 17; the

comparable values for 18 are 0.516(7) Å/0.009 Å and 0.857(5) Å/0.063 Å, respectively.

The dihedral angle between the chelate rings in 17 of 27.8(2)° is less than 35.8(2)° in 18.

With the exception of the very recently reported diorganotin derivatives bearing

ligands related to those in the present study, i.e. with a methoxy substituent adjacent to

the phenoxide-oxygen atom, and which present similar distorted five coordinate

geometries,178 there are only two other crystallographically-characterised tin structures

available in the Cambridge Structural Database.261 In each of these, the phenoxide ring

is unsubstituted and there is an ethyl group bound to the ester-sulphur atom, and similar

tridentate modes of coordination are noted. A distorted octahedral geometry based on a

N2O2S6 donor set with oxygen and sulphur atoms trans to each other is found in the

homoleptic species.118 In the other available structure, a distorted octahedral geometry is

also found with three positions occupied by the tridentate ligand, two by iodide atoms and

the sixth site by an oxygen atom derived from a dimethylformamide ligand.118

In the molecular packing of 12, centrosymmetric molecules assemble into dimeric

aggregates via hydroxyl-O‒H…O(hydroxyl) hydrogen bonds leading to 10-membered

…HOC2O2 synthons. These are connected into a linear supramolecular chain along [1

0 4] by thioamide-N‒H…S(thione) hydrogen bonds between centrosymmetrically related

molecules via eight-membered …HNCS2 synthons, Figure 4.6. The chains are

assembled into a three-dimensional architecture by hydroxyphenyl-C‒H…π(benzyl)

interactions; a view of the unit cell contents is shown Appendix Figure A4.3. Geometric

parameters characterising these interactions and those for the other structures discussed

123

herein are included in the respective figure captions of the Supplementary information, in

this case Appendix Figure A4.3.

Figure 4.6: A view of the linear supramolecular chain along [1 0 4] in the crystal of 12. The hydroxyl-O‒H…O(hydroxyl) and thioamide-N‒H…S(thione) hydrogen bonds are shown as orange and green dashed lines, respectively.

Along with the structures of 15, 17 and 18, there is an intramolecular charge-

assisted hydroxyl-O‒H…O(phenoxide) hydrogen bond in each molecule of 14 which

precludes the participation of this atom in hydrogen bonding in these crystals. An

interesting feature of the molecular packing of 14 is the presence of π…π interactions

involving the only planar chelate ring observed among the diorganotin species reported

herein (see above) and a symmetry-related hydroxyphenyl ring. The distance between

the ring centroids is 3.874(2) Å and the angle between rings is 0.95(15)° for symmetry

operation x, ½-y, -½+z; the slippage is 1.69 Å, indicating the usually-observed off-set

disposition of the rings.239 Such interactions are becoming increasingly apparent in the

supramolecular chemistry of heavy-element compounds and can impart energies of

stabilisation to their crystals equal to or greater than conventional hydrogen bonding

interactions.237,239 In the present structure, these interactions lead to the presence of

supramolecular chains with a zig-zag topology (glide symmetry) along the c-axis, Figure

4.7. The connections between the chains are tin-bound phenyl- and hydroxyphenyl-C‒

H…π(Sn-bound phenyl) and contribute to the stabilisation of a layer in the bc-plane; layers

124

stack along the a-axis without directional interactions between them, Appendix Figure

A4.3.

In the crystal of 15, distinct supramolecular motifs are observed containing one of

the independent molecules exclusively. Thus, a linear supramolecular chain containing

Sn1-molecules only is sustained by benzyl-phenyl-C‒H…O(hydroxyl) interactions,

Appendix Figure A4.4a, and a supramolecular dimer of Sn1a-molecules and sustained by

tin-bound-phenyl-C‒H…O(hydroxyl) interactions, Appendix Figure A4.4b, are noted.

The chains are connected into a three-dimensional architecture by a combination of

hydroxyphenyl-C‒H…π(tin-bound phenyl) and π(tin-bound phenyl)…π(tin-bound phenyl)

interactions, Appendix Figure A4.4c .

The crystals of 17 and 18 are isomorphous, as shown in Table 4.1. In the crystal of

17, a three-dimensional architecture is stabilised by tin-bound-methyl-C‒H...π(benzyl-

phenyl) and hydroxyphenyl-C‒H…O(hydroxyl) interactions, as detailed in Appendix

Figure A4.5. Analogous interactions are apparent in the crystal of 18, Appendix Figure

A4.6.

The gas phase B3LYP geometries of 12, 14, 15, 17 and 18 are in good agreement

with the values obtain from crystallographic analysis above (see Appendix Table A4.5),

consistent with the absence of strong directional of inter- and intramolecular interactions

in the solid form of the crystals. The deviation in selected bond lengths and bond angles

of 12, 14, 15, 17 and 18 are less than 0.038 Å and 8.0°, respectively, where the maximum

bond length deviation corresponded to the Sn-O1 bond and the maximum bond angle

deviation corresponded to the organo group attached to the tin, C-Sn-C angle.

125

Figure 4.7: A view of the zig-zag chain along [0 0 1] in the crystal of 14. The hydroxyl-O‒H…O(hydroxyl) and thioamide-N‒H…S(thione) hydrogen bonds are shown as orange and green dashed lines, respectively. For reasons of clarity, the hydrogen atoms have been removed and only the ipso-carbon atoms of the tin-bound phenyl rings shown.

4.3.7 In Vitro Cytotoxicity

Compounds 10-18 were tested for their cytotoxic activity against ten cancer cell

lines (HT29, U87, SJ-G2, MCF-7, A2780, H460, A431, DU145, BE2-C, MIA) and one

normal breast cell line (MCF-10A) using the MTT metabolic assay. The growth inhibition

concentration of the Schiff bases (10-12) and their organotin(IV) compounds (13-18)

required to inhibit 50% cell proliferation relative to the cells that were treated with DMSO

(control) are tabulated in Table 4.3. The stability of these compounds was evaluated in

DMSO as well as in a mixture of DMSO-H2O by monitoring UV-vis absorbance. The

unchanged pattern in the spectra indicated that the compounds were stable in both

solvents.

Compound 10 showed selectivity against the BE2-C cell line and the highest

potency (GI50 = 0.38 ± 0.07 μM, refer Table 4.3) of the Schiff base analogues 11 and 12.

However, 10 exhibited the lowest cytotoxicity against HT29, U87, MCF-7, A431 and

Du145 cell lines. Moderate cytotoxic activity was observed for 11 with no significant

selectivity, with slightly lower activity than 12 in all cancer cell lines investigated here,

suggesting that the presence of the methyl group at the ortho and para positions did not

126

impact cytotoxicity significantly, an observation that correlated well with what was

reported in previous work.178 Complexation with diphenyltin(IV) dichloride improved the

activity of the Schiff bases, with diphenyltin(IV) compounds (13, 14 and 15)

demonstrating good cytotoxicity against A2780, BE2-C, SJ-G2 and MIA cells. The

highest level of selectivity of 13, 14 and 15 was observed for the A2780 cells, with GI50

values of 0.31 ± 0.05, 0.23 ± 0.02 and 0.19 ± 0.01 μM, respectively, which corresponded

to 6- and 10-fold higher potencies compared to 10, 11 and 12 (the latter had GI50 values

of 1.8 ± 1.2, 2.8 ± 0.20 and 2.2 ± 0.25 μM, respectively). Furthermore, 13 also showed

selectivity against HT29, MCF-7 and Du145 cell lines, with 10-fold, 15-fold and 57-fold

higher potency than its Schiff base, 10. A 3-fold greater activity against U87 was observed

for 15 with a GI50 value of 0.69 ± 0.03 μM as compared to its Schiff base, 12 (GI50 = 2.7

± 0.47 μM). The selectivity of 14 and 15 for A2780 was 8- and 5-fold higher, respectively,

over the normal breast cell line, MCF10A. There was a significant contrast in the

cytotoxic activity of dimethyltin(IV) compounds (16, 17 and 18) as compared to the

diphenyltin(IV) compounds, which suggested that the presence of the two phenyl groups

coordinating the tin(IV) centre enhanced the cytotoxicity of the compounds. Compounds

16, 17 and 18 exhibited no obvious enhancement in cytotoxicity as compared to their

respective Schiff bases, which was similar to that observed in previous work.178

4.3.8 DNA Binding Analysis

A number of studies66,142,152,262 have reported that cytotoxic agents can cleave or

bind to DNA. The interaction between organotin(IV) compounds (13-18) and CT-DNA

is therefore examined here via UV-vis absorption spectroscopy. Figure 4.8 shows the UV-

vis spectra observed where dimethyltin(IV) compounds (16, 17 and 18) in solutions of

varying DNA concentrations. These compounds had absorption peaks at 304 - 312, 358 -

127

362 and 426 - 434 nm, which indicated π-π* (due to aromatic group), n-π* (due to either

C=N or OH) and ligand-metal charge transfer excitations, respectively. By increasing the

concentrations of CT-DNA, significant hypochromism and red shifts were observed in

the spectra, see Table 4.4. This hypochromism was attributed to the strong interactions

between the organotin(IV) π electrons and DNA base pair π electrons. Therefore, the

energy level of the π→π* electron transition decreased and shifted to a higher

wavelength.263 The change in transitions were observed in the spectra of compounds 16,

17 and 18, but only slight changes were observed in diphenyltin(IV) compounds (13, 14

and 15) (Figure A4.7). It is therefore conceivable that the cytotoxic activities of

compounds 13, 14 and 15 could be partly due to the interaction between the compounds

and some proteins or organelles in cancer cells, not exclusively only via binding to

DNA.224 Nevertheless, negative Gibbs free energies (∆G) obtained from this

spectroscopic analysis indicate that the DNA - organotin(IV) compound interaction is

spontaneous.

4.3.9 Molecular Docking Analysis

In order to understand specific interactions between the six organotin(IV)

compounds (13-18) and CT-DNA, we turn to molecular docking simulations using a

model DNA duplex of sequence d(CGCGAATTCGCG)2. Results are detailed in Table

4.5. The organotin(IV) compounds mainly bound to the groove of the DNA helix via non-

covalent interactions, viz., hydrogen bonding, hydrophobic, van der Waals and

electrostatic interactions. These non-covalent interactions are usually stabilised by π-

interactions between aromatic groups of the compounds 13-18 and the DNA duplex, and

therefore provide a structural basis for the observed hypochromism detailed in the

previous section. Figure 4.9 shows the most favourable conformation of the docked pose

128

which exhibited the lowest binding energy, revealing that compounds 13-18 interact with

the AT-rich region of the minor groove. The DG10 and DT20 DNA residues played a

major role in the interaction with diphenyltin(IV) compounds (13-15), whereas DG10 and

DT8 of the DNA residues predominantly interacted with dimethyltin(IV) compounds (16-

18). The interactions mainly occurred through the phosphate backbone of the DNA, as

well as via nitrogenous bases of DNA.

129

Table 4.3: Summary of the in vitro cytotoxicity of the Schiff bases (10-12) and their organotin(IV) compounds (13-18) in several cell lines, determined by MTT assay, expressed as GI50 values with standard errors. GI50 is the concentration at which cell growth was inhibited by 50% over 72 h.

Compounds Growth inhibition concentration, GI50 (µM)

HT29 U87 MCF-7 A2780 H460 A431 Du145 BE2-C SJ-G2 MIA MCF10A

10 10 ± 3.5 20 ± 5.2 10 ± 3.1 1.8 ± 1.2 8.7 ± 1.9 14 ± 1.5 23 ± 3.6 0.38 ± 0.07 6.4 ± 2.2 5.5 ± 1.1 3.7 ± 0.50

11 2.8 ± 0.03

4.0 ± 0.43

2.8 ± 0.07

2.8 ± 0.20

3.7 ± 0.10

3.7 ± 0.20

4.2 ± 0.59

3.2 ± 0.30

3.3 ± 0.30 6.6 ± 1.7 2.9 ± 0.03

12 2.1 ± 0.35

2.7 ± 0.47

2.3 ± 0.32

2.2 ± 0.25

2.6 ± 0.27

2.4 ± 0.26

2.8 ± 0.15

2.0 ± 0.23

2.2 ± 0.47

2.2 ± 0.64 2.9 ± 0.21

13 0.85 ± 0.47

1.2 ± 0.49

0.54 ± 0.16

0.31 ± 0.05

1.2 ± 0.20

1.0 ± 0.20

0.43 ± 0.22

0.38 ± 0.08

0.27 ± 0.09

0.42 ± 0.12

0.47 ± 0.13

14 1.8 ± 0.06

1.1 ± 0.27

1.6 ± 0.20

0.23 ± 0.02

2.3 ± 0.07

2.1 ± 0.09

1.9 ± 0.03

0.67 ± 0.09

0.94 ± 0.13

0.69 ± 0.03 1.9 ± 0.09

15 1.3 ± 0.29

0.69 ± 0.03

1.5 ± 0.20

0.19 ± 0.01

1.3 ± 0.54

1.5 ± 0.24

1.5 ± 0.18

0.35 ± 0.04

0.89 ± 0.36

0.27 ± 0.08 1.1 ± 0.48

16 5.7 ± 2.4 6.0 ± 1.6 2.9 ± 0.54

2.9 ± 0.39

3.5 ± 0.18

3.2 ± 0.38

2.5 ± 0.09

2.3 ± 0.21

1.4 ± 0.26

3.6 ± 0.83 3.0 ± 0.15

17 2.3 ± 0.19

3.0 ± 0.06

2.7 ± 0.32

2.6 ± 0.19

2.5 ± 0.03

2.9 ± 0.18

3.1 ± 0.46

2.2 ± 0.27

1.9 ± 0.10

3.9 ± 0.94 2.4 ± 0.12

18 1.6 ± 0.34

2.7 ± 0.10

2.0 ± 0.32

2.3 ± 0.31

1.9 ± 0.17

2.2 ± 0.13

3.5 ± 0.26

1.7 ± 0.23

1.4 ± 0.15

2.1 ± 0.23 2.2 ± 0.12

Cisplatin 11.0 ± 2.0 4.0 ± 1.0 6.5 ± 0.8 1.0 ± 0.1 0.9 ± 0.2 2.4 ± 0.3 1.2 ± 0.1 1.9 ± 0.2 0.4 ± 0.1 8.0 ± 1.0 nd

GI50 (μM): (the colors indicate) = 0.1–0.99 (strong); = 1.0–9.9 (moderate); = 10-100 (weak); = nd (not determined)

130

Figure 4.8: (a) Electronic absorption spectra of (i) 16, (ii) 17 and (iii) 18 ; (b) Plot of [DNA]/εa - εf vs [DNA] for absorption titration of DNA with (i) 16, (ii) 17 and (iii) 18. (The arrow indicates the change in absorbance in tandem with increasing DNA concentration).

131

Table 4.4: Binding constants (Kb), hypochromism (%), bathochromic shifts (nm) and Gibbs free energy (kJ mol-1) values for the interaction of organotin(IV) compounds with CT-DNA.

Compounds Binding constant

(Kb)

Hypochromism, (%)

Bathochromism shift (nm)

Gibbs free energy (kJ

mol-1)

13 4.71 x 104 7.46 2 -26.7 14 4.17 x 104 5.23 1 -26.4 15 1.68 x 106 6.25 1 -35.5 16 1.1 x 105 50.89 10 -28.8 17 2.88 x 105 56.65 16 -31.1 18 1.99 x 106 43.01 17 -35.9

132

Figure 4.9: (a) Schematic representation of organotin(IV) compounds that fit well in the grooves of the DNA strand obtained by docking simulations. The two double-stranded DNA comprised of the phosphate deoxyribose backbone with guanine (DG, red), cytosine (DC, blue), adenine (DA, pink) and thymine (DT, orange). (b) Molecular interactions of organotin(IV) compounds within the grooves of double stranded DNA residues.

133

Table 4.5: Molecular docking data for the organotin(IV) compounds with B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2.

Complex

Final Intermolecular Energy, kcal mol−1 Final total Internal

Energy (2), kcal mol−1

Torsional Free

Energy (3), kcal mol−1

Unbound System’s

Energy [=(2)] (4), kcal mol−1

Estimated Free Energy of

Binding [(1) + (2) + (3) − (4)],

kcal mol−1

Estimated Free Energy of Binding,

kJ mol−1

vdW + Hbond + desolv. Energy

Electrostatic Energy

Total (1)

13 -10.30 -0.05 -10.36 -2.41 1.79 -2.41 -8.57 -35.86

14 -12.25 -0.12 -12.36 -2.20 1.79 -2.20 -10.57 -44.22

15 -10.42 0.00 -10.43 -2.34 1.79 -2.34 -8.64 -36.15

16 -10.27 -0.07 -10.24 -1.15 1.19 -1.15 -9.04 -37.82

17 -10.73 -0.06 -10.80 -0.94 1.19 -0.94 -9.60 -40.17

18 -10.58 -0.09 -10.67 -0.90 1.19 -0.90 -9.48 -39.66

134

4.4 Conclusions

Diphenyltin(IV) (13-15) and dimethyltin(IV) (16-18) compounds containing ONS-

tridentate 2,3-dihydroxybenzyldithiocarbazate Schiff bases (10-12) were synthesised and

characterised by analytical, spectroscopic and single crystal X-ray diffraction studies. The

cytotoxic activities of these compounds against a panel of ten cancer cell lines were

evaluated. A good correlation was found between calculated geometries, electronic

absorptions and vibrational frequencies from experimental data, single crystal XRD, UV-

visible and FT-IR spectroscopy. X-ray crystallography established the anticipated

tridentate mode of coordination between the dinegative Schiff bases and the diorganotin

centres, resulting in highly distorted five-coordinate C2NOS geometries. In vitro

cytotoxic assays revealed remarkable cytotoxicity of diphenyltin(IV) compounds with

GI50 values in the range 0.19–2.3 μM for all cancer cell lines. This study has also

demonstrated that diphenyltin(IV) compounds exhibited selective potency against the

U87, A2780, BE2-C, SJ-G2 and MIA cell lines, with higher cytotoxicity than their Schiff

bases and dimethyltin(IV) compounds which were comparable with the diphenyltin

compounds discussed in Chapter 3. The DNA binding data obtained using UV-vis

spectroscopy showed that the dimethyltin(IV) compounds bound to the DNA via groove

binding, resulting in hypochromism and redshift in the principal UV-vis excitation. The

molecular docking results showed that DG10 and DT20 of the DNA residues were crucial

for the binding to diphenyltin(IV) compounds, whereas all dimethyltin(IV) compounds

were bound to DG10 and DT8 of the DNA residues.

Exciting cytotoxicity outcomes were observed for dithiocarbazate Schiff bases and

their organotin(IV) compounds which was previously discussed in this chapter and

previous chapters. It was also demonstrated that complexation enhanced the

antiproliferative effects against a panel of cancer cell lines tested, particularly

135

diphenyltin(IV) compounds suggesting that the presence of two phenyl groups attached

to the tin atom played a main role in inhibiting cancer cells, particularly by interacting

aromatic phenyl ring of compounds with nitrogenous bases of DNA of cancer cells.

Slightly higher cytotoxicity were also observed for dimethyltin(IV) compounds. Based

on this observation, further investigation required to study the absence of organo groups

attached to the tin atom. Recently, tin(IV) compound coordinated with two molecules of

dithiocarbazate Schiff base showed high potency against MCF-7 cell due to its ability to

bind to human serum albumin protein.118

Further investigations will be discussed in Chapter 5, which reports structural and

cytotoxicity studies of six homoleptic tin(IV) compounds coordinated with two molecules

of dithiocarbazate Schiff bases 1, 2, 3, 10, 11, 12 (discussed in Chapter 2 and 4). The

theoretical prediction of tin(IV) compound-DNA binding will also discussed in Chapter

5.

136

CHAPTER 5

HOMOLEPTIC TIN(IV) COMPOUNDS OF DINEGATIVE ONS DITHIOCARBAZATE SCHIFF BASES: SYNTHESIS, X-RAY

CRYSTALLOGRAPHY, DFT AND CYTOTOXICITY STUDIES

5.1 Introduction

Two decades ago, octahedral homoleptic tin(IV) compounds containing S-benzyl-

β-N-(2-hydroxyphenyl)methylene264, S-methyl-β-N-(2-hydroxyphenyl) methylene264, S-

methyl-β-N-(2-hydroxy-3-methoxy-phenyl)methylenedithiocar-bazate,265 S-methyl-β-N-

(2-hydroxy-1-naphthyl)methylenedithiocarbazate,265 S-benzyl-β-N-(2-hydroxy-3-

methoxyphenyl)methylenedithiocarbazate,266 S-benzyl-β-N-(2-hydroxy-1-naphthyl)

methylene dithiocarbazate266 and S-benzyl-β-N-(2-ferrocenoylacetone) methylene

dithiocarbazate267 were synthesised and structurally characterised. However, the

cytotoxic studies of these compounds remain undisclosed. Recently, the discovery of

homoleptic tin(IV) compounds coordinated with dithiocarbazate Schiff bases were

reported by Yekke-Ghasemi et al.118 They found that tin(IV) compound of

2,2’(disulfanediylbis((ethylthio) methylene)bis(hydrazin-2-yl-1-ylidene) bis(methanyl

ylidene))diphenol exhibited good cytotoxicity against MCF-7 cell compared to the

clinical drug, cisplatin. This compound was also observed to bind to the most abundant

protein in blood plasma, human serum albumin protein.118 However, the cytotoxicity

studies of tin(IV) compounds coordinated with two molecules of oVa or catechol

dithiocarbazate Schiff bases remains unexplored. It is hypothesised here that homoleptic

tin(IV) compounds containing Schiff bases would have good cytotoxic properties, and

are able to interact with biomacromolecules.

Following on the cytotoxicity results reported in Chapter 3 and 4, this chapter

investigates this hypothesis by focusing on the synthesis, structural evaluation and

137

cytotoxic studies of homoleptic tin(IV) compounds containing two molecules of oVa or

catechol dithiocarbazate Schiff bases with general formula, Sn(L)2 (L= Schiff bases 1, 2,

3, 10, 11 and 12), see Figure 5.1.

Figure 5.1: The structures of homoleptic tin(IV) compounds (19-24).

This chapter contains material from submitted article for publication in Journal of

Molecular Structure on 27 September 2019.

5.2 Experimental

5.2.1 Materials and Physical Measurements

All solvents and reagents were of analytical reagent grade and used without further

purification as described in Chapter 2 (§2.2.1) and Chapter 3 (§3.2.1). Tin(II) dichloride,

97% (Merck, New York, NY, USA) and triethylamine, > 99% (Sigma Aldrich, St. Louis,

RS N

SN

OR'

RSN

SN

OR' Sn

Where: R = o-CH3, R' = O-CH3 (19) R = p-CH3, R' = O-CH3 (20) R = H, R' = O-CH3 (21) R = o-CH3, R' = OH (22) R = p-CH3, R' = OH (23) R = H, R' = O-H (24)

138

MO, USA). The instrumentations used are as outlined in Chapter 2 (§2.2.3) and Chapter

3 (§3.2.3).

5.2.2 Synthesis

5.2.2.1 Synthesis of Schiff Bases

Schiff bases were synthesised following references 177,178. The details of this

synthesis procedure are described in Chapter 2 (§2.2.2).)

5.2.2.2 Synthesis of Tin(IV) Compounds (19-24)

The Schiff base (0.35 g (1, 2); 0.33 g (3, 10, 11); 0.32 g (12), 1 mmol) was dissolved

in absolute ethanol (50 cm3) and dichloromethane (20 cm3) and was mixed with an

ethanolic solution (2 mmol) of triethylamine. With continuous stirring, an ethanolic

solution of SnCl2 (0.20 g, 1 mmol) was added to the mixture. A cloudy solution formed

which was filtered. The yellow filtrate was then heated under reflux for ca. 6 h. The

mixture was left overnight and the reduction of the volume of the reaction mixture

resulted in an orange precipitate, which was filtered off. The precipitate was recrystallised

in dry ether to remove the remaining triethylamine hydrochloride salt from the mixture.

2.3.2.1. Tin(IV) [S-2-methybenzyl-β-N-(2-hydroxy-3-methoxybenzylmethylene)

dithiocarbazate] (19)

Orange solid. Yield: 82 %. M.p.: >300 °C. Analysis calculated for

C34H32N4O4S4Sn: C, 50.56; H, 3.99; N, 6.94%. Found: C, 50.87; H, 3.58; N, 5.27%. FT-

IR (ATR, cm-1): 1587, v(C=N); 1088, v(N-N); 955, v(C-S). 1H NMR (CDCl3) δ (ppm.):

8.81, 8.44 (s, 2H, CH), 6.77-7.33 (multiplet, 14H, Ar-H), 4.47, 4.33 (s, 4H, CH2), 3.84,

3.57 (s, 3H, O-CH3), 2.40, 2.36 (s, 6H, CH3); 13C NMR (CDCl3) δ (ppm.): 177.2 (C-S),

139

171.0 (C=N); 115.1-171.0 (aromatic-C), 57.0, 55.1 (O-CH3), 19.4, 19.4 (CH3), 36.3, 33.9

(CH2).119Sn NMR (CDCl3) δ (ppm.): -446.3.

2.3.2.2. Tin(IV) [S-4-methybenzyl-β-N-(2-hydroxy-3-methoxybenzyl methylene)

dithiocarbazate] (20)

Orange crystals. Yield: 87 %. M.p.: 211-214 °C. Analysis calculated for

C34H32N4O4S4Sn: C, 51.56; H, 3.99; N, 6.94%. Found: C, 51.89; H, 4.22; N, 6.53%. FT-

IR (ATR, cm-1): 1593, v(C=N); 1083, v(N-N); 958, v(C=S). 1H NMR (CDCl3) δ (ppm.):

8.78, (s, 2H, CH), 6.76-7.36 (multiplet, 14H, Ar-H), 4.45 (s, 4H, CH2), 3.56 (s,3H, O-

CH3), 1.61 (s, 6H, CH3); 13C NMR (CDCl3) δ (ppm.): 170.8 (C-S), 165.4 (C=N); 117.2-

156.7 (aromatic-C), 56.9 (O-CH3), 21.2 (CH3), 35.5 (CH2).119Sn NMR (CDCl3) δ (ppm.):

-446.2.

2.3.2.3. Tin(IV) [S-benzyl-β-N-(2-hydroxy-3-methoxybenzylmethylene)

dithiocarbazate] (21)

Orange crystals. Yield: 53 %. M.p.: >300 °C. Analysis calculated for

C32H28N4O4S4Sn: C, 49.30; H, 3.62; N, 7.19%. Found: C, 49.26; H, 3.26; N, 6.02%. FT-

IR (ATR, cm-1): 1582, v(C=N); 1031, v(N-N); 958, v(C-S). 1H NMR (CDCl3) δ (ppm.):

8.78 (s, 2H, CH), 6.76-7.36 (multiplet, 16H, Ar-H), 4.45 (s, 2H, CH2), 3.56 (s, 3H, O-

CH3); 13C NMR (CDCl3) δ (ppm.): 170.7 (C-S), 165.5 (C=N); 117.2-156.7 (aromatic-C),

56.9 (CH3), 35.7 (CH2).119Sn NMR (CDCl3) δ (ppm.): -447.1.

140

2.3.2.4. Tin(IV)[S-2-methybenzyl-β-N-(2,3-dihydroxybenzylmethylene)

dithiocarbazate] (22)

Orange solid. Yield: 40 %. M.p.: 188-192 °C. Analysis calculated for

C32H28N4O4S4Sn: C, 49.30; H, 3.62; N, 7.19%. Found: C, 49.14; H, 3.43; N, 7.50%. FT-

IR (ATR, cm-1): 1610, v(C=N); 1035, v(N-N); 964, v(C-S). 1H NMR (CDCl3) δ (ppm.):

8.90 (s, 2H, CH), 6.02-7.34 (multiplet, 14H, Ar-H), 4.48 (s, 4H, CH2), 2.42 (s, 6H, CH3);

13C NMR (CDCl3) δ (ppm.): 171.3 (C-S), 165.6 (C=N), 115.1, 119.0, 119.6, 125.4, 126.5,

128.3, 130.5, 130.8, 133.2, 137.4, 148.2, 152.1 (aromatic-C), 34.3 (CH2), 19.5

(CH3).119Sn NMR (DMSO-d6) δ (ppm.): -440.9.

2.3.2.5. Tin(IV) [S-4-methybenzyl-β-N-(2,3-dihydroxybenzylmethylene)

dithiocarbazate] (23)

Orange solid. Yield: 83 %. M.p.: 206-209°C. Analysis calculated for

C32H28N4O4S4Sn: C, 49.30; H, 3.62; N, 7.19%. Found: C, 49.02; H, 3.16; N, 7.85%. FT-

IR (ATR, cm-1): 1620, v(C=N); 1006, v(N-N); 965, v(C-S). 1H NMR (DMSO-d6) δ

(ppm.): 8.65, (s, 2H, CH), 6.50-7.35 (multiplet, 14H, Ar-H), 4.41 (s, 4H, CH2), 2.29 (s,

6H, CH3); 13C NMR (DMSO-d6) δ (ppm.): 150.8 (C-S), 145.2 (C=N); 118.4, 129.5, 129.6,

134.2, 136.8 (aromatic-C), 21.2 (CH3), 37.8 (CH2).119Sn NMR (DMSO-d6) δ (ppm.): -

443.1.

2.3.2.6. Tin(IV) [S-benzyl-β-N-(2,3-dihydroxy-benzylmethylene)

dithiocarbazate] (24)

Orange solid. Yield: 67 %. M.p.: 206-208 °C. Analysis calculated for

C30H24N4O4S4Sn: C, 47.95; H, 3.22; N, 6.46%. Found: C, 47.72; H, 2.93; N, 6.25 %. FT-

IR (ATR, cm-1): 1612, v(C=N); 1016, v(N-N); 960, v(C-S). 1H NMR (DMSO-d6) δ

141

(ppm.): 8.60, (s, 2H, CH), 6.27-7.41 (multiplet, 16H, Ar-H), 4.44 (s, 4H, CH2); 13C NMR

(DMSO-d6) δ (ppm.): 155.9 (C-S), 153.4 (C=N); 111.6, 114.3, 115.9, 116.9, 127.5, 128.9,

129.7, 137.6, 146.9 (aromatic-C), 38.1 (CH2).119Sn NMR (DMSO-d6) δ (ppm.): -461.9.

5.2.3 Single Crystal X-ray Structure Determination

An Oxford Diffraction Gemini Eos CCD diffractometer fitted with Mo Kα radiation

source (λ = 0.71073 Å) was employed to measure intensity data for the orange crystals of

20 and 21 at T = 150 K. The data reduction and analytical absorption corrections were

accomplished with CrysAlisPro180. The structures were solved by direct-methods181 and

refined (anisotropic displacement parameters, C-bound H atoms in the riding model

approximation) on F2 182. A weighting scheme w = 1/[σ 2(Fo2) + (aP)2 + bP] where P =

(Fo2 + 2Fc2)/3) was introduced in each case. In the final cycles of the refinement of 21, a

reflection, i.e. (-6 -4 8), was omitted owing to poor agreement. The molecular structure

diagrams were generated with ORTEP for Windows183 with 50% displacement ellipsoids

and the packing diagrams were drawn with DIAMOND184. Additional data analysis was

made with PLATON216. Crystal data and refinement details are given in Table 5.1.

5.2.4 DFT Calculations and Molecular Docking Studies

DFT and molecular docking simulations were performed using parameters detailed

in Chapter 2 (§2.2.6) and Chapter 3 (§3.2.5.1), respectively. The coordinates of 20 and

21 were taken from their respective crystal structures, whereas the coordinates for 19, 22,

23 and 24 were obtained after DFT optimisation. The parameters for molecular docking

simulations are described in Chapter 3 (§3.2.5.2).

142

Table 5.1: Crystal data and refinement details for complexes 20 and 21. Complex 20 21 Formula C34H32N4O4S4Sn C32H28N4O4S4Sn Formula weight 807.56 779.51 Crystal system Monoclinic Monoclinic Space group P21/n P21/c a/Å 12.8540(2) 12.8656(3) b/Å 19.4686(5) 8.4693(2) c/Å 27.2466(6) 29.8639(7) β/° 101.109(2) 96.634(2) V/Å3 6690.7(3) 3232.26(13) Z 8 4 Dc/g cm-3 1.603 1.602 F(000) 3280 1576 µ(MoKα)/mm-1 1.059 1.093 Measured data 44247 36728

range/° 3.3 – 25.2 3.4 – 25.2 Unique data 15972 7934 Observed data (I 2.0σ(I)) 10737 6324 No. parameters 855 408 R, obs. data; all data 0.041; 0.079 0.035; 0.072 a; b in weighting scheme 0.036; 1.389 0.032; 2.061 Rw, obs. data; all data 0.076; 0.093 0.052; 0.080 GoF 1.05 1.02 Range of residual electron density peaks/eÅ-3 -0.75 – 0.91 -0.57 – 0.50

5.2.5 MTT Assays

The protocol used for MTT assay followed that outlined in Chapter 2 (§2.2.6).

5.3 Results and Discussion

5.3.1 Synthesis

The six new tin(IV) compounds were synthesised by the condensation reaction of

S-substituted dithiocarbazate Schiff bases of 2-hydroxy-3-methoxybenzaldehyde or 2,3-

dihydroxybenzaldehyde with tin(II) chloride (Scheme 5.1), whereby the tin(II) precursor

underwent oxidation to tin(IV) after reflux step for 6 hours.159 The tin(IV) compounds

were stable at room temperature and soluble in most organic solvents, especially

dimethylsulfoxide (DMSO) and dimethylformamide (DMF). However, 20 was not

143

soluble in DMSO; consequently, cytotoxic activity of this compound was not determined.

The molar conductance of 10-3 M solutions of synthesised compounds in DMSO were in

the range 1.12-6.07 Ω-1 cm2 mol-1, indicating their non-electrolytic nature.106

Scheme 5.1: Synthetic pathway for the synthesis of 19-24.

5.3.2 IR Spectral Analysis

Complete experimental and calculated IR data are provided in Appendix, Table

A5.1. The FT-IR spectra of 19-24 were measured between 4000-280 cm−1 and verified

using DFT calculations. All DFT vibrational frequencies were scaled using a scaling

factor of 0.9682.189 For the uncomplexed Schiff bases, v(O-H), v(N-H), v(C=N), v(N-N)

and v(C=S) bands are observed at 3488-3501, 3084-3106, 1598-1608, 1114-1125 and

1012-1030 cm-1, respectively. However, v(O-H) of 1, 2 and 3 were not observed in the IR

spectra due to the intra- and inter-molecular interactions between the molecules, which

were reported in Chapter 2.178 The disappearance of the v(N-H) band for all Schiff bases

and the v(O-H) band for 10, 11 and 12 indicated the deprotonation of N-H and O-H groups

and its subsequent coordination to the central tin atom. The shifting of the v(C=N) band

RS N

H

SN

HOR'

RS N

SN

OR'

RSN

SN

OR' Sn

Et3N, EtOHreflux (6 hours)

Where: R = o-CH3, R' = O-CH3 (19) R = p-CH3, R' = O-CH3 (20) R = H, R' = O-CH3 (21) R = o-CH3, R' = OH (22) R = p-CH3, R' = OH (23) R = H, R' = O-H (24)

SnCl2

144

observed in the spectra of tin(IV) compounds further proves that the complexation

occurred through the azomethine nitrogen. The IR group assignments for the

experimental and theoretical calculation in gas phase appeared to be in a good agreement

with experimental data.

5.3.3 NMR Spectroscopic Analysis

All 1H NMR and 13C NMR spectra are tabulated in Appendix Tables A5.2 and A5.3,

respectively. The important signals observed in the range of 9.51-9.61 ppm and 13.32-

13.41 ppm were assigned to the OH and -NH protons in the Schiff base, respectively, as

reported previously in Chapter 2 and 4. Both of the signals disappeared in the 1H NMR

of the tin(IV) compounds indicating the deprotonation of the -OH and -NH groups upon

complexation. This confirmed that the coordination of the tin ion occurs via the oxygen

and nitrogen donor atom of the Schiff bases. Moreover, the signal of the HC=N proton

shifted to the downfield region due to the coordination of –NH with the tin centre, which

was a result of the formation of the C=N-N=C conjugated systems.66 The downfield shift

of the HC=N signal in the 13C NMR spectra, due to the electron density transfer from the

Schiff bases and the acceptor (tin atom), was consistent with that observed in earlier

reports.268,269 Furthermore, the signal attributable to the C-S moiety in the 13C NMR

spectra of the tin(IV) compounds was shifted upfield compared to their Schiff bases,

suggesting coordination of the sulphur atom to the tin centre. These observations

supported the FTIR analysis discussed above. 119Sn NMR spectroscopy provided further

validation of the geometries of 19-24. A sharp signal was observed in the range δ -441 to

-462 in the 119Sn NMR spectra of compounds 19-24, which strongly supported a six-

coordinated, distorted octahedral geometry around tin, comparable to reported

literature.100

145

5.3.4 UV-vis Absorption Spectral Analysis

Complete experimental and calculated UV-vis data are listed in Table A5.4.

Frontier MOs of 19-24 are presented in Figure 5.2. The experimental UV-vis spectra of

tin(IV) compounds showed a prominent absorption peak at 343-355 nm, in excellent

correlation with DFT calculations (334-348 nm, gas phase). The other experimental

absorption peak was observed at 415-427 nm, which indicated the presence of the

S→SnIV LMCT band. This was also verified by the excitations between 428-438 nm

observed from DFT calculations. The HOMOs in these compounds were primarily

located on the dithiocarbazate backbones, the phenyl ring of the aldehyde moieties and

the coordinating oxygen atoms. The LUMOs centred on the dithiocarbazate backbones,

and the phenyl rings of oVa or catechol. It is evident then that the UV-vis absorption

corresponds to a n→π* excitation, due to the presence of the electron lone pairs in the

azomethine nitrogen, thiolate sulphur and phenoxide oxygen atoms. The π→π*

transitions observed in all tin(IV) compounds corresponded to the electron delocalisation

around the aromatic rings.

5.3.5 X-ray Crystallography

Crystal structure determinations were achieved for the homoleptic compounds 20

and 21; selected geometric parameters are collected in Table 5.2. The crystallographic

asymmetric unit of 20 comprises two independent molecules, with the first shown in

Figure 5.3(a) and the second molecule shown in Appendix, Figure A5.1. For 21, one

molecule comprises the asymmetric unit, Figure 5.3(b). For the first independent

molecule of 20, the dinegative tridentate ligand coordinates the tin atom via thiolate-S,

phenoxide-O and imine-N atoms to establish five-membered Sn,S,C,N2 and six-

membered Sn,O,C3,N rings. Evidence for the presence of thiolate-sulphur atoms is found

146

in the elongation of the C1−S1 and C18−S3 bond lengths (Table 5.2) from 1.670(2) Å,

which is found in the most closely related acid molecule for which a crystallographic

analysis has been reported, i.e. the benzyl ester, rather than the 4-tolylCH2 ester.177

Concomitantly, the C1−N1 and C18−N3 bond lengths (Table 5.2) have also decreased

considerably from 1.338(2) Å in the acid.177 The conformation of the two tridentate

ligands is such that all like atoms are mutually trans in a distorted octahedral environment.

The major distortions from the ideal octahedral geometries are due to the acute chelate

angles in the five-membered rings, i.e. 78.88(7)° for S1−Sn−N2 and 78.94(7)° for the

S3−Sn−N4 angle.

Figure 5.2: Frontier MOs of (a) 19, (b) 20, (c) 21, (d) 22, (e) 23 and (f) 24.

147

The S1-five-membered chelate ring formed by the tridentate ligand is strictly planar

with the RMSD of the fitted atoms being 0.0036 Å. However, the S3-ring is less planar

with the RMSD being 0.0629 Å. A better description for the latter is an envelope in which

the tin atom lies 0.308(4) Å out of the least-squares plane defined by the remaining four

atoms (RMSD = 0.0060 Å). Envelope conformations also apply for the six-membered

chelate rings. For the O1-ring, the tin atom lies 0.579(4) Å out the plane through the five

remaining atoms. For the O3-ring the envelope is somewhat flattened with the tin atom

0.135(4) Å out of plane (RMSD = 0.0234 Å). The dihedral angle between the five-

membered chelate rings is 81.56(8)°.

Table 5.2: Selected geometric parameters (Å, °) for 20 and 21. Complex 20 20a 21 (x = 18, y = 19) (x = 18, y = 19) (x = 17, y = 18) Parameter Sn‒S1 2.4828(8) 2.4871(9) 2.4932(7) Sn‒S3 2.4961(9) 2.4898(8) 2.4873(7) Sn‒O1 2.025(2) 2.026(2) 2.0304(18) Sn‒O3 2.032(2) 2.032(2) 2.0242(18) Sn‒N2 2.198(2) 2.189(2) 2.169(2) Sn‒N4 2.189(2) 2.196(2) 2.173(2) N1‒N2 1.398(3) 1.399(3) 1.395(3) C1‒S1 1.753(3) 1.749(3) 1.744(3) C1‒S2 1.744(3) 1.743(3) 1.741(3) C1‒N1 1.288(4) 1.286(4) 1.289(3) C2‒N2 1.299(4) 1.309(4) 1.302(3) N3‒N4 1.393(3) 1.396(3) 1.397(3) C(x)‒S3 1.744(3) 1.754(3) 1.745(3) C(x)‒S4 1.743(3) 1.745(3) 1.751(2) C(x)‒N3 1.295(4) 1.289(4) 1.288(3) C(y)‒N4 1.307(4) 1.303(4) 1.304(3) S1‒Sn‒O1 162.46(6) 165.07(6) 164.51(5) S3‒Sn‒O3 164.42(6) 162.30(6) 165.46(5) N2‒Sn‒N4 170.18(10) 170.45(9) 177.76(8)

As seen from the overlay diagram in Figure 5.3(c), there are conformational

differences between the independent molecules of 20, at least with respect to the terminal

148

thioester residues. The data in Table 5.2 and Appendix Table A5.5 confirm the close

similarity between the independent molecules in terms of the tin-atom geometries.

Figure 5.3: Molecular structures of the molecules in (a) 20 (first independent molecule; the structure for the second independent molecule is shown in Appendix Figure A5.1) and (b) 21, showing atom labelling schemes and 50% displacement ellipsoids. (c) Overlay diagram of the molecules in 20 (red image for the first independent molecule), 20a (green, inverted second molecule) and 21 (blue). Molecules have been overlapped so the Sn,S1,N2 chelate rings are coincident.

149

The trends in the structure of 21 follow those established for 20, with the most

obvious difference relating to the coordination of the imine-N2 and N4 atoms. The angle

they subtend at the tin atom is wider by ca. 6-7°, and there is evidence to suggest the Sn–

N2, N4 bond lengths are marginally shorter than the equivalent bonds in the molecules of

20. Each of the chelate rings adopts an envelope conformation with data presented in

Appendix Table A5.5. The dihedral angle between the five-membered chelate rings in

21 is 81.77(5)°, i.e. within experimental error of the value computed for 20.

There are three related homoleptic tin(IV) compounds that have been structurally

characterised in the literature, namely the ethyl thioester and unsubstituted phenoxide

residue (Figure 5.4(a)),118 benzyl thioester, unsubstituted phenoxide residue and methyl

bound to the imine-carbon(Figure 5.4(b))270 and benzyl thioester with a ferrocenyl

substituent adjacent to the alkoxide-oxygen atom (Figure 5.4(c)).267 The contrasting and

curious feature of the literature structures is that the thiolate-sulphur atoms are mutually

cis, as are the alkoxide-oxygen atoms. The reasons for the different conformations

observed in 20 and 21, and those in the literature remain unclear.

In the absence of conventional hydrogen bonding, the crystals of 20 and 21 are

sustained by a variety of other non-covalent interactions; the geometric parameters

characterising these are included in the captions to the respective figures in the Appendix

Tables A5.6 and A5.7. In the molecular packing of 20, each of the independent molecules

form equivalent intermolecular contacts with the other independent molecule to sustain a

supramolecular layer in the ab-plane. These are imine-C2-H…S3(thiolate), imine-C19-

H…S1(thiolate), methoxybenzene-C7-H…O2(methoxy), methoxybenzene-C24-

H…O4(methoxy) and π(C3-C8)…π(C20-C25). The connections between layers along the

c-axis direction include methyl-C16a-H…O2a, methyl-C16-H…π(C10-C15) and

methoxy-C34-H…π(C10-C15) interactions. These occur between like-molecules and

150

hence, differentiate the independent molecules comprising the asymmetric unit in terms

of their supramolecular association. Images of the supramolecular association operating

in the crystal of 20 and intermolecular interactions are shown in Appendix Figure A5.2

and Table A5.6, respectively.

S NN

SO

SNN

SO Sn

S NN

SO

SNN

SO Sn

S NH

NS

Sn O Fc

SHN

NS

OFc

Fc = ferrocene

(a) (b)

(c)

Figure 5.4: Homoleptic tin(IV) compounds that have been studied.

As seen in Appendix Figure A5.3, supramolecular layers in the ab-plane are also

formed in the crystal of 21 via a combination of methoxybenzene-C7–H…O4(methoxy),

methoxybenzene-C23–H…O2(methoxy), methylene-C25–H…S1(thiolate), methoxy-

C16–H…π(C10-C15) and π(C3-C8)…π(C3…C8) interactions. Layers inter-digitate along

the c-axis direction but, without directional interactions between them. The geometric

characteristic of intermolecular interactions is supplemented in Appendix Table A5.7.

151

5.3.6 In Vitro Cytotoxic Activity

The cytotoxic assays of 19, 21, 22, 23 and 24 were carried out against a panel ten

of cancer cells (HT29, U87, SJ-G2, MCF-7, A2780, H460, A431, Du145, BE2-C, MIA)

and one normal breast cell line (MCF-10A). Cytotoxicity for 20 could not be studied due

to its insolubility in 100% DMSO at 1 mM. Cisplatin was used as a positive control and

DMSO was used as a negative control to monitor the experiments. Table 5.3 shows that

the cytotoxicity of these tin(IV) compounds and their respective Schiff bases, 19, 21, 22,

23 and 24 exhibited moderate activity against all the tested cancer cell lines. The

cytotoxicities of these compounds against the HT29, MCF7 and MIA cell lines were

higher than cisplatin. Taking the Schiff bases into account, 19-24 were equipotent to their

corresponding Schiff bases, with the exception of 22. Compound 22 demonstrated higher

potency than their corresponding Schiff bases against the HT29, U87, MCF-7, A431 and

Du145 cell lines. The mechanism of action of 22 needs further in-depth investigation. In

general, for compounds 19, 20, 21, 23 and 24, it can be inferred that the presence of tin

does not influence cytotoxicity. This is possibly due to the bulkiness of the compound

and the possibility that only the coordinated Schiff bases playing a role in binding to the

biomacromolecules.271

5.3.7 Molecular Docking Analysis

Molecular docking analyses of compounds 19, 21, 22, 23 and 24 with the DNA

duplex sequence were performed, and the lowest binding energy conformation of the

docked structure are summarised in Table 5.4. From the docked structures obtained, it

was clear that 19, 21, 22, 23 and 24 considered here were fitted with the dimensions of

the grooves of the DNA duplex sequence. From the docked structures obtained here, it

was proposed that compounds 19-24 fit well into the cytosine and guanine region in the

152

minor groove of the targeted DNA (Figure 5.4(a)) with a relative binding energy of -35.82

(1), -37.15 (2), -33.56 (3), -31.42 (4), -32.07 (5) and -31.67 (6) kJ mol-1. These compounds

were predicted to interact with the DNA strand, suggesting that the planar aromatic ring

of the Schiff bases interacts with the nitrogenous bases of DNA via π-π stacking

interactions. These interactions are stabilised by other interactions, such as van der Waals,

hydrogen bonding, hydrophobic and electrostatic interactions. These interactions of the

energetically favourable conformation of the docked model are visualised in Figure

5.5(b). Is it worth noting that these interactions are consistent with the fact that the

cytotoxicity is the result of the Schiff base, and not necessarily the tin.

153

Table 5.3: Summary of the in vitro cytotoxicity of tin(IV) compounds in several cell lines, determined by the MTT assay and expressed as a GI50 value with standard error. GI50 is the concentration of tin(IV) compounds at which cell growth is inhibited at 50% over 72 hours.

Compounds Growth inhibition concentration, GI50 (µM)

HT29 U87 MCF-7 A2780 H460 A431 Du145 BE2-C SJ-G2 MIA MCF-10A

1 3.9 ± 0.71 4.3 ± 0.30 2.7 ± 0.21 3.2 ± 0.15 4.3 ± 0.12 3.8 ± 0.033 4.1 ± 0.09 3.1 ± 0.00 3.0 ± 0.09 4.9 ± 0.37 3.1 ± 0.07

19 3.9 ± 0.53 4.2 ± 0.03 2.9 ± 0.07 3.3 ± 0.10 4.0 ± 0.09 3.8 ± 0.15 4.6 ± 0.17 3.0 ± 0.13 2.9 ± 0.12 4.2 ± 0.27 3.2 ± 0.21 2 8.0 ± 3.5 3.9 ± 0.00 3.5 ± 0.00 3.1 ± 0.40 5.5 ± 0.73 4.2 ± 0.74 5.4 ± 0.23 3.6 ± 0.27 3.3 ± 0.87 11 ± 2.3 3.5 ± 0.23 20 nd nd nd nd nd nd nd nd nd nd nd

3 2.2 ± 0.033 3.1 ± 0.26 2.5 ± 0.12 3.0 ± 0.06 3.3 ± 0.21 3.0 ± 0.15 3.4 ± 0.30 2.8 ± 0.20 2.4 ± 0.26 4.5 ± 0.25 3.0 ± 0.067

21 2.5 ± 0.30 3.1 ± 0.13 2.5 ± 0.23 2.6 ± 0.12 3.0 ± 0.12 2.5 ± 0.32 3.5 ± 0.26 2.1 ± 0.00 1.4 ± 0.36 2.7 ± 0.21 2.8 ± 0.19

10 10 ± 3.5 20 ± 5.2 10 ± 3.1 1.8 ± 1.2 8.7 ± 1.9 14 ± 1.5 23 ± 3.6 0.38 ± 0.07 6.4 ± 2.2 5.5 ± 1.1 3.7 ± 0.50

22 4.4 ± 0.83 9.3 ± 1.8 4.4 ± 1.9 2.7 ± 0.20 4.7 ± 0.23 3.7 ± 0.25 6.4 ± 1.4 1.7 ± 0.33 1.9 ± 0.36 4.6 ± 0.97 3.3 ± 0.19

11 2.8 ± 0.033 4.0 ± 0.43 2.8 ± 0.067 2.8 ± 0.20 3.7 ± 0.10 3.7 ± 0.20 4.2 ± 0.59 3.2 ± 0.30 3.3 ± 0.30 6.6 ± 1.7 2.9 ±

0.033 23 3.3 ± 0.26 4.8 ± 0.52 3.0 ± 0.15 3.3 ± 0.26 4.2 ± 0.12 3.7 ± 0.32 6.0 ± 0.58 2.7 ± 0.43 3.0 ± 0.27 11 ± 0.88 2.8 ± 0.33 12 2.1 ± 0.35 2.7 ± 0.47 2.3 ± 0.32 2.2 ± 0.25 2.6 ± 0.27 2.4 ± 0.26 2.8 ± 0.15 2.0 ± 0.23 2.2 ± 0.47 2.2 ± 0.64 2.9 ± 0.21

24 2.1 ± 0.44 3.3 ± 0.23 2.2 ± 0.20 2.4 ± 0.38 2.1 ± 0.28 2.8 ± 0.23 3.8 ± 0.52 2.1 ± 0.17 1.4 ± 0.23 2.5 ± 0.64 2.8 ± 0.088

Cisplatin 11.0 ± 2.0 4.0 ± 1.0 6.5 ± 0.8 1.0 ± 0.1 0.9 ± 0.2 2.4 ± 0.3 1.2 ± 0.1 1.9 ± 0.2 0.4 ± 0.1 8.0 ± 1.0 nd GI50 (μM): (the colors indicate) = 0.1–0.99 (strong); = 1.0–9.9 (moderate); = 10-100 (weak); = nd (not determined)

154

Table 5.4: Molecular docking data for the tin(IV) compounds with B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2.

Compound

Final intermolecular energy, kcal mol-1 Final total internal

energy (2), kcal mol-1

Torsional free energy

(3), kcal mol-1

Unbound system's energy

[=(2)] (4), kcal mol-1

Estimated free energy of binding

[(1)+(2)+(3)-(4)], kcal mol-1

Estimated free energy of binding,

kJ mol-1

vdW + Hbond + desolv. Energy

Electrostatic energy

Total (1)

19 10.82 -0.15 -10.98 -1.36 2.39 -1.36 -8.56 -35.82 21 -10.3 -0.1 -10.4 -1.29 2.39 -1.29 -8.02 -33.56 22 -9.86 -0.03 -9.89 -3.42 2.39 -3.42 -7.51 -31.42 23 -10.02 -0.03 -10.06 -3.12 2.39 -3.12 -7.67 -32.09 24 -9.78 -0.18 -9.96 -3.06 2.39 -3.06 -7.57 -31.67

155

Figure 5.5: (a) Schematic representation of 19 (cyan), 21 (yellow), 22 (blue), 23 (grey) and 24 (green) that fit well in the grooves of the DNA strand obtained by docking simulations. The two double-stranded DNA comprised of the phosphate deoxyribose backbone (grey) with guanine (green), cytosine (purple), adenine (DA, red) and thymine (cyan). (b) Molecular interactions of 19, 21, 22, 23 and 24 within the grooves of double stranded DNA residues. Hydrogen bond, electrostatic and hydrophobic interactions are depicted by green, orange and pink dot lines, respectively.

156

5.4 Conclusions

In summary, six octahedral tin(IV) compounds were synthesised by the

condensation of tin(II) chloride with dithiocarbazate Schiff bases (1, 2, 3, 10, 11 and 12).

The oxidation of tin(II) to tin(IV) occurred as two dinegatively charge tridentate ONS

Schiff bases coordinated the tin centre with general formulae Sn(L)2 (L = 1, 2, 3, 10, 11

and 12). X-ray crystallography indicated that the dinegative tridentate ligands of 20 and

21 coordinated to the tin atoms via the thiolate-S, phenoxide-O and imine-N atoms,

leading to octahedral geometries in which the like-atoms were mutually trans. The in

vitro cytotoxicity against a panel of cancer cell lines viz., HT29, U87, SJ-G2, MCF-7,

A2780, H460, A431, Du145, BE2-C and MIA cancer cell lines revealed that compounds

19, 21, 22, 23 and 24 showed moderate cytotoxicity, similar to that of their respective

Schiff bases. However, compound 22 exhibited a higher potency against HT29, U87,

MCF-7, A431 and Du145 cells as compared to its Schiff base. Molecular docking predicts

that the tin(IV) compounds discussed in this chapter have a higher affinity wards guanine–

cytosine rich regions than the adenine–thymine rich regions of DNA due to their bulky

octahedral structures. The binding interactions of tin(IV) compounds-DNA were

stabilised by hydrogen, electrostatic and hydrophobic interactions.

Chapter 2, 3 and 4 demonstrated the synthesis, structural evaluation and

cytotoxicity studies of dithiocarbazate Schiff bases and their tin(IV) compounds, where

all the compounds showed good cytotoxic activity, particularly diphenyltin(IV)

compounds. From the cytotoxicity results obtained here, it is hypothesised that this

activity could be enhanced by modification of the backbone of dithiocarbazate (see Figure

5.6).

157

S NH

SN

RNH

NH

SN

HOR1

HOR1

R

R = o-CH3; R1 = OCH3 (1)R = p-CH3; R1 = OCH3 (2)R = H; R1 = OCH3 (3)R = o-CH3; R1 = OH (10)R = p-CH3; R1 = OH (11)R = H; R1 = OH (12)

R = CH3; R1 = OCH3 (25)R = C6H3; R1 = OCH3 (26)R = CH3; R1 = OH (27)R = C6H3; R1 = OH (28)

(a) (b)

4

Figure 5.6: The small modification (blue) of backbone of (a) dithiocarbazate Schiff bases to produce (b) thiosemicarbazone Schiff bases.

This hypothesis will be tested in Chapter 6, which reports the synthesis, structural

and cytotoxic studies of thiosemicarbazone Schiff bases (25-28) and their tin(IV)

compounds (29-40). Previous studies reported that the variation of the groups on the N(4)

atom (see Figure 5.6b) of the thiosemicarbazone Schiff bases affected the biological

action.272 This motivated further investigations on the relationship between

thiosemicarbazone Schiff base structures and cytotoxic activity, by changing the

substituents at N(4) position with methyl and phenyl groups, which will be discussed in

Chapter 6.

158

CHAPTER 6

TIN(IV) COMPOUNDS OF THIOSEMICARBAZONE SCHIFF BASES: SYNTHESIS, STRUCTURAL CHARACTERISATION AND IN VITRO

CYTOTOXICITY

6.1 Introduction

Thiosemicarbazones are an important class of compounds which have gained

attention owing to their remarkable biological and pharmacological properties.

Thiosemicarbazones differ from dithiocarbazates by replacing the sulphur atom with a

nitrogen atom attached to C=S (see Figure 6.1). One of the promising uses of

thiosemicarbazones is as anticancer agents. They play a key role in transporting and

directing molecules to the target site.98,119–121 In 1956, 2-formylpyridine

thiosemicarbazone was reported to exhibit a good effect on leukemic cells, but this

compound was found to be highly toxic.273 Almost a decade later, French and Blanz

formulated a hypotheses about the mode of action of tridentate α(N)-heterocyclic

thiosemicarbazones, postulating that the modification of the ring system while retaining

the ligand pattern could improve the cytotoxicity effect and decrease toxicity.274 The main

factors that contributed to the activity were electron densities, substituents and geometry

of the thiosemicarbazones. The anticancer activity of thiosemicarbazones also depend on

the topology of the cancer cells, which renders the selectivity of the target compounds.

Nevertheless, the presence of metal ions, and specifically tin, in this study almost

systematically increased the cytotoxicity and lowered the toxicity. Presently, the main

mechanism of action related to thiosemicarbazones is RR inhibition, which was discussed

in Chapter 1. Furthermore, thiosemicarbazones were also found to inhibit RNA-

dependent DNA polymerase and dihydrofolate reductase.275

159

S NH

SN

RNH

NH

SN

HOR1

HOR1

R

R = o-CH3; R1 = OCH3 (1)R = p-CH3; R1 = OCH3 (2)R = H; R1 = OCH3 (3)R = o-CH3; R1 = OH (10)R = p-CH3; R1 = OH (11)R = H; R1 = OH (12)

R = CH3; R1 = OCH3 (25)R = C6H3; R1 = OCH3 (26)R = CH3; R1 = OH (27)R = C6H3; R1 = OH (28)

(a) (b)

Figure 6.1: The structures of (a) dithiocarbazate and (b) thiosemicarbazone Schiff bases.

While the synthesis of thiosemicarbazone Schiff bases containing 4-methyl-3-

thiosemicarbazide or 4-phenyl-3-thiosemicarbazide with oVa or catechol have been

reported,276–280 the cytotoxic evaluation for these compounds and their tin(IV) compounds

remains largely unexplored. Following the discussion in previous chapters on the design

and synthesis of dithiocarbazate Schiff bases and their tin(IV) compounds, it was

hypothesised that their activity could be enhanced by modification of the dithiocarbazate

backbone. This hypothesis is investigated in this chapter, which presents the synthesis,

structural characterisation and cytotoxic activities of oVa and catechol thiosemicarbazone

Schiff bases and their tin(IV) compounds.

6.2 Experimental

6.2.1 Materials and Physical Measurements

All solvents and reagents were of analytical reagent grade and used without further

purification as described in Chapter 2 (§2.2.1), Chapter 3 (§3.2.1) and Chapter 4 (§4.2.1).

Chemicals: 4-methyl-3-thiosemicarbazide (Sigma Aldrich, St. Louis, MO, USA), 4-

160

phenyl-3-thiosemicarbazide (Sigma Aldrich, St. Louis, MO, USA). The instrumentation

used are as outlined in Chapter 2 (§2.2.3) and Chapter 3 (§3.2.3).

6.2.2 Synthesis

6.2.2.1 Schiff Bases

6.2.2.1.1 2-(2-hydroxy-3-methoxybenzylidene)-N-methylhydrazine

carbothioamide (25) and 2-(2-hydroxy-3-methoxybenzylidene)-N-phenylhydrazine

carbothioamide (26)

Compounds 25 and 26 were prepared according to the procedure described in the

literature,123,278 without reflux. 4-Methylthiosemicarbazide (1.05 g, 10 mmol)/ 4-

phenylthiosemicarbazide (1.67 g, 10 mmol) was dissolved in 40 cm3 of methanol with

stirring and heating over a period of 30 min. o-Vanillin (1.52 g, 10 mmol) in 10 cm3 of

methanol was added to the thiosemicarbazide solution and was stirred at room

temperature for 4 h. Upon cooling, a crystalline product began to separate and the product

was filtered, washed with cold methanol and dried in a desiccator over anhydrous silica

gel.

2-(2-hydroxy-3-methoxybenzylidene)-N-methylhydrazinecarbothioamide (25)

White crystalline solid. Yield: 92 %. Melting point: 242-243°C. Analysis calculated

for C10H13N3O2S: C, 50.19; H, 5.48; N, 17.56. Found: C, 49.93; H, 5.38; N, 17.22 %. FT-

IR (ATR, cm-1): 3337 v(OH), 3304 v(NH), 1610 v(C=N), 1109 v(N-N), 1037 v(C=S). 1H

NMR (DMSO-d6) δ (ppm.): 2.99 (d, 3H, CH3), 3.79 (s, 3H, O-CH3), 6.93-7.54 (multiplet,

3H, Ar-H), 8.37 (s, 1H, CH), 8.39 (q, 1H, C(=S)-NH), 9.18 (s, 1H, OH), 11.42 (s, 1H,

NH-N). 13C NMR (DMSO-d6) δ (ppm.): 30.8 (N-CH3), 56.4 (O-CH3), 113.1, 118.2,

119.2, 121.2, 139.6, 146.7 (6 x aromatic-C), 148.4 (C=N), 177.6 (C=S).

161

2-(2-hydroxy-3-methoxybenzylidene)-N-phenylhydrazinecarbothioamide (26)

White crystalline solid. Yield: 90 %. Melting point: 209-210°C. Analysis calculated

for C15H15N3O2S: C, 59.78; H, 5.02; N, 13.94. Found: C, 59.99; H, 5.15; N, 13.80 %. FT-

IR (ATR, cm-1): 3300 v(NH), 1609 v(C=N), 1103 v(N-N), 908 v(C=S). 1H NMR (DMSO-

d6) δ (ppm.): 3.80 (s, 3H, O-CH3), 6.77-7.69 (multiplet, 6H, Ar-H), 9.26 (s, 1H, CH), 8.50

(s, 1H, C(=S)-NH), 10.02 (s, 1H, OH), 11.78 (s, 1H, NH-N). 13C NMR (DMSO-d6) δ

(ppm.): 56.3 (O-CH3), 113.4, 118.8, 119.5, 121.2, 125.4, 126.4, 128.6, 139.7, 140.4, 146.6

(10 x aromatic-C), 148.6 (C=N), 176.1 (C=S).

6.2.2.1.2 2-(2,3-dihydroxybenzylidene)-N-methylhydrazinecarbothioamide

(27) and 2-(2,3-dihydroxybenzylidene)-N-phenylhydrazine carbothioamide (28)

Compounds 27 and 28 were prepared according to the procedure described in the

literature.280,281 A 25 cm3 ethanolic solution of 2,3-dihydroxybenzaldehyde (1.38 g, 10

mmol) was added to an equimolar ethanolic solution (10 cm3) of 4-methyl-3-

thiosemicarbazide (1.05 g, 10 mmol) / 4-phenyl-3-thiosemicarbazide (1.67 g, 10 mmol).

The mixture was heated (78°C) for 3 hours and the title compound was filtered. The title

compound was then recrystallised from methanol to remove all the impurities and kept in

a desiccator over anhydrous silica gel.

2-(2,3-dihydroxybenzylidene)-N-methylhydrazinecarbothioamide (27)

Pale yellow solid. Yield: 83 %. Melting point: 231-232°C. Analysis calculated for

C9H11N3O2S: C, 47.99; H, 4.92; N, 18.65. Found: C, 46.88; H, 4.85; N, 18.38 %. FT-IR

(ATR, cm-1): 3418 v(OH), 3140 v(NH), 1601 v(C=N), 1112 v(N-N), 1035 v(C=S). 1H

NMR (DMSO-d6) δ (ppm.): 3.00 (d, 3H, CH3), 6.67-7.38 (multiplet, 3H, Ar-H), 9.02 (s,

1H, CH), 8.37, 8.38 (s, 2H, OH), 9.49 (s, 1H, C(=S)-NH), 11.40 (s, 1H, NH-N). 13C NMR

162

(DMSO-d6) δ (ppm.): 31.3 (N-CH3), 116.7, 117.4, 119.4, 121.5, 140.1, 145.6 (6 x

aromatic-C), 146.0 (C=N), 178.0 (C=S).

2-(2,3-dihydroxybenzylidene)-N-phenylhydrazine carbothioamide (28)

Pale yellow solid. Yield: 70 %. Melting point: 215-216°C. Analysis calculated for

C14H13N3O2S: C, 58.52; H, 4.56; N, 14.62. Found: C, 57.61; H, 4.69; N, 14.67 %. FT-IR

(ATR, cm-1): 3443 v(OH), 3129 v(NH), 1597 v(C=N), 1047 v(N-N), 1029 v(C=S). 1H

NMR (DMSO-d6) δ (ppm.): 6.67-7.57 (multiplet, 8H, Ar-H), 8.49 (s, 1H, CH), 8.96, 9.52

(s, 2H, OH), 10.01 (s, 1H, C(=S)-NH), 11.75 (s, 1H, NH-N). 13C NMR (DMSO-d6) δ

(ppm.): 117.1, 117.9, 119.5, 121.3, 125.6, 126.0, 128.5, 139.6, 141.3, 145.9 (12 x

aromatic-C), 146.0 (C=N), 176.0 (C=S).

6.2.2.2 Tin(IV) Compounds derived from 25 and 27

To a solution of 25 (0.24 g, 1 mmol)/ 27 (0.23 g, 1 mmol) in 100 cm3 of methanol,

potassium hydroxide (0.11 g, 2 mmol) was added and the mixture was stirred and heated

for 30 minutes in methanol. 1 mmol of tin salt (Ph2SnCl2 (0.34 g)/ Me2SnCl2 (0.22 g)/

SnCl2 (0.19 g)) was added to the mixture and heated for 2 hours until the solution reduced

by half. The mixture was filtered while hot and then the filtrate was placed in the freezer

until bright yellow solids formed. The solid residue obtained was recrystallised from

methanol.

Diphenyltin(IV) compound containing 2-(2-hydroxy-3-methoxybenzylidene)-N-

methylhydrazinecarbothioamide (29)

Bright yellow solid. Yield: 74 %. Melting point: 138-139°C. Analysis calculated

for C22H21N3O2SSn: C, 51.79; H, 4.15; N, 8.24. Found: C, 51.03; H, 4.23; N, 8.16 %. FT-

163

IR (ATR, cm-1): 3299 v(N-H), 1596 v(C=N), 1066 v(N-N), 973 v(C=S). 1H NMR (CDCl3)

δ (ppm.): 3.01 (d, 3H, N-CH3), 3.96 (s, 3H, O-CH3), 7.96 (s, 1H, CH), 7.36-8.12 (m, 13H,

Ar-H), 8.59 (s, 1H, NH). 13C NMR (CDCl3) δ (ppm.): 29.8 (NH-CH3), 56.5 (O-CH3),

115.7, 116.7, 117.2, 125.1, 128.6, 129.9, 135.9, 142.5, 151.6 (aromatic-C), 157.1 (C=N),

160.3 (S-C-S). 119Sn NMR (CDCl3) δ (ppm.): -236.2.

Dimenyltin(IV) compound containing 2-(2-hydroxy-3-methoxybenzylidene)-N-

methylhydrazinecarbothioamide (30)

Bright yellow solid. Yield: 42 %. Melting point: 164-168°C. Analysis calculated

for C12H17N3O2SSn: C, 37.33; H, 4.44; N, 10.88. Found: C, 39.00; H, 4.76; N, 11.00 %.

FT-IR (ATR, cm-1): 3222 v(N-H), 1590 v(C=N), 1066 v(N-N), 973 v(C=S). 1H NMR

(CDCl3) δ (ppm.): 0.91 (s, 6H, Sn-CH3), 2.97 (d, 3H, N-CH3), 3.94 (s, 3H, O-CH3), 7.26

(s, 1H, CH), 6.66-6.87 (m, 3H, Ar-H), 8.55 (s, 1H, NH). 13C NMR (CDCl3) δ (ppm.): 8.7

(Sn-CH3), 31.3 (NH-CH3), 56.4 (O-CH3), 115.7, 115.8, 117.8, 118.5, 125.8, 151.3

(aromatic-C), 156.5 (C=N), 178.0 (S-C-S). 119Sn NMR (DMSO-d6) δ (ppm.): -154.4.

Tin(IV) compound containing 2-(2-hydroxy-3-methoxybenzylidene)-N-

methylhydrazinecarbothioamide (31)

Compound 31 was prepared following the same procedure as described for 29,

using 25 (0.48 g, 2 mmol). Bright yellow solid. Yield: 31 %. Melting point: 118-119°C.

Analysis calculated for C20H22N6O4S2Sn: C, 40.49; H, 3.74; N, 14.17. Found: C, 40.50;

H, 3.33; N, 14.17 %. FT-IR (ATR, cm-1): 3308 v(N-H), 1590 v(C=N), 1066 v(N-N), 973

v(C=S). 1H NMR (DMSO-d6) δ (ppm.): 2.36 (d, 6H, N-CH3), 3.82 (s, 6H, O-CH3), 7.60

(s, 2H, CH), 6.67-7.31 (m, 6H, Ar-H), 8.88 (s, 2H, NH). 13C NMR (DMSO-d6) δ (ppm.):

19.3 (NH-CH3), 56.7 (O-CH3), 105.4, 116.7, 117.4, 126.5, 128.6, 129.0, 130.7, 130.8,

164

134.9, 149.7, 151.8, 158.3 (aromatic-C), 163.7 (C=N), 170.6 (S-C-S). 119Sn NMR

(DMSO-d6) δ (ppm.): -354.2.

Diphenyltin(IV) compound containing 2-(2,3-dihydroxybenzylidene)-N-

methylhydrazinecarbothioamide (32)

Yellow solid. Yield: 49 %. Melting point: 186-192°C. Analysis calculated for

C21H19N3O2SSn: C, 50.83; H, 3.86; N, 8.47. Found: C, 48.06; H, 4.27; N, 8.02 %. FT-IR

(ATR, cm-1): 1593 v(C=N), 1006 v(N-N), 953 v(C=S). 1H NMR (DMSO-d6) δ (ppm.):

3.00 (s, 3H, NH-CH3), 8.57 (s, 1H, CH), 6.37-8.20 (m, 13H, Ar-H), 11.27 (s, 1H, NH).

13C NMR (DMSO-d6) δ (ppm.): 30.9 (NH-CH3), 112.8, 113.7, 115.7, 118.4, 128.3, 129.2,

136.0, 136.1, 141.1, 145.4, 153.3 (aromatic-C), 154.3 (C=N), 177.0 (S-C-S). 119Sn NMR

(DMSO-d6) δ (ppm.): -227.0.

Dimethyltin(IV) compound containing 2-(2,3-dihydroxybenzylidene)-N-

methylhydrazinecarbothioamide (33)

Yellow solid. Yield: 32 %. Melting point: 223-225°C. Analysis calculated for

C11H15N3O2SSn: C, 35.51; H, 4.06; N, 11.29. Found: C, 35.87; H, 3.86; N, 11.35 %. FT-

IR (ATR, cm-1): 1596 v(C=N), 1006 v(N-N), 951 v(C=S). 1H NMR (DMSO-d6) δ (ppm.):

0.62 (s, 6H, Sn-CH3), 3.00 (d, 3H, N-CH3), 6.32-7.10 (m, 3H, Ar-H), 8.38 (s, 1H, CH),

11.30 (s, 1H, NH). 13C NMR (DMSO-d6) δ (ppm.): 6.8 (Sn-CH3), 30.7 (NH-CH3), 112.6,

113.7, 115.6, 118.6, 140.7, 153.9 (aromatic-C), 155.0 (C=N), 177.2 (S-C-S). 119Sn NMR

(DMSO-d6) δ (ppm.): -122.9.

Tin(IV) compound containing 2-(2,3-dihydroxybenzylidene)-N-

methylhydrazinecarbothioamide (34)

165

Compound 34 was prepared following the same procedure as described for 29,

using 27 (0.46 g, 2 mmol). Orange solid. Yield: 79 %. Melting point: >300°C. Analysis

calculated for C18H18N6O4S2Sn: C, 38.25; H, 3.21; N, 14.87. Found: C, 36.62; H, 2.92;

N, 14.21 %. FT-IR (ATR, cm-1): 1585 v(C=N), 993 v(N-N), 951 v(C=S). 1H NMR

(DMSO-d6) δ (ppm.): 2.83 (d, 6H, N-CH3), 6.65-7.52 (m, 6H, Ar-H), 8.34 (s, 1H, OH),

8.77 (s, 2H, CH), 11.27 (s, 1H, NH). 13C NMR (DMSO-d6) δ (ppm.): 30.6 (NH-CH3),

113.8, 114.3, 116.7, 118.9, 140.0. 150.1 (aromatic-C), 150.6 (C=N), 177.5 (S-C-S). 119Sn

NMR (DMSO-d6) δ (ppm.): -519.4.

6.2.2.3 Tin(IV) Compounds derived from 26

Compound 26 (0.30 g, 1 mmol)/ 28 (0.29 g, 0.001 mol) was dissolved in methanol

(100 cm3) and triethylamine (Et3N) (0.28 cm3, 2 mmol) was added dropwise to the

solution of 26 or 28. The mixture was heated for about 2 hours until the solution was

reduced by half. Next 1 mmol of tin salt (Ph2SnCl2 (0.34 g)/ Me2SnCl2 (0.22 g)/ SnCl2

(0.19 g)) was added to the mixture. The mixture was heated (64°C) for about 2 hours and

filtered while hot to remove triethylamine salt and the filtrate was kept at room

temperature until the bright yellow product formed.

Diphenyltin(IV) compound containing 2-(2-hydroxy-3-methoxybenzylidene)-N-

phenylhydrazinecarbothioamide (35)

Yellow crystals. Yield: 73 %. Melting point: 205-207°C. Analysis calculated for

C27H23N3O2SSn: C, 56.67; H, 4.05; N, 7.34%. Found: C, 57.53; H, 4.26; N, 7.87%. FT-

IR (ATR, cm-1): 3331 v(N-H), 1586 v(C=N), 1075 v(N-N), 832 v(C=S). 1H NMR (CDCl3)

δ (ppm.): 3.96 (s, 3H, O-CH3), 7.99 (s, 1H, CH), 6.69-7.56 (m, 18H, Ar-H), 8.70 (s, 1H,

NH). 13C NMR (CDCl3) δ (ppm.): 56.7 (O-CH3), 115.2, 116.2, 116.9, 119.4, 120.7, 123.4,

166

124.1, 125.4, 128.7, 128.9, 130.0, 135.9, 139.3, 142.1, 148.3, 149.7, 151.7 (aromatic-C),

162.5 (C=N), 164.8 (S-C-S). 119Sn NMR (CDCl3) δ (ppm.): -241.6.

Dimethyltin(IV) compound containing 2-(2-hydroxy-3-methoxybenzylidene)-N-

phenylhydrazinecarbothioamide (36)

Yellow crystals. Yield: 64 %. Melting point: 176-179°C. Analysis calculated for

C17H19N3O2SSn: C, 45.56; H, 4.27; N, 9.38%. Found: C, 45.86; H, 4.40; N, 7.08%. FT-

IR (ATR, cm-1): 3294 v(N-H), 1577 v(C=N), 1059 v(N-N), 824 v(C=S). 1H NMR (CDCl3)

δ (ppm.): 3.85 (s, 3H, O-CH3), 7.53 (s, 1H, CH), 6.69-7.32 (m, 18H, Ar-H), 8.65 (s, 1H,

NH). 13C NMR (CDCl3) δ (ppm.): 6.5 (Sn-CH3), 56.2 (O-CH3), 115.3, 116.6, 116.7,

120.5, 123.3, 125.4, 128.9, 139.4, 151.3, 156.8 (aromatic-C), 162.5 (C=N), 163.9 (S-C-

S). 119Sn NMR (CDCl3) δ (ppm.): -114.8.

Tin(IV) compound containing 2-(2-hydroxy-3-methoxybenzylidene)-N-

phenylhydrazinecarbothioamide (37)

Compound 37 was prepared following the same procedure as described for 35,

using 26 (0.60 g, 2 mmol). Yellow solid. Yield: 50 %. Melting point: 293-294°C. Analysis

calculated for C30H26N6O4S2Sn: C, 50.23; H, 3.65; N, 11.71%. Found: C, 49.85; H, 3.73;

N, 11.60%. FT-IR (ATR, cm-1): 3303 v(N-H), 1581 v(C=N), 1063 v(N-N), 824 v(C=S).

1H NMR (DMSO-d6) δ (ppm.): 3.58 (s, 6H, O-CH3), 9.08 (s, 2H, CH), 6.80-7.73 (m, 16H,

Ar-H), 9.70 (s, 2H, NH). 13C NMR (DMSO-d6) δ (ppm.): 56.4 (O-CH3), 117.9, 118.2,

121.0, 123.4, 126.8, 129.2, 140.4, 151.4, 154.8 (aromatic-C), 160.3 (C=N), 162.2 (S-C-

S). 119Sn NMR (DMSO-d6) δ (ppm.): -451.1.

167

6.2.2.4 Tin(IV) Compounds derived from 28

Compound 28 (0.29 g, 0.001 mol) was dissolved in methanol (100 cm3) and

potassium hydroxide (0.11 g, 2 mmol) was added dropwise to the solution of 28. The

mixture was refluxed for about 30 minutes, where the colour changed from light yellow

to orange. Next 1 mmol of tin salt (Ph2SnCl2 (0.34 g)/ Me2SnCl2 (0.22 g)/ SnCl2 (0.19 g))

was added to the mixture. The mixture was refluxed for 6 hour and filtered while hot to

remove triethylamine salt and the filtrate was kept at room temperature until the product,

an orange precipitate, formed.

Diphenyltin(IV) compound containing 2-(2,3-dihydroxybenzylidene)-N-

phenylhydrazinecarbothioamide (38)

Orange solid. Yield: 71 %. Melting point: 133-137 °C. Analysis calculated for

C26H21N3O2SSn: C, 55.94; H, 3.79; N, 7.53%. Found: C, 56.30; H, 3.99; N, 7.42%. FT-

IR (ATR, cm-1): 1587 v(C=N), 998 v(N-N), 957 v(S-C-S). 1H NMR (DMSO-d6) δ (ppm.):

6.57-9.52 (m, 18H, Ar-H), 9.87 (s, 2H, CH), 11.69 (s, 2H, NH). 13C NMR (DMSO-d6) δ

(ppm.): 120.5, 120.7, 125.4, 125.5, 125.6, 125.7, 128.5, 128.6, 128.7, 128.8, 128.9, 129.0,

129.1, 129.5, 134.6, 135.2, 135.5, 136.3, 136.4, 136.7, 139.7 (aromatic-C), 148.8 (C=N),

175.2 (S-C-S). 119Sn NMR (DMSO-d6) δ (ppm.): -327.8.

Dimethyltin(IV) compound containing 2-(2,3-dihydroxybenzylidene)-N-

phenylhydrazinecarbothioamide (39)

Orange solid. Yield: 42 %. Melting point: 153-156°C. Analysis calculated for

C16H17N3O2SSn: C, 44.27; H, 3.95; N, 9.68. Found: C, 44.32; H, 3.72; N, 9.90%. FT-IR

(ATR, cm-1): 1575 v(C=N), 1001 v(N-N), 917 v(S-C-S). 1H NMR (DMSO-d6) δ (ppm.):

0.63 (s, 6H, CH3), 8.51 (s, 1H, CH), 6.33-7.75 (m, 8H, Ar-H), 9.84 (s, 1H, OH). 13C NMR

168

(DMSO-d6) δ (ppm.): 7.1 (Sn-CH3), 110.2, 114.7, 128.3, 128.4, 128.7, 128.8, 135.3,

152.1, 152.7, 153.6 (aromatic-C), 153.8 (C=N), 167.5 (S-C-S). 119Sn NMR (DMSO-d6) δ

(ppm.): -103.3.

Tin(IV) of 2-(2,3-dihydroxybenzylidene)-N-phenylhydrazinecarbothioamide (40)

Compound 40 was prepared following the same procedure as described for 38,

using 28 (0.58 g, 2 mmol). Orange solid. Yield: 48 %. Melting point: >300 °C. Analysis

calculated for C28H22N6O4S2Sn: C, 48.78; H, 3.22; N, 12.19%. Found: C, 48.30; H, 2.97;

N, 12.52 %. FT-IR (ATR, cm-1): 1588 v(C=N), 1001 v(N-N), 939 v(S-C-S). 1H NMR

(DMSO-d6) δ (ppm.): 6.66-7.55 (m, 16H, Ar-H), 8.48 (s, 2H), 10.00 (s, 2H, CH), 11.74

(s, 2H, NH). 13C NMR (DMSO-d6) δ (ppm.): 117.1, 119.5, 121.2, 125.6, 125.7, 125.9,

126.0, 128.4, 128.5, 128.5, 139.6 (aromatic-C), 145.9, 146.0 (C=N), 175.7, 176.1 (S-C-

S). 119Sn NMR (DMSO-d6) δ (ppm.): -540.7.

6.2.3 DFT Calculations and Molecular Docking Simulations

DFT were performed using parameters detailed in Chapter 2 (§2.2.6) and Chapter

3 (§3.2.5.1), respectively. The coordinates for all compounds (25-40) were obtained after

minimisation of energy using the DFT method. The parameters for molecular docking

studies are described in Chapter 3 (§3.2.5.2).

6.2.4 MTT Assay

The protocol used for MTT assay followed that outlined in Chapter 2 (§2.2.6).

169

6.3 Results and discussion

6.3.1 Synthesis

The synthetic pathway of Schiff bases 25-28 and their tin(IV) compounds (29-40)

are described in Schemes 6.1 and 6.2, respectively. The Schiff bases were synthesised by

the condensation reaction of oVa or catechol and the corresponding thiosemicarbazide

(4-methyl-3-thiosemicarbazide and 4-phenyl-3-thiosemicarbazide) in alcoholic

solution.278,282,283 The Schiff bases were then reacted with Ph2SnCl2, Me2SnCl2 and SnCl2

in the presence of potassium hydroxide (KOH) or triethyamine by conventional or reflux

methods (outlined in §6.2.2). The isolated tin(IV) compounds were in good yields with

yellow or orange colours. The resultant tin(IV) compounds were soluble in most organic

solvents especially DMSO and DMF, except 31 which was not soluble in DMSO and

DMF. Due to the insolubility of 31, cytotoxic activity was not determined. The molar

conductance values of the tin(IV) compounds ranged from 0.88-7.85 Ω-1 cm2 mol-1, which

were well below than 25 Ω-1 cm2 mol-1 indicating that all of them were non-electrolytes

in nature and no counter ions were present around the lattice of the compounds.106

NH

NH

R1

SN

HOR2

NH

NH

R1

SNH3 HO

R2

O

MeOH/EtOH

heat and stir (3-4 hours)

R1 = CH3, R2 = OCH3 (25)R1 = C6H5, R2 = OCH3 (26)R1 = CH3, R2 = OH (27)R1 = C6H5, R2 = OH (28)

Scheme 6.1: Synthetic pathway of thiosemicarbazone Schiff bases 25-28.

170

NH

NR1

SN

OR2

Sn

NH

NR1

SN

OR2

Sn

NH

NR1

SN

OR2

Sn

HNN

R1

SN

OR2

NH

NH

R1

SN

HOR2

methanol

methanol

methanol

SnPh2Cl2

SnMe2Cl2

SnCl2

R1 = CH3, R2 = OCH3 (29)R1 = CH3, R2 = OH (32)R1 = C6H5, R2 = OCH3 (35)R1 = C6H5, R2 = OH (38)

R1 = CH3, R2 = OCH3 (30)R1 = CH3, R2 = OH (33)R1 = C6H5, R2 = OCH3 (36)R1 = C6H5, R2 = OH (39)

R1 = CH3, R2 = OCH3 (31)R1 = CH3, R2 = OH (34)R1 = C6H5, R2 = OCH3 (37)R1 = C6H5, R2 = OH (40)

Scheme 6.2: Synthetic pathway of tin(IV) compounds (29-40) of thiosemicarbazone Schiff bases

6.3.2 IR Spectral Analysis

The experimental and calculated frequencies in the infrared spectra of

thiosemicarbazone Schiff bases (25-28) and their tin(IV) compounds (29-40) were

recorded in the range of 4000-280 cm-1 and 4000-0 cm-1. All DFT vibrational frequencies

were scaled using a scaling factor of 0.9682.189 Their important frequencies and their

assignments are summarised in Appendix Table A6.1, for both experimental and

calculated frequencies.

In the spectra of 25 and 26, v(OH) was not observed in the spectra, which suggested

that the v(OH) band overlapped with the v(N-H) band due to the formation of hydrogen

bonding (NH...OH) between the two groups.284 Conversely, v(OH) was observed in the

spectra of 27 and 28 which were comparable with previous literature.65 As a result, the

loss of v(OH) upon complexation was difficult to assign by FTIR, due to the intra- and

intermolecular hydrogen bonding between the molecules. The v(N-H) band of

thiosemicarbazone Schiff bases disappeared upon complexation due to the deprotonation

171

of the NH group, and the involvement of this group in the coordination to the tin centre.

Furthermore, the FTIR spectra of 25, 26, 27 and 28 exhibited an intense band due to the

presence of the v(C=N)azomethine stretch at 1610, 1609, 1601 and 1597 cm-1, respectively.

This band shifted to lower frequencies in the spectra of tin(IV) compounds suggesting

that the coordination with the tin centre occurred via the azomethine nitrogen atom. The

v(C=S) and v(N-N) were also shifted to lower frequencies upon complexation, indicating

the coordination via the thiolate sulphur and azomethine nitrogen atoms, forming a five-

membered chelate ring.

6.3.3 NMR Spectroscopic Analysis

The 1H NMR and 13C NMR spectra of thiosemicarbazone Schiff bases 25-28 were

recorded in DMSO-d6 solution, while spectra of tin(IV) compounds 29-40 were recorded

in DMSO-d6/CDCl3 solution at room temperature. The assignments of the relevant signals

are compiled in Appendix Tables A6.2 and A6.3.

In the 1H NMR spectra of 25, 26, 27 and 28, a singlet signal observed at 11.42,

11.48, 9.49 and 10.01 ppm respectively indicated the presence of NH protons. This NH

proton signal was not observed in the spectra of tin(IV) compounds indicating that the

Schiff bases were coordinated to the tin atoms via the nitrogen donor atom. A proton

signal appeared at 9.18 and 10.02 ppm for 25 and 26, corresponding to a hydroxyl proton,

which disappeared in the 1H NMR spectra of the tin(IV) compounds indicating that the

coordination through the hydroxyl group.229 By contrast, the two signals for the hydroxyl

group at 8.32, 8.38 ppm (27) and 11.75, 10.01 ppm (28) disappeared upon complexation,

which indicated the presence of intra- and intermolecular hydrogen bonding between the

molecules.65,285

172

The 13C NMR spectra of 25, 26, 27 and 28 showed carbon signals at 177.6, 176.1,

178.0 and 176.0 ppm respectively, at the downfield region attributed to -S-C(=S)N. The

position of these carbon signals proved that 25, 26, 27 and 28 predominated as the thione

tautomers, even in DMSO-d6 solution. This signal was shifted upfield in the spectra of

the respective tin(IV) compounds, which indicated a decrease of electron density at the

carbon atom of -S-C(-S)N when sulphur was chelated to the tin centre. The C=N signal

was observed at 148.4 ppm (25), 148.6 ppm (26), 146.0 ppm (27) and 146.0 ppm (28)

appeared downfield, as the carbon atom was attached an electronegative atom. However,

this C=N signal was shifted downfield in the spectra of tin(IV) compounds, due to the

increased electron density around the atom upon complexation. The methoxy -CH3-

signal appeared upfield, at 56.4 ppm (25) and 56.3 ppm (26). The same carbon signal was

observed in the spectra of tin(IV) compounds indicating that the methoxy group did not

coordinate the tin centre.

119Sn NMR was also used to assist in the structural elucidation of compounds 29-

40. The 119Sn chemical shift strongly depends on the alkyl/aryl group attached to the tin

atom and electronegativity of the ligand coordinated to the tin atom as well as temperature

employed in the experiments (room temperature here). Theoretically, as the coordination

number increases, the 119Sn chemical shift moves towards the shielding region.230 The

spectra of 29-40 each showed one sharp signal which strongly supported that the tin(IV)

compounds contain single species of the tin atom. The 119Sn NMR values of penta-

coordinated diphenyl- (29, 32, 35 and 38) and dimethyltin(IV) (30, 33, 36 and 39)

compounds fell in the range of -227 to -328 and -103 to -154 ppm, respectively. The 119Sn

NMR values of hexa-coordinated tin(IV) compounds (31, 34, 37 and 40) were observed

in the range of -354 to -541 ppm, which supported the hypothesis that by increasing the

coordination sphere, the chemical shift decreased.151

173

6.3.4 Mass Spectral Analysis

Mass spectral data for 25-28 were recorded in DMSO and were found to be

consistent with the proposed formulation of the Schiff bases. Mass spectra displayed

prominent peaks at m/z 239, 301, 225 and 287 for Schiff bases 25, 26, 27 and 28,

respectively, which corresponded to [C10H13N3O2S]+, [C15H15N3O2S]+, [C9H11N3O2S]+

and [C14H13N3O2S]+. The mass spectra for 25, 26, 27 and 28 are supplied in Appendix

Figure A6.1.

6.3.5 UV-vis Absorption Spectroscopy

The experimental and calculated electronic spectra of the thiosemicarbazone Schiff

bases and their tin(IV) compounds in DMSO are tabulated in Appendix Table A6.4. The

prominent experimental electronic absorptions for the thiosemicarbazone Schiff bases 25-

28 were observed at 328-334 nm, which were closely correlated with B3LYP excitations

at 317-330 nm. The frontier MOs of the thiosemicarbazone Schiff bases and their tin(IV)

compounds are depicted in Figure 6.2, which shows that the HOMOs in these compounds

are dominated by sulphur and nitrogen, whereas their LUMOs are largely centred on the

thiosemicarbazone backbones and the oVa or catechol groups. Thus, UV-vis absorption

maxima are assigned as n→π* and π→π* excitations for 25-28. For the tin(IV)

compounds 29-40, the HOMOs are largely centred on the thiosemicarbazone Schiff base,

whereas the LUMOs are centred on the entire thiosemicarbazone Schiff base, except the

coordinating sulphur atom. The overlapped n→π* and π→π* transitions were observed

at 328-360 nm (experimental) and 347-380 nm (DFT). The intraligand absorption

corresponded to S→SnIV LMCT transition was observed at 401-417 nm (experimental)

and 395-424 nm (DFT), consistent with coordination of tin by the sulphur atoms.

174

Figure 6.2: Frontier MOs (a) 25, (b) 26, (c) 27, (d) 28, (e) 29, (f) 30, (g) 31, (h) 32, (i) 33, (j) 34, (k) 35, (l) 36, (m) 37, (n) 38, (o) 39 and (p) 40.

6.3.6 In Vitro Cytotoxic Activity

The four thiosemicarbazone Schiff bases and twelve tin(IV) compounds reported in

this chapter were screened for their cytotoxicity against a panel of ten cancer cell lines,

HT29, U87 and SJ-G2, MCF-7, A2780, H460, A431, Du145, BE2-C, MIA and one

175

normal cell line, MCF-10A (Table 6.1). The cytotoxicity of 31 was unable to be

determined due to its insolubility in 100% of DMSO at 1 mM concentration.

The cytotoxicity of oVa thiosemicarbazone Schiff bases (25 and 26) revealed an

increase in potency with the substituted methyl group at the α-nitrogen atom, where 25

exhibited 10 to 20 times higher antiproliferative activity than 26, 27 and 28 in all cancer

cell lines tested. Table 6.1 shows that the HT-29, A2780, A431, BE2-C and MIA cell

lines were more sensitive after treatment with 26. Compound 26 was approximately ~10-

100 times more potent than 3 (discussed in Chapter 2) against all cancer cell lines tested,

except for the Du145 cell line, which was more resistant towards 3. No obvious

cytotoxicity pattern was observed for the cytotoxicity of 26 which was similar to 3. In

contrast, 2,3-dihydroxybenzyl thiosemicarbazone Schiff bases (27 and 28) showed a

different pattern of cytotoxicity, which was attributed to the phenyl group attached to the

α-nitrogen. For instance, 28 was more active than the compound than 27 in all cancer cell

lines tested. Compound 27 showed poorer cytotoxicity at the 25 µM single point dose

evaluation pre-screening and was not selected for further GI50 determination, as it was

considered inactive.

The cytotoxicity studies of the tin(IV) compounds was comparable with the series

reported in Chapter 3 and 4.178 Diphenyltin(IV) compounds exhibited higher activities as

compared to their Schiff bases and other tin(IV) compounds, except 29, which exhibited

2.5-fold lower activity than its Schiff base (25) in MIA cells. Furthermore, 32 showed

promising cytotoxic activity with no specific selectivity against all cancer cell lines. The

cytotoxicity of dimethyltin(IV) compounds (30, 33, 36 and 39) and tin(IV) compounds

with the absence of organo group (31, 34, 37 and 40) exhibited no significant difference

as compared to their Schiff bases. It can be concluded that the presence of two phenyl

groups attached to the tin atom at the centre yields improved cytotoxicity against all

176

cancer cell lines considered here. Overall, the cytotoxicity results showed that HT29,

MCF-7, A2780, A431, BE2-C and MIA were more sensitive, whereas H460 and Du145

were less sensitive cancer cell lines to the Schiff bases and tin(IV) compounds in this

work.

6.3.7 DNA Binding Studies

In this chapter, DNA binding studies of compounds 27, 32, 33 and 34 were

investigated by absorption spectroscopy (Figure 6.3). These compounds were selected for

DNA binding evaluation in order to investigate the interactions of the inactive Schiff base

27 and its active tin(IV) compounds 32, 33 and 34 with DNA. The spectra were recorded

for a constant concentration of compounds (50 μM) with increasing concentrations of CT-

DNA. No significant changes were noted in the absorption spectra of DNA for 27 (Figure

6.3(i)) and 33 (Figure 6.3(iii)), which indicated that both compounds did not interact with

DNA. However, small changes were observed in the absorption spectra of DNA with

compounds 32 (Figure 6.3(ii)) and 34 (Figure 6.3(iv)), notably the decrease of the

absorption at λmax = 319 nm and 315 nm (i.e. hypochromism), respectively with absence

of any appreciable red shifts. Groove binding interactions is suggested as the most likely

mode of interaction, which was comparable to reported literature.286 The observed order

of hypochromism, viz 32 > 34 > 27 = 33, reflected a decrease in DNA binding affinity of

the tested compounds. This is consistent with the respective Kb values of 32 and 34, 4.62

x 104 and 7.51 x 103. The values obtained strongly supported that both compounds bound

to phosphate groups of the DNA surface.225 The DNA binding results correlated well with

cytotoxicity data, where compounds 32 and 34 showed more potency than compounds 27

and 33.

177

6.3.8 Molecular Docking Studies

As can be seen from the presented data in Table 6.2, there was an excellent

correlation between the growth inhibitions of cytotoxicity studies and free binding energy

of molecular docking simulations. A clear correlation was observed that 32 and 34 had

better activity than 27 and 33. Amongst the docked conformations, 32 showed the lowest

free binding energy (-31.42 kJ mol-1), which was in agreement with the DNA binding

studies discussed above. The compounds fitted well in the cytosine-guanine (G-C) rich

grooves (see Figure 6.4), and this region has a higher affinity than the adenine-thymine

(A-T) region. The G-C region is very important in terms of DNA stability, because the

guanine and cytosine are stabilised by three hydrogen bonds, and it has been suggested

that these regions could have a key role in anticancer activity.287 The compounds fitted in

the G-C region were stabilised by hydrogen, electrostatic and hydrophobic interactions.

Even though compounds 27 and 33 were not observed to have good DNA binding,

molecular docking simulations suggested that both compounds fitted at the grooves of

DNA. This can be explained by the fact that molecular docking simulations were

performed in rigid-rigid compounds-DNA conditions without explicit of water molecules

and other biological entities. Of note, significant potency was observed for compound 33,

where DNA binding studies revealed that 33 did not exclusively bind to the DNA. The

mechanism of action of 33 remains unknown and a subject for future research.

178

Table 6.1. In vitro cytotoxicity of tin(IV) compounds (29-40) derived thiosemicarbazone Schiff bases (25-28) in several cell lines, determined by the MTT assay and expressed as a GI50 value with standard error. GI50 is the concentration at which cell growth is inhibited by 50% over 72 hours.

Compounds

Growth inhibition concentration, GI50 (µM)

HT29 U87 MCF-7 A2780 H460 A431 Du145 BE2-C SJ-G2 MIA MCF-10A

25 0.09 ± 0.06

0.26 ± 0.11

0.12 ± 0.08

0.04 ± 0.07

0.42 ± 0.19

0.09 ± 0.06 10.2 ± 9.9 0.03 ±

0.02 0.15 ± 0.03

0.04 ± 0.02

0.43 ± 0.29

26 1.80 ± 0.22

3.5 ± 0.40

1.9 ± 0.46

2.0 ± 0.40

1.5 ± 0.27 2.4 ± 0.26 5.1 ± 0.58 1.0 ± 0.15 2.5 ± 0.77 2.1 ± 0.49 3.3 ±

0.26 27 > 25 > 25 > 25 > 25 > 25 > 25 > 25 > 25 > 25 > 25 > 25

28 0.29 ± 0.25

1.6 ± 0.29

0.03 ± 0.01

0.04 ± 0.02

1.0 ± 0.03 1.4 ± 0.43 2.1 ± 0.09 0.7 ± 0.51 0.8 ± 0.49 2.5 ± 1.2 1.5 ±

0.40

29 0.05 ± 0.02

0.14 ± 0.06

0.02 ± 0.00

0.02 ± 0.00

0.22 ± 0.06

0.07 ± 0.04

0.86 ± 0.42

0.02 ± 0.00

0.09 ± 0.04

0.11 ± 0.09

0.15 ± 0.08

30 0.30 ± 0.05 6.9 ± 5.1 0.13 ±

0.03 0.20 ± 0.01

1.3 ± 0.53

0.35 ± 0.12 5.0 ± 3.1 0.15 ±

0.03 0.48 ± 0.14

0.15 ± 0.03

1.0 ± 0.71

31 nd nd nd nd nd nd nd nd nd nd nd

32 0.15 ± 0.07

0.38 ± 0.10

0.10 ± 0.05

0.15 ± 0.04

0.60 ± 0.10

0.49 ± 0.31

0.50 ± 0.18

0.15 ± 0.04

0.42 ± 0.18

0.23 ± 0.07

0.34 ± 0.15

33 0.36 ± 0.07

2.2 ± 0.33

0.22 ± 0.04

0.26 ± 0.01

1.4 ± 0.67 1.2 ± 0.81 9.1 ± 3.4 1.5 ± 0.39 1.6 ± 1.2 0.79 ±

0.36 3.2 ± 0.33

34 0.09 ± 0.06 3.5 ± 3.3 0.09 ±

0.07 0.10 ± 0.08 >50 20.38 ±

15.1 25 ± 4.00 0.15 ± 0.11

0.22 ± 0.04

0.95 ± 0.21

0.81 ± 0.60

35 1.0 ± 0.35

0.57 ± 0.13

0.31 ± 0.13

0.27 ± 0.01

1.1 ± 0.03 1.1 ± 0.26 1.4 ± 0.20 0.29 ±

0.06 0.44 ± 0.11

0.35 ± 0.08

1.2 ± 0.41

36 1.8 ± 0.38

4.8 ± 0.75

1.6 ± 0.75

2.1 ± 0.43

2.0 ± 0.18 2.8 ± 0.46 4.8 ± 0.73 1.1 ± 0.20 1.3 ± 0.54 2.0 ± 0.88 3.2 ±

0.07

179

37 1.7 ± 0.74 3.7 ± 1.3 2.6 ±

0.48 2.6 ± 0.37

3.2 ± 0.15 2.6 ± 0.22 5.3 ± 0.67 1.8 ± 0.07 5.0 ± 1.97 3.3 ± 0.72 4.4 ±

0.60

38 0.19 ± 0.05

0.32 ± 0.04

0.12 ± 0.03

0.16 ± 0.04

0.53 ± 0.19

0.38 ± 0.19

0.20 ± 0.05

0.13 ± 0.02

0.20 ± 0.03

0.10 ± 0.02

0.37 ± 0.08

39 0.12 ± 0.03

0.40 ± 0.07

0.10 ± 0.03

0.21 ± 0.05

0.83 ± 0.37

0.27 ± 0.11

0.56 ± 0.27

0.20 ± 0.05

0.23 ± 0.03

0.17 ± 0.05

0.30 ± 0.02

40 0.18 ± 0.02

0.44 ± 0.06

0.17 ± 0.08

0.27 ± 0.023

0.47 ± 0.07

0.26 ± 0.03

0.42 ± 0.15

0.25 ± 0.05

0.26 ± 0.02

0.23 ± 0.07

0.34 ± 0.02

Cisplatin 11.0 ± 2.0 4.0 ± 1.0 6.5 ± 0.8 1.0 ± 0.1 0.9 ± 0.2 2.4 ± 0.3 1.2 ± 0.1 1.9 ± 0.2 0.4 ± 0.1 8.0 ± 1.0 nd

GI50 (μM): (the colors indicate) = 0.01-0.099 (very strong); = 0.1–0.99 (strong); = 1.0–9.9 (moderate); = 10-100 (weak); = nd (not determined)

180

Figure 6.3: Electronic absorption spectra of (i) 27, (ii) 32, (iii) 33 and (iv) 34. The arrow indicates the change in absorbance in tandem with increasing DNA concentration.

Figure 6.4: Molecular interactions of (a) 27, (b) 32, (c) 33 and (d) 34 within the grooves of double stranded DNA residues. Hydrogen bond, electrostatic and hydrophobic interactions are depicted by green, orange and pink dot lines, respectively.

181

Table 6.2: Molecular docking data for 27, 32, 33 and 34 with B-DNA (PDB ID: 1BNA) dodecamer d(CGCGAATTCGCG)2.

Compounds

Final intermolecular energy, kcal mol-1 Final total internal

energy (2), kcal mol-1

Torsional free energy

(3), kcal mol-1

Unbound system's energy

[=(2)] (4), kcal mol-1

Estimated free energy of binding

[(1)+(2)+(3)-(4)], kcal mol-1

Estimated free energy of binding,

kJ mol-1 vdW + Hbond

+ desolv. Energy

Electrostatic energy

Total (1)

27 -7.90 -0.50 -8.41 -1.14 1.79 -1.14 -6.62 -27.70 32 -8.53 -0.18 -8.71 -1.95 1.19 -1.95 -7.51 -31.42 33 -7.10 -0.44 -7.55 -0.07 0.60 -0.07 -6.95 -29.08 34 -8.07 -0.49 -8.56 -1.42 1.19 -1.42 -7.36 -30.79

182

6.4 Conclusions

This chapter presents the synthesis and characterisation of four thiosemicarbazone

Schiff bases (25-28) containing 4-methyl-3-thiosemicarbazide or 4-phenyl-3-

thiosemicarbazide with oVa or catechol. Their penta-coordinated diphenyl- (29, 32, 35,

38) and dimethyltin(IV) (30, 33, 36, 39) and hexa-coordinated tin(IV) (31, 34, 37, 40)

compounds of 25, 26, 27 and 28 are also discussed in this chapter. The Schiff bases in

this work behaved as tridentate binegatively charged ONS which coordinated to the tin

ion via thiolate-S, phenoxide-O and imine-N atoms. The in vitro cytotoxicity of the

compounds were determined using a panel of cancer cell lines viz., HT29, U87, SJ-G2,

MCF-7, A2780, H460, A431, Du145, BE2-C and MIA and one normal breast cell line,

MCF-10A. Schiff bases 25 and 28 showed selective strong cytotoxicity against certain

cancer cells tested. 25 showed good selectivity against HT29, A2780, A431, BE2-C and

MIA, whereas 28 exhibited strong selectivity against MCF-7 and A2780. 26 displayed

moderate cytotoxicity with no selectivity. Very strong selective cytotoxicity against

HT29, MCF-7, A2780, A431, BE2-C and SJ-G2 were observed for 29, suggesting that

the complexation with diphenyltin(IV) enhanced the cytotoxicity. With the exception of

MIA, the same pattern was found here as that observed for the compounds discussed in

Chapter 3 and 4. Good cytotoxic activity with no selectivity was observed in tin(IV)

compounds 32, 36, 37, 38, 39 and 40. In general, H460, A431 and Du145 were more

resistant towards 25, 28, 30, 33, 35, 38, 39 and 40. DNA binding studies demonstrated

the existence of groove binding interactions between compounds 32 and 34 with DNA,

which suggested that binding with DNA was one of the mechanisms of action that

enhanced the cytotoxicity for both compounds. Molecular docking simulations predicted

that the compounds bound to the G-C rich groove of DNA and they were stabilised by

hydrogen, electrostatic and hydrophobic interactions.

183

The summary of the synthesis, structural evaluation and cytotoxicity studies for all

Schiff bases and tin(IV) compounds discussed in this research will be outlined in Chapter

7, which also contains of future direction of this research, such as to enhance solubility,

mechanism of action and toxicity evaluation and also molecular dynamic simulations in

order to strengthen the value of the compounds discussed in this research.

184

CHAPTER 7

CONCLUSIONS AND FUTURE RECOMMENDATIONS

7.1 Conclusions

The aim of this project was to synthesise, structurally characterise and evaluate the

cytotoxicity of tin(IV) compounds derived from six dithiocarbazate and four

thiosemicarbazone Schiff bases. The compounds were characterised by elemental

analysis, molar conductivity, various spectroscopic techniques and X-ray

crystallography. In vitro cytotoxicity was investigated against a panel of cancer cell lines

(EJ-28, RT-112, HT29, U87, SJ-G2, MCF-7, A2780, H460, A431, Du145, BE2-C, MIA)

and one normal breast cell (MCF-10A) using a MTT metabolic method based on their

mitochondrial dehydrogenase activity. The interactions between DNA and compounds

was also investigated via UV-vis absorption spectroscopy. Molecular docking simulation

was carried out in order to theoretically understand the potential types of interactions that

occurred. The compounds were classed in five libraries and discussed in Chapter 2-6.

Chapter 2 detailed the synthesis of three dithiocarbazate Schiff bases (1, 2 and 3)

derived from S1, S2 and S3 with oVa. The crystal structure of 3 comprised two almost

planar residues, i.e. the phenyl ring and the remaining 14 non-H atoms (r.m.s. deviation

= 0.0410 Å) which were orientated perpendicularly, forming a dihedral angle of

82.72(5)°. An analysis of the geometric parameters was consistent with the molecule

existing as the thione tautomer, and the conformation about the C=N bond was E. In vitro

cytotoxicity of 1, 2 and 3 showed moderate activity against all cancer cells (EJ-28, RT-

112, HT29, U87, SJ-G2, MCF-7, A2780, H460, A431, Du145, BE2-C, MIA) tested with

no specificity or selectivity. To enhance the cytotoxicity, these compounds were

complexed with tin ion and the findings were reported in Chapter 3.

185

Chapter 3 detailed the synthesis of nine organotin(IV) compounds (4-9) derived

from Schiff bases synthesised in Chapter 2. Three crystal structures of 7, 8 and 9 which

supported the coordination geometry of the compounds were obtained. Schiff bases

behaved as tridentate ONS donor ligands coordinating to the tin centre via azomethine

nitrogen, thiolo sulphur and oxygen atoms. Compounds 7, 8 and 9 adopted a five-

coordinated distorted trigonal bipyramidal environment, where one molecule of Schiff

base and two methyl groups were coordinated to the central tin. Organotin(IV)

compounds 4-9 were evaluated for their in vitro cytotoxicity against twelve cancer cell

lines (EJ-28, RT-112, HT29, U87, SJ-G2, MCF-7, A2780, H460, A431, Du145, BE2-C,

MIA) and one normal breast cell (MCF-10A). Selective strong cytotoxicity was observed

for MCF-7, A2780, BE2-C, SJ-G2 and MIA cell lines treated with diphenyltin

compounds (4, 5 and 6). Selected tin(IV) compounds, 4, 5 and 6 were evaluated for the

mechanism of death, ROS and Annexin V against RT-112 cell line, where these

compounds were found to induce ROS production and were marked as apoptosis inducer.

The significance of these target compounds was clearly shown by the high binding

interactions with DNA, where generally chemotherapeutic agents that are approved for

clinical use are compounds having the ability to damage DNA, block DNA synthesis by

inhibiting nucleic acid precursor biosynthesis or disrupting hormonal stimulation that

controls the cell growth.288 Theoretical understanding by molecular docking simulations

supported that the compounds bound to the groove of DNA strand via hydrogen bonding,

hydrophobic and electrostatic interactions.

Chapter 4 detailed the synthesis, characterisation of three dithiocarbazate Schiff

bases (10, 11 and 12) derived from catechol and their diphenyl- (13, 14 and 15) and

dimethyltin(IV) (16, 17 and 18) compounds. One crystal structure of Schiff base 12 and

four crystal structure data of five-coordinated distorted trigonal bipyramidal

186

organotin(IV) compounds 14, 15, 17 and 18 were obtained. All compounds were

investigated for their cytotoxicity against ten cancer cell lines (HT29, U87, SJ-G2, MCF-

7, A2780, H460, A431, Du145, BE2-C, MIA) and one normal breast cell (MCF-10A). A

similar selectivity pattern was observed for 13, 14 and 15, however, DNA was not one of

the cytotoxic targets, where the compounds had only weak DNA binding interactions.

Chapter 5 detailed the synthesis of homoleptic tin(IV) compounds (19, 20, 21, 22,

23 and 24) coordinated with two molecules of dithiocarbazate Schiff bases synthesised in

Chapter 2 and 4. Two crystal structure data of distorted octahedral 20 and 21 were

obtained. All compounds exhibited similar cytotoxic activity compared to their respective

Schiff bases, possibly due to the bulky structure of the compounds and hence the difficulty

of interaction with the biomacromolecules of the cancer cells. The mechanisms of action

of these compounds remain unknown and warrant further biochemical investigations.

Lastly, in Chapter 6, the synthesis of four thiosemicarbazone Schiff bases derived

from oVa and catechol and their tin(IV) compounds is reported. Amongst the Schiff

bases, thiosemicarbazone derived Schiff bases, 25 and 28 had the best cytotoxic potency

against HT29, A2780, A431, BE2-C and MIA cells. These compounds are useful lead

candidates for future organic drug design development to treat cancers. However, the

mechanism of action of these compounds have yet to be verified. Good cytotoxic activity

with no selectivity pattern was observed in tin(IV) compounds 32, 36, 37, 38, 39 and 40.

H460, A431 and Du145 were observed to be more resistant towards 25, 28, 30, 33, 35,

38, 39 and 40. DNA binding studies demonstrated that 32 and 34 had good binding ability

compared to 27 and 33, which suggested that DNA was one of the cytotoxic targets for

32 and 34.

187

7.2 Future Recommendations

A number of recommendations for future research have resulted from the detailed

investigations in this thesis. The poor water solubility of synthesised Schiff bases and

tin(IV) compounds is a major issue in the course of applications, particularly anticancer

activity in this work. Pre-dissolution in high concentration of DMSO was used to

overcome their low solubility in aqueous media for anticancer studies. Due to the toxicity

of final DMSO concentration is higher than 3% that has been shown to retard cancerous

cells at high concentrations, there is a need of highly water soluble compounds for future

use, where they will be able to enter preclinical and clinical investigations without the use

of any solubilisers. To overcome this problem, peptides as targeting vectors for the direct

transportation of Schiff bases and tin(IV) compounds can be utilised.289 The attachment

of synthesised compounds to one or more polyethylene glycol (PEG) chains might also

enhance the compounds aqueous solubility, concomitant with the increase in bioactivity.

PEG is a non-toxic, non-immunogenic and non-antigenic polymer that is highly water

soluble. The introduction of FDA-approved PEG into pharmaceuticals could enhance

their pharmacokinetics which “prolong residence in body, decrease degradation by

metabolic enzymes and reduce or eliminate protein immunogenicity”.290

In addition, the mechanism of action of Schiff bases and tin(IV) compounds has yet

to be verified. Cellular uptake will be the next crucial analysis in order to investigate the

compound uptake in the cells. Therefore, efforts to determine the concentration and the

mechanism of action of the compounds are important in the line of succeeding in vivo

studies and clinical trials. Given the exciting activity of synthesised Schiff bases and

tin(IV) compounds, future work could investigate the toxicity effect of the compounds

towards normal cells. In this case, MCF-10A is positive for telomerase reverse

transcriptase289 which is known to be up-regulated in many cancer cells as well. The

188

decrease in cell viability after treatment with synthesised compounds may be due to the

inactivation of this enzyme. The most relevant option to determine toxicity in vitro is by

using bone marrow or intestinal epithelial cells.291 Furthermore, the most accurate relative

toxicity can be measured by in vivo studies. Zebrafish is one of the model organisms used

for in vivo experiment, where the zebrafish have morphological and physiological

similarities to mammals. Zebrafish are grown rapidly on a large scale which is sufficient

for toxicity screening. Other than that, zebrafish promises to contribute to several other

aspects of the drug development process, including target identification, target validation,

morpholino oligonucleotide screens and structure-activity relationship studies.292–294

Hitherto, medicinal chemists face contemporary challenges in molecular docking

in order to evaluate the performance of synthesised compounds towards

biomacromolecules, such as DNA in this work. Even though software solutions for

molecular docking based on end point calculations (i.e. evaluation of binding energy and

interactions using a single, rigid structure of synthesised compounds and

biomacromolecules) have achieved performance that guarantees their great involvement

in drug discovery pipelines, several issues are to be addressed. High-accuracy of

molecular modelling is crucial in order to create the system mimics with biological

systems, and one of the prediction techniques is molecular dynamic (MD) simulations.

This approach could be an excellent method for both accuracy and cost-effective as

compared to the molecular docking method.

189

REFERENCES

[1] Cooper, G. M. The Cell: A Molecular Approach; 2nd ed.; Sinauer Associates Inc.: Massachusetts, United States, 2000.

[2] Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R. L.; Torre, L. A.; Jemal, A. Global Cancer Statistics 2018: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA. Cancer J. Clin. 2018, 68, 394–424.

[3] Walker, B. J.; Gerber, A. Occupational Exposure to Aromatic Amines: Benzidine and Benzidine-Based Dyes. Natl. Cancer Inst. Monogr. 1982, 58, 11–13.

[4] Cancer treatment https://www.mayoclinic.org/tests-procedures/cancer-treatment/about/pac-20393344 (accessed May 2, 2017).

[5] Understanding Targeted Therapy https://www.cancer.net/navigating-cancer-care/how-cancer-treated/personalized-and-targeted-therapies/understanding-targeted-therapy (accessed Apr 3, 2018).

[6] Nicholson, L. B. The Immune System. Essays Biochem. 2016, 60, 275–301.

[7] Helissey, C.; Vicier, C.; Champiat, S. The Development of Immunotherapy in Older Adults: New Treatments, New Toxicities? J. Geriatr. Oncol. 2016, 7, 325–333.

[8] Baldo, B. A. Adverse Events to Monoclonal Antibodies Used for Cancer Therapy. Oncoimmunology 2013, 2, e26333.

[9] Scott, A. M.; Allison, J. P.; Wolchok, J. D. Monoclonal Antibodies in Cancer Therapy. Cancer Immun. 2012, 12, 1–8.

[10] Shepard, H. M.; Phillips, G. L.; Thanos, C. D.; Feldmann, M. Developments in Therapy with Monoclonal Antibodies and Related Proteins. Clin. Med. 2017, 17, 220–232.

[11] Akam, E. A.; Tomat, E. Targeting Iron in Colon Cancer via Glycoconjugation of Thiosemicarbazone Prochelators. Bioconjug. Chem. 2016, 27, 1807–1812.

[12] Palma, E.; Mendes, F.; Morais, G. R.; Rodrigues, I.; Santos, I. C.; Campello, M. P. C.; Raposinho, P.; Correia, I.; Gama, S.; Belo, D.; Alves, V.; Abrunhosa, A. J.; Santos, I.; Paulo, A. Biophysical Characterization and Antineoplastic Activity of New Bis(thiosemicarbazonato) Cu(II) Complexes. J. Inorg. Biochem. 2017, 167, 68–79.

[13] Refat, M. S.; El-Deen, I. M.; Anwer, Z. M.; El-Ghol, S. Bivalent Transition Metal Complexes of Coumarin-3-yl Thiosemicarbazone Derivatives: Spectroscopic, Antibacterial Activity and Thermogravimetric Studies. J. Mol. Struct. 2009, 920, 149–162.

[14] Prasad, K. S.; Kumar, L. S.; Prasad, M.; Revanasiddappa, H. D. Novel

190

Organotin(IV)-Schiff Base Complexes: Synthesis, Characterization, Antimicrobial Activity, and DNA Interaction Studies. Bioinorg. Chem. Appl. 2010, 854514.

[15] Lebwohl, D.; Canetta, R. Clinical Development of Platinum Complexes in Cancer Therapy: An Historical Perspective and an Update. Eur. J. Cancer 1998, 34, 1522–1534.

[16] Galanski, M. Recent Developments in the Field of Anticancer Platinum Complexes. Recent Pat. Anticancer. Drug Discov. 2006, 1, 285–295.

[17] Alama, A.; Tasso, B.; Novelli, F.; Sparatore, F. Organometallic Compounds in Oncology: Implications of Novel Organotins as Antitumor Agents. Drug Discov. Today 2009, 14, 500–508.

[18] Fuertes, M. A.; Alonso, C.; Pérez, J. M. Biochemical Modulation of Cisplatin Mechanisms of Action: Enhancement of Antitumor Activity and Circumvention of Drug Resistance. Chem. Rev. 2003, 103, 645–662.

[19] Jung, Y.; Lippard, S. J. Direct Cellular Responses to Platinum-Induced DNA Damage Direct Cellular Responses to Platinum-Induced DNA Damage. Chem. Rev. 2007, 107, 1387–1407.

[20] Dasari, S.; Bernard, T. P. Cisplatin in Cancer Therapy: Molecular Mechanisms of Action. Eur. J. Pharmacol. 2014, 740, 364–378.

[21] Kim, J.; Pramanick, S.; Lee, D.; Park, H.; Kim, W. J. Polymeric Biomaterials for the Delivery of Platinum-Based Anticancer Drugs. Biomater. Sci. 2015, 3, 1002–1017.

[22] Subbaraj, P.; Ramu, A.; Raman, N.; Dharmaraja, J. Synthesis, Characterization, DNA Interaction and Pharmacological Studies of Substituted Benzophenone Derived Schiff Base Metal(II) Complexes. J. Saudi Chem. Soc. 2015, 19, 207–216.

[23] Drug resistance reversed in head and neck cancer https://www.labonline.com.au/content/life-scientist/news/drug-resistance-reversed-in-head-and-neck-cancer-1212457248 (accessed Feb 27, 2019).

[24] Pattan, S. R.; Pawar, S. B.; Vetal, S. S.; Gharate, U. D.; Bhawar, S. B. The Scope of Metal Complexes in Drug Design–a Review. Indian drugs 2012, 49, 5–12.

[25] Ott, I.; Gust, R. Non Platinum Metal Complexes as Anti-Cancer Drugs. Arch. Pharm. 2007, 340, 117–126.

[26] Ott, I.; Gust, R. Preclinical and Clinical Studies on the Use of Platinum Complexes for Breast Cancer Treatment. Anticancer. Agents Med. Chem. 2007, 7, 95–110.

[27] Ott, I.; Gust, R. Review Article Non Platinum Metal Complexes as Anti-Cancer Drugs. Arch Pharm Chem. Life Sci. 2007, 340, 117–126.

[28] Korfel, A.; Scheulen, M. E.; Hans-Joachim, S.; Grundel, O.; Harstrick, A.; Knoche, M.; Fels, L. M.; Skorzec, M.; Bach, F.; Baumgart, J.; Safi, G.; Seeber, S.; Thiel, E.; Benjamin, W. E. Phase I Clinical and Pharmacokinetic Dichloride in

191

Adults with Advanced Study of Titanocene Solid Tumors Agnieszka. Clin. Cancer Res. 1998, 4, 2701–2708.

[29] Kroger, N.; Kleeberg, U. R.; Mross, K.; Edler, L.; Sass, G.; Hossfeld, D. K. Phase II Clinical Trial of Titanocene Dichloride in Patients with Metastatic Breast Cancer. Onkologie 2000, 23, 60–62.

[30] Sun, H.; Li, H.; Weir, R. A.; Sadler, P. J. The First Specific Ti(IV)-Protein Complex: Potential Relevance to Anticancer Activity of Titanocenes. Angew. Chemie - Int. Ed. 1998, 37, 1577–1579.

[31] Guo, M.; Harvey, I.; Campopiano, D. J.; Sadler, P. J. Short Oxo-titanium(IV) Bond in Bacterial Transferrin: A Protein Target for Metalloantibiotics. Angew. Chemie - Int. Ed. 2006, 45, 2758–2761.

[32] Hartinger, C. G.; Jakupec, M. a; Zorbas-Seifried, S.; Groessl, M.; Egger, A.; Berger, W.; Zorbas, H.; Dyson, P. J.; Keppler, B. K. KP1019, a New Redox-Active Anticancer Agent--Preclinical Development and Results of a Clinical Phase I Study in Tumor Patients. Chem. Bodiversity 2008, 5, 2140–2155.

[33] Vacca, A.; Bruno, M.; Boccarelli, A.; Coluccia, M.; Ribatti, D.; Bergamo, A.; Garbisa, S.; Sartor, L.; Sava, G. Inhibition of Endothelial Cell Functions and of Angiogenesis by the Metastasis Inhibitor NAMI-A. Br. J. Cancer 2002, 86, 993–998.

[34] Morbidelli, L.; Donnini, S.; Filippi, S.; Messori, L.; Piccioli, F.; Orioli, P.; Sava, G.; Ziche, M. Antiangiogenic Properties of Selected ruthenium(III) Complexes That Are Nitric Oxide Scavengers. Br. J. Cancer 2003, 88, 1484–1491.

[35] Rademaker-Lakhai, J. M.; Bongard, D. Van Den; Pluim, D. A Phase I and Pharmacological Study with Imidazolium- Trans-DMSO-Imidazole-Tetrachlororuthenate, a Novel Ruthenium Anticancer Agent. Clin. Cancer Res. 2004, 10, 3717–3727.

[36] Hartinger, C. G.; Zorbas-Seifried, S.; Jakupec, M. A.; Kynast, B.; Zorbas, H.; Keppler, B. K. From Bench to Bedside - Preclinical and Early Clinical Development of the Anticancer Agent Indazolium Trans-[tetrachlorobis(1H-indazole)ruthenate(III)] (KP1019 or FFC14A). J. Inorg. Biochem. 2006, 100, 891–904.

[37] Bernstein, L. R. Mechanisms of Therapeutic Activity for Gallium. Pharmacol. Rev. 1998, 50, 665–682.

[38] Senderowicz, A. M.; Reid, R.; Headlee, D.; Abornathy, T.; Horti, J.; Lush, R. M.; Reed, E.; Figg, W. D.; Sausville, E. A. A Phase II Trial of Gallium Nitrate in Patients with Androgen-Metastatic Prostate Cancer. Urol. Int. 1999, 63, 120–125.

[39] Galanski, M.; Arion, V. B.; Jakupec, M. A.; Keppler, B. K. Recent Developments in the Field of Tumor-Inhibiting Metal Complexes. Curr. Pharm. Des. 2003, 9, 2078–2089.

192

[40] Louie, A. Y.; Meade, T. J. Metal Complexes as Enzyme Inhibitors. Chem. Rev. 1999, 99, 2711–2734.

[41] Timerbaev, A. R. Advances in Developing tris(8-quinolinolato)gallium(III) as an Anticancer Drug: Critical Appraisal and Prospects. Metallomics 2009, 1, 193–198.

[42] Kopf-Maier, P.; Kopf, H.; Neuse, E. W. Ferricenium Complexes : A New Type of Water-Soluble Antitumor Agent. J. Cancer Res. Clin. Oncol. 1984, 108, 336–340.

[43] Hillard, E.; Vessières, A.; Thouin, L.; Jaouen, G.; Amatore, C. Ferrocene-Mediated Proton-Coupled Electron Transfer in a Series of Ferrocifen-Type Breast-Cancer Drug Candidates. Angew. Chemie - Int. Ed. 2006, 45, 285–290.

[44] Wong, E. L.-M.; Fang, G.-S.; Che, C.-M.; Zhu, N. Highly Cytotoxic Iron(II) Complexes with Pentadentate Pyridyl Ligands as a New Class of Anti-Tumor Agents. Chem. Commun. 2005, 36, 4578–4580.

[45] Jung, M.; Kerr, D. E.; Senter, P. D. Bioorganometallic Chemistry--Synthesis and Antitumor Activity of Cobalt Carbonyl Complexes. Arch. Pharm. (Weinheim). 1997, 330, 173–176.

[46] Ott, I.; Schmidt, K.; Kircher, B.; Schumacher, P.; Wiglenda, T.; Gust, R. Antitumor-Active Cobalt-Alkyne Complexes Derived from Acetylsalicylic Acid: Studies on the Mode of Drug Action. J. Med. Chem. 2005, 48, 622–629.

[47] Ott, I.; Abraham, A.; Schumacher, P.; Shorafa, H.; Gastl, G.; Gust, R.; Kircher, B. Synergistic and Additive Antiproliferative Effects on Human Leukemia Cell Lines Induced by Combining Acetylenehexacarbonyldicobalt Complexes with the Tyrosine Kinase Inhibitor Imatinib. J. Inorg. Biochem. 2006, 100, 1903–1906.

[48] Simon, T. M.; Kunishima, D. H.; Vibert, G. J.; Lorber, A. Inhibitory Effects of a New Oral Gold Compound on HELA Cells. Cancer 1979, 44, 1965–1975.

[49] Hoke, G. D.; Rush, G. F.; Mirabelli, C. K. The Mechanism of Acute Cytotoxicity of Triethylphosphine Gold(l) Complexes. Toxicol. Appl. Pharmacol. 1989, 99, 50–60.

[50] McKeage, M. J.; Maharaj, L.; Berners-Price, S. J. Mechanisms of Cytotoxicity and Antitumor Activity of Gold(I) Phosphine Complexes: The Possible Role of Mitochondria. Coord. Chem. Rev. 2002, 232, 127–135.

[51] Powis, G.; Kirkpatrick, D. L. Thioredoxin Signaling as a Target for Cancer Therapy. Curr. Opin. Pharmacol. 2007, 7, 392–397.

[52] Gromer, S.; Urig, S.; Becker, K. The Thioredoxin System - From Science to Clinic. Med. Res. Rev. 2004, 24, 40–89.

[53] Liu, K.; Yan, H.; Chang, G.; Li, Z.; Niu, M.; Hong, M. Organotin(IV) Complexes Derived from Hydrazone Schiff Base: Synthesis, Crystal Structure, In Vitro Cytotoxicity and DNA/BSA Interactions. Inorganica Chim. Acta 2017, 464, 137–146.

193

[54] Saeed, A.; Channar, P. A.; Larik, F. A.; Jabeen, F.; Muqadar, U.; Saeed, S.; Flörke, U.; Ismail, H.; Dilshad, E.; Mirza, B. Design, Synthesis, Molecular Docking Studies of Organotin-Drug Derivatives as Multi-Target Agents against Antibacterial, Antifungal, α-Amylase, α-Glucosidase and Butyrylcholinesterase. Inorganica Chim. Acta 2017, 464, 204–213.

[55] Wang, X.; Qiang, J.; Ni, X.; Jia, A.; Ma, X.; Tong, B.; Zhang, Q. Synthesis, Structure and DFT Calculation of a Facial Tris-Cyclometalated Phosphorescent Iridium(III) Complex Containing Substituted Phthalazine Ligands. Inorganica Chim. Acta 2017, 466, 343–348.

[56] Davies, A. G.; Gielen, M.; Pannell, K. H.; Tiekink, E. R. T. Tin Chemistry: Fundamentals, Fronteir, and Applications; Wiley: United Kingdom, 2008.

[57] Arjmand, F.; Parveen, S.; Tabassum, S.; Pettinari, C. Organo-Tin Antitumor Compounds : Their Present Status in Drug Development and Future Perspectives. Inorganica Chim. Acta 2014, 423, 26–37.

[58] Pellerito, L.; Nagy, L. Organotin(IV)n+ Complexes Formed with Biologically Active Ligands: Equilibrium and Structural Studies, and Some Biological Aspects. Coord. Chem. Rev. 2002, 224, 111–150.

[59] Pellerito, C.; Nagy, L.; Pellerito, L.; Szorcsik, A. Biological Activity Studies on organotin(IV)n+ Complexes and Parent Compounds. J. Organomet. Chem. 2006, 691, 1733–1747.

[60] Casas, J. S.; Castellano, E. E.; Couce, M. D.; Ellena, J.; Sanchez, A.; Sanchez, J. L.; Sordo, J.; Taboada, C. The Reaction of Dimethyltin(IV) Dichloride with Thiamine Diphosphate (H2TDP): Synthesis and Structure of [SnMe2(HTDP)(H2O]Cl·H2O, and Possibility of a Hitherto Unsuspected Role of the Metal Cofactor in the Mechanism of Vitamin-B 1 -Dependent Enzymes. Inorg. Chem. 2004, 43, 1957–1963.

[61] Gennari, A.; Viviani, B.; Galli, C. L.; Marinovich, M.; Pieters, R.; Corsini, E. Organotins Induce Apoptosis by Disturbance of [Ca2+]i and Mitochondrial Activity, Causing Oxidative Stress and Activation of Caspases in Rat Thymocytes. Toxicol. Appl. Pharmacol. 2000, 169, 185–190.

[62] Jaiswal, N. Organotin (IV) Complexes and DNA Interaction : A Promising Future for Tin Based Metallodrugs. Int. J. ChemTech Res. 2017, 10, 495–499.

[63] Collier, W. A. Zur Experimentellen Therapie Der Tumoren. H , Die Wirksamkeit Einiger Metallorganischen Blei- Uud Zinnverbindungen. Zeit. Hyg. Infektkr 1929, 10, 169–174.

[64] Crowe, A. J.; Smith, P. J. Investigations into the Antitumour Activity of Organotin Compounds. 2.* Diorganotin Dihalide and Di-Pseudohalide Complexes. Inorganica Chim. Acta 1984, 93, 179–184178.

[65] Haque, R. A.; Salam, M. A.; Arafath, M. A. New organotin(IV) Complexes with N (4)-Methylthiosemicarbazone Derivatives Prepared from 2,3-

194

Dihydroxybenzaldehyde and 2-Hydroxy-5-Methylbenzaldehyde: Synthesis, Characterization, and Cytotoxic Activity. J. Coord. Chem. 2015, 68, 2953–2967.

[66] Yang, Y.; Hong, M.; Xu, L.; Cui, J.; Chang, G.; Li, D.; Li, C. Organotin(IV) Complexes Derived from Schiff Base N’-[(1E)-(2-Hydroxy-3- methoxyphenyl)methylidene]pyridine-3-carbohydrazone: Synthesis, In Vitro Cytotoxicities and DNA/BSA Interaction. J. Organomet. Chem. 2016, 804, 48–58.

[67] Arakawa, Y. Invasion of Biofunctions by Organotins-Immune System, Brain Nervous System and Endocrine System. Biomed. Res. Trace Elem. 2000, 11, 269–286.

[68] Shang, X.; Meng, X.; Alegria, E. C. B. A.; Li, Q.; Guedes Da Silva, M. F. C.; Kuznetsov, M. L.; Pombeiro, A. J. L. Syntheses, Molecular Structures, Electrochemical Behavior, Theoretical Study, and Antitumor Activities of Organotin(IV) Complexes Containing 1-(4-Chlorophenyl)-1-Cyclopentanecarboxylato Ligands. Inorg. Chem. 2011, 50, 8158–8167.

[69] Hadi, S.; Rilyanti, M. Synthesis and in Vitro Anticancer Activity of Some Organotin(IV) Benzoate Compounds. Orient. J. Chem. 2010, 26, 775–779.

[70] Rehman, W.; Badshah, A.; Rahim, F.; Baloch, M. K.; Ullah, H.; Abid, O. U. R.; Nawaz, M.; Tauseef, I. Synthesis, Spectral Characterization, Antibacterial and Antitumor Studies of Some diorganotin(IV) Complexes Derived from 2-Phenylmonomethylglutarate. Inorganica Chim. Acta 2014, 423, 177–182.

[71] Jiménez-Pérez, V. M.; García-López, M. C.; Muñoz-Flores, B. M.; Chan-Navarro, R.; Berrones-Reyes, J. C.; Dias, H. V. R.; Moggio, I.; Arias, E.; Serrano-Mireles, J. A.; Chavez-Reyes, A. New Application of Fluorescent Organotin Compounds Derived from Schiff Bases: Synthesis, X-Ray Structures, Photophysical Properties, Cytotoxicity and Fluorescent Bioimaging. J. Mater. Chem. B 2015, 3, 5731–5745.

[72] Agiorgiti, M. S.; Evangelou, A.; Vezyraki, P.; Hadjikakou, S. K.; Kalfakakou, V.; Tsanaktsidis, I.; Batistatou, A.; Zelovitis, J.; Simos, Y. V.; Ragos, V.; Karkabounas, S.; Peschos, D. Cytotoxic Effect, Antitumour Activity and Toxicity of Organotin Derivatives with Ortho- or Para-Hydroxy-Benzoic Acids. Med. Chem. Res. 2018, 27, 1122–1130.

[73] Shang, X.; Zhao, B.; Xiang, G.; Guedes Da Silva, M. F. C.; Pombeiro, A. J. L. Dimeric Diorganotin(IV) Complexes with Arylhydrazones of β-Diketones: Synthesis, Structures, Cytotoxicity and Apoptosis Properties. RSC Adv. 2015, 5, 45053–45060.

[74] Gholivand, K.; Salami, R.; Shahsavari, Z.; Torabi, E. Novel Binuclear and Polymeric Diorganotin(IV) Complexes with N-Nicotinyl Phosphoramides: Synthesis, Characterization, Structural Studies and Anticancer Activity. J. Organomet. Chem. 2016, 819, 155–165.

[75] Zhong, G.-Y. Synthesis of [C5H5NH][Ph2Sn(μ2-SCH2COO)CI] with Good Biological Activity In Vitro Anti-Tumor. Adv. Mater. Res. 2013, 726–731, 222–224.

195

[76] Pellerito, L.; Prinzivalli, C.; Casella, G.; Fiore, T.; Pellerito, O.; Giuliano, M.; Scopelliti, M.; Pellerito, C. Diorganotin(IV) N-Acetyl-L-Cysteinate Complexes: Synthesis, Solid State, Solution Phase, DFT and Biological Investigations. J. Inorg. Biochem. 2010, 104, 750–758.

[77] Xanthopoulou, M. N.; Hadjikakou, S. K.; Hadjiliadis, N.; Kubicki, M.; Skoulika, S.; Bakas, T.; Baril, M.; Butler, I. S. Synthesis, Structural Characterization, and Biological Studies of Six- and Five-Coordinate Organotin(IV) Complexes with the Thioamides. Inorg. Chem. 2007, 46, 1187–1195.

[78] Xanthopoulou, M. N.; Hadjikakou, S. K.; Hadjiliadis, N.; Milaeva, E. R.; Gracheva, J. A.; Tyurin, V. Y.; Kourkoumelis, N.; Christoforidis, K. C.; Metsios, A. K.; Karkabounas, S.; Charalabopoulos, K. Biological Studies of New Organotin(IV) Complexes of Thioamide Ligands. Eur. J. Med. Chem. 2008, 43, 327–335.

[79] Milaeva, E. R.; Shpakovsky, D. B.; Gracheva, Y. A.; Antonenko, T. A.; Osolodkin, D. I.; Palyulin, V. A.; Shevtsov, P. N.; Neganova, M. E.; Vinogradova, D. V.; Shevtsova, E. F. Some Insight into the Mode of Cytotoxic Action of Organotin Compounds with Protective 2,6-Di- Tert -Butylphenol Fragments. J. Organomet. Chem. 2014, 782, 96–102.

[80] Adeyemi, J. O.; Onwudiwe, D. C.; Singh, M. Synthesis, Characterization, and Cytotoxicity Study of Organotin(IV) Complexes Involving Different Dithiocarbamate Groups. J. Mol. Struct. 2019, 1179, 366–375.

[81] Pillai, V.; Patel, S. K.; Buch, L.; Singh, V. K. Binuclear Diphenyltin(IV)dithiocarbamate Complexes Bearing Functionalized Linkers: Synthesis, Spectral Characterization, DFT and In Vitro Anticancer Activity. Appl. Organomet. Chem. 2018, 32, 1–12.

[82] Tian, L.; Zheng, X.; Zheng, X.; Sun, Y.; Yan, D.; Tu, L. Synthesis, Structure and Cytotoxic Activity of Binuclear Phenyltin(IV) Complexes with N,N-bis(2-Hydroxybenzyl)-1,2-ethanebis(dithiocarbamate) Ligand. Appl. Organomet. Chem. 2011, 25, 785–790.

[83] Yousefi, M.; Safari, M.; Torbati, M. B.; Kazemiha, V. M.; Sanati, H.; Amanzadeh, A. New Mononuclear Diorganotin(IV) Dithiocarboxylates: Synthesis, Characterization and Study of Their Cytotoxic Activities. Appl. Organomet. Chem. 2012, 26, 438–444.

[84] Ali, M. A.; Livingstone, S. E. Metal Complexes of Sulphur-Nitrogen Chelating Agents. Coord. Chem. Rev. 1974, 13, 101–132.

[85] Hamid, M. H. S. A.; Said, A. N. A. H.; Mirza, A. H.; Karim, M. R.; Arifuzzaman, M.; Akbar Ali, M.; Bernhardt, P. V. Synthesis, Structures and Spectroscopic Properties of Some Tin(IV) Complexes of the 2-Acetylpyrazine Schiff Bases of S-Methyl- and S-Benzyldithiocarbazates. Inorganica Chim. Acta 2016, 453, 742–750.

[86] Kumar, A.; Chaudhary, P.; Singh, R.; Kaushik, N. K. Organotin(IV) Complexes

196

of Thiohydrazones of Phenethylamine: Synthesis, Characterization, Biological and Thermal Study. Main Gr. Chem. 2016, 15, 163–178.

[87] Belwal, S.; Singh, R. V. Inventive Triorgano Tin(IV) Complexes of Biologically Potent Schiff Base Derivatives. Eur. J. Adv. Eng. Technol. 2015, 2, 34–42.

[88] Beshir, A. B.; Guchhait, S. K.; Gascón, J. A.; Fenteany, G. Synthesis and Structure-Activity Relationships of Metal-Ligand Complexes That Potently Inhibit Cell Migration. Bioorganic Med. Chem. Lett. 2008, 18, 498–504.

[89] Vijayan, P.; Viswanathamurthi, P.; Sugumar, P.; Ponnuswamy, M. N.; Balakumaran, M. D.; Kalaichelvan, P. T.; Velmurugan, K.; Nandhakumar, R.; Butcher, R. J. Unprecedented Formation of Organo-ruthenium(II) Complexes Containing 2-Hydroxy-1-Naphthaldehyde S-Benzyldithiocarbazate: Synthesis, X-ray Crystal Structure, DFT Study and Their Biological Activities In Vitro. Inorg. Chem. Front. 2015, 2, 620–639.

[90] Tian, L.; Yu, H.; Zheng, X.; Liu, X. Synthesis, Crystal Structure and Cytotoxic Activity of Tricyclohexyltin Complexes of Benzenedioxyacetic Acids. Appl. Organomet. Chem. 2015, 29, 725–729.

[91] Akbar Ali, M.; Huq Mirza, A.; Kok Wei, L.; Bernhardt, P. V.; Atchade, O.; Song, X.; Eng, G.; May, L. Synthesis and Characterization of Pentagonal Bipyramidal Organotin(IV) Complexes of 2,6-Diacetylpyridine Schiff Bases of S-Alkyl- and Aryldithiocarbazates. J. Coord. Chem. 2010, 63, 1194–1206.

[92] Ali, M. A.; Majumder, S. M. M.-H.; Butcher, R. J.; Jasinski, J. P.; Jasinski, J. M. The Preparation and Characterization of Bis-chelated Nickel(ll) Complexes of the 6-Methylpyridine-2-carboxaldehyde Schiff Bases of S-Alkyldithiocarbazates and the X-ray Crystal Structure of the BisS-methyl-β-N-(6-methylpyrid-2-yl)- methylenedithiocarbazatonickel(ll) Complex.. Polyhedron 1997, 16, 2749–2754.

[93] Akbar Ali, M.; Mirza, A. H.; Hamid, M. H. S. A.; Bernhardt, P. V. Diphenyltin(IV) Complexes of the 2-Quinolinecarboxaldehyde Schiff Bases of S-Methyl- and S-Benzyldithiocarbazate (Hqaldsme and Hqaldsbz): X-ray Crystal Structures of Hqaldsme and Two Conformers of Its Diphenyltin(IV) Complex. Polyhedron 2005, 24, 383–390.

[94] Ali, M. A.; Mirza, A. H.; Butcher, R. J.; Crouse, K. A. The Preparation, Characterization and Biological Activity of Palladium(II) and Platinum(II) Complexes of Tridentate NNS Ligands Derived from S-Methyl- and S-Benzyldithiocarbazates and the X-ray Crystal Structure of the [Pd(mpasme)Cl] Complex. Transit. Met. Chem. 2006, 31, 79–87.

[95] Elsayed, S. A.; Noufal, A. M.; El-Hendawy, A. M. Synthesis, Structural Characterization and Antioxidant Activity of Some Vanadium(IV), Mo(VI)/(IV) and Ru(II) Complexes of Pyridoxal Schiff Base Derivatives. J. Mol. Struct. 2017, 1144, 120–128.

[96] Low, M. L.; Maigre, L.; Tahir, M. I. M.; Tiekink, E. R. T.; Dorlet, P.; Guillot, R.; Ravoof, T. B.; Rosli, R.; Pages, J. M.; Policar, C.; Delsuc, N.; Crouse, K. A. New

197

Insight into the Structural, Electrochemical and Biological Aspects of Macroacyclic Cu(II) Complexes Derived from S-Substituted Dithiocarbazate Schiff Bases. Eur. J. Med. Chem. 2016, 120, 1–12.

[97] Omar, S. A.; Ravoof, T. B. S. A.; Tahir, M. I. M.; Crouse, K. A. Synthesis and Characterization of Mixed-Ligand copper(II) Saccharinate Complexes Containing Tridentate NNS Schiff Bases. X-Ray Crystallographic Analysis of the Free Ligands and One Complex. Transit. Met. Chem. 2013, 39, 119–126.

[98] Pavan, F. R.; Maia, P. I. da S.; Leite, S. R. A.; Deflon, V. M.; Batista, A. A.; Sato, D. N.; Franzblau, S. G.; Leite, C. Q. F. Thiosemicarbazones, Semicarbazones, Dithiocarbazates and Hydrazide/hydrazones: Anti - Mycobacterium Tuberculosis Activity and Cytotoxicity. Eur. J. Med. Chem. 2010, 45, 1898–1905.

[99] Ravoof, T. B. S. A.; Crouse, K. A.; Tahir, M. I. M.; How, F. N. F.; Rosli, R.; Watkins, D. J. Synthesis, Characterization and Biological Activities of 3-Methylbenzyl 2-(6-Methylpyridin-2-ylmethylene)hydrazine Carbodithioate and Its Transition Metal Complexes. Transit. Met. Chem. 2010, 35, 871–876.

[100] Singh, H. L.; Varshney, A. K. Synthetic, Structural, and Biochemical Studies of Organotin(IV) with Schiff Bases Having Nitrogen and Sulphur Donor Ligands. Bioinorg. Chem. Appl. 2006, 1–7.

[101] Yusof, E. N. M.; Ravoof, T. B. S. A.; Jamsari, J.; Tiekink, E. R. T.; Veerakumarasivam, A.; Crouse, K. A.; Tahir, M. I. M.; Ahmad, H. Synthesis, Characterization and Biological Studies of S-4-Methylbenzyl-β-N-(2-Furylmethylene)dithiocarbazate (S4MFuH) Its Zn2+, Cu2+, Cd2+ and Ni2+ Complexes. Inorganica Chim. Acta 2015, 438, 85–93.

[102] Shahzadi, S.; Ali, S.; Fettouhi, M. Synthesis, Spectroscopy, In Vitro Biological Activity and X-ray Structure of (4-Methylpiperidine-Dithiocarbamato-S,S′)triphenyltin(IV). J. Chem. Crystallogr. 2008, 38, 273–278.

[103] Ravoof, T. B. S. A.; Crouse, K. A.; Tahir, M. I. M.; Rosli, R.; Watkin, D. J.; How, F. N. F. Synthesis, Characterisation and Biological Activities of 2-Methylbenzyl 2-(Dipyridin-2-ylmethylene)hydrazinecarbodithioate. J. Chem. Crystallogr. 2011, 41, 491–495.

[104] Tarafder, M. T. H.; Ali, M. A.; Wee, D. J.; Azahari, K.; Silong, S.; Crouse, K. A. Complexes of a Tridentate ONS Schiff Base. Synthesis and Biological Properties. Transit. Met. Chem. 2000, 25, 456–460.

[105] Chew, K.-B.; Tarafder, M. T. .; Crouse, K. A.; Ali, A. .; Yamin, B. .; Fun, H.-K. Synthesis, Characterization and Bio-Activity of Metal Complexes of Bidentate N–S Isomeric Schiff Bases Derived from S-Methyldithiocarbazate (SMDTC) and the X-Ray Structure of the bis[S-Methyl-β-N-(2-Furyl-methylketone)dithiocarbazato]cadmium(II) Complex. Polyhedron 2004, 23, 1385–1392.

[106] Tarafder, M. T. H.; Chew, K.; Crouse, K. A.; Ali, A. M.; Yamin, B. M.; Fun, H. K. Synthesis and Characterization of Cu(II), Ni(II) and Zn(II) Metal Complexes of

198

Bidentate NS Isomeric Schiff Bases Derived from S-Methyldithiocarbazate (SMDTC): Bioactivity of the Bidentate NS Isomeric Schiff Bases , Some of Their Cu(II), Ni(II) and Zn(II). Polyhedron 2002, 21, 2683–2690.

[107] How, F. N.-F.; Crouse, K. A.; Tahir, M. I. M.; Tarafder, M. T. H.; Cowley, A. R. Synthesis, Characterization and Biological Studies of S-Benzyl-β-N-(Benzoyl) Dithiocarbazate and Its Metal Complexes. Polyhedron 2008, 27, 3325–3329.

[108] Crouse, K. A.; Chew, K. B.; Tarafder, M. T. H.; Kasbollah, A.; Ali, A. M.; Yamin, B. M.; Fun, H. K. Synthesis, Characterization and Bio-Activity of S-2-Picolyldithiocarbazate (S2PDTC), Some of Its Schiff Bases and Their Ni(II) Complexes and X-Ray Structure of S-2-Picolyl-β-N-(2-Acetylpyrrole)dithiocarbazate. Polyhedron 2004, 23, 161–168.

[109] Begum, M. S.; Zangrando, E.; Sheikh, M. C.; Miyatake, R.; Howlader, M. B. H.; Rahman, M. N.; Ghosh, A. Bischelated Complexes of a Dithiocarbazate N,S Schiff Base Ligand: Synthesis, Characterization and Antimicrobial Activities. Transit. Met. Chem. 2017, 42, 553–563.

[110] Yusof, E. N. M.; Ravoof, T. B. S. A.; Tiekink, E. R. T.; Veerakumarasivam, A.; Crouse, K. A.; Tahir, M. I. M.; Ahmad, H. Synthesis, Characterization and Biological Evaluation of Transition Metal Complexes derived from N, S Bidentate Ligands. Int. J. Mol. Sci. 2015, 16, 11034–11054.

[111] Islam, M. A. A. A. A.; Sheikh, M. C.; Alam, M. S.; Zangrando, E.; Alam, M. A.; Tarafder, M. T. H.; Miyatake, R. Synthesis, Characterization and Bio-Activity of a Bidentate NS Schiff Base of S-Allyldithiocarbazate and Its Divalent Metal Complexes: X-Ray Crystal Structures of the Free Ligand and Its nickel(II) Complex. Transit. Met. Chem. 2014, 39, 141–149.

[112] Khoo, T.-J.; Break, M. K. B.; Crouse, K. A.; Tahir, M. I. M.; Ali, A. M.; Cowley, A. R.; Watkin, D. J.; Tarafder, M. T. H. Synthesis, Characterization and Biological Activity of Two Schiff Base Ligands and Their Nickel(II), Copper(II), Zinc(II) and Cadmium(II) Complexes derived from S-4-Picolyldithiocarbazate and X-Ray Crystal Structure of Cadmium(II) Complex derived from Pyridine-2-carboxaldehyde. Inorganica Chim. Acta 2014, 413, 68–76.

[113] Zangrando, E.; Begum, M. S.; Sheikh, M. C.; Miyatake, R.; Hossain, M. M.; Alam, M. M.; Hasnat, M. A.; Halim, M. A.; Ahmed, S.; Rahman, M. N.; Gosh, A. Synthesis, Characterization, Density Functional Study and Antimicrobial Evaluation of a Series of Bischelated Complexes with a Dithiocarbazate Schiff Base Ligand. Arab. J. Chem. 2017, 10, 172–184..

[114] Zangrando, E.; Begum, M. S.; Miyatake, R.; Sheikh, M. C.; Hossain, M. M. Crystal Structure of bisS-Hexyl 3-[4-Dimethylamino)benzylidene]dithiocarbazato-κ 2 N 3,S -copper(II). Acta Crystallogr. Sect. E Crystallogr. Commun. 2015, E71, 706–708.

[115] Li, H.-Q.; Luo, Y.; Li, D.-D.; Zhu, H.-L. (E)-4-Chlorobenzyl 3-(3-Nitrobenzylidene)dithiocarbazate. Acta Crystallogr. Sect. E Struct. Reports Online 2009, E65, o3101.

199

[116] Ali, M. A.; Mirza, A. H.; Tan, A. L.; Wei, L. K.; Bernhardt, P. V. The Preparation and Characterization of tin(IV) Complexes of 2-Quinolinecarboxaldehyde Schiff Bases of S-Methyl- and S-Benzyldithiocarbazates and the X-Ray Crystal and Molecular Structures of the 2-Quinolinecarboxaldehyde Schiff Base of S-Benzyldithiocarb. Polyhedron 2004, 23, 2405–2412.

[117] Ali, M. A.; Huq, A.; Haniti, M.; Hamid, S. A.; Bernhardt, P. V; Atchade, O.; Song, X.; Eng, G.; May, L. Synthesis , Spectroscopic and Structural Characterization of Diphenyltin(IV) Complexes of Acetone Schiff Bases of S -Alkyldithiocarbazates. Polyhedron 2008, 27, 977–984.

[118] Yekke-Ghasemi, Z.; Takjoo, R.; Ramezani, M.; Mague, J. T. Molecular Design and Synthesis of New Dithiocarbazate Complexes; Crystal Structure, Bioactivities and Nano Studies. RSC Adv. 2018, 8, 41795–41809.

[119] Scovill, J. P.; Klayman, D. L.; Lambros, C.; Childs, G. E.; Notsch, J. D. 2-Acetylpyridine Thiosemicarbazones. 9. Derivatives of 2-Acetylpyridine 1-Oxide as Potential Antimalarial Agents. J. Med. Chem. 1984, 27, 87–91.

[120] Hu, W.; Zhou, W.; Xia, C.; Wen, X. Synthesis and Anticancer Activity of Thiosemicarbazones. Bioorg. Med. Chem. Lett. 2006, 16, 2213–2218.

[121] Kovala-demerzi, D.; Domopoulou, A.; Demertzis, M. A.; Valle, G.; Papageorgiou, A. Palladium(II) Complexes of 2- Acetylpyridine N(4)-Methyl, N(4)-Ethyl and N(4)-Phenyl-Thiosemicarbazones. Crystal Structure of Chloro(2-Acetylpyridine N(4)-Methylthiosemicarbazonato) Palladium(II). Synthesis, Spectral Studies, In Vitro and In Vivo Antitumour. J. Inorg. Biocemistry 1997, 68, 147–155.

[122] Khan, S. A.; Yusuf, M. Synthesis, Spectral Studies and In Vitro Antibacterial Activity of Steroidal Thiosemicarbazone and Their Palladium (Pd(II)) Complexes. Eur. J. Med. Chem. 2009, 44, 2270–2274.

[123] Đilović, I.; Rubčić, M.; Vrdoljak, V.; Pavelić, S. K.; Kralj, M.; Piantanida, I.; Cindrić, M. Novel Thiosemicarbazone Derivatives as Potential Antitumor Agents: Synthesis, Physicochemical and Structural Properties, DNA Interactions and Antiproliferative Activity. Bioorg. Med. Chem. 2008, 16, 5189–5198.

[124] Palanimuthu, D.; Poon, R.; Sahni, S.; Anjum, R.; Hibbs, D.; Lin, H. Y.; Bernhardt, P. V.; Kalinowski, D. S.; Richardson, D. R. A Novel Class of Thiosemicarbazones Show Multi-Functional Activity for the Treatment of Alzheimer’s Disease. Eur. J. Med. Chem. 2017, 139, 612–632.

[125] Parrilha, G. L.; Da Silva, J. G.; Gouveia, L. F.; Gasparoto, A. K.; Dias, R. P.; Rocha, W. R.; Santos, D. A.; Speziali, N. L.; Beraldo, H. Pyridine-Derived Thiosemicarbazones and Their Tin(IV) Complexes with Antifungal Activity against Candida Spp. Eur. J. Med. Chem. 2011, 46, 1473–1482.

[126] Satheesh, D.; Jayanthi, K. An In Vitro Antibacterial and Antifungal Activities of Copper(II) and Zinc(II) Complexes of N4-Methyl-3-Thiosemicarbazones. Int. J. Chem. Pharm. Anal. 2017, 1–7.

200

[127] Singh, H. L.; Singh, J. B.; Bhanuka, S. Synthesis, Spectroscopic Characterization, Biological Screening, and Theoretical Studies of Organotin(IV) Complexes of Semicarbazone and Thiosemicarbazones Derived from (2-Hydroxyphenyl)(pyrrolidin-1-yl)methanone. Res. Chem. Intermed. 2016, 42, 997–1015.

[128] Wiecek, J.; Dokorou, V.; Ciunik, Z.; Kovala-Demertzi, D. Organotin Complexes of Pyruvic Acid Thiosemicarbazone: Synthesis, Crystal Structures and Antiproliferative Activity of Neutral and Cationic Diorganotin Complexes. Polyhedron 2009, 28, 3298–3304.

[129] Oliveira, A. A.; Franco, L. L.; dos Santos, R. G.; Perdigão, G. M. C.; da Silva, J. G.; Souza-Fagundes, E. M.; Beraldo, H. Neutron Activation of In(III) Complexes with Thiosemicarbazones Leads to the Production of Potential Radiopharmaceuticals for the Treatment of Breast Cancer. New J. Chem. 2017, 41, 9041–9050.

[130] Giles, F. J.; Fracasso, P. M.; Kantarjian, H. M.; Cortes, J. E.; Brown, R. A.; Verstovsek, S.; Alvarado, Y.; Thomas, D. A.; Faderl, S.; Garcia-Manero, G.; Wright, L. P.; Samson, T.; Cahill, A.; Lambert, P.; Plunkett, W.; Sznol, M.; DiPersio, J. F.; Gandhi, V. Phase I and Pharmacodynamic Study of Triapine®, a Novel Ribonucleotide Reductase Inhibitor, in Patients with Advanced Leukemia. Leuk. Res. 2003, 27, 1077–1083.

[131] Karp, J. E.; Giles, F. J.; Gojo, I.; Morris, L.; Greer, J.; Johnson, B.; Thein, M.; Sznol, M.; Low, J. A Phase I Study of the Novel Ribonucleotide Reductase Inhibitor 3-Aminopyridine-2-Carboxaldehyde Thiosemicarbazone (3-AP, Triapine®) in Combination with the Nucleoside Analog Fludarabine for Patients with Refractory Acute Leukemias and Aggressive Myelopro. Leuk. Res. 2008, 32, 71–77.

[132] Kunos, C. A.; Radivoyevitch, T.; Waggoner, S.; Debernardo, R.; Zanotti, K.; Resnick, K.; Fusco, N.; Adams, R.; Redline, R.; Faulhaber, P.; Dowlati, A. Radiochemotherapy plus 3-Aminopyridine-2-Carboxaldehyde Thiosemicarbazone (3-AP, NSC #663249) in Advanced-Stage Cervical and Vaginal Cancers. Gynecol. Oncol. 2013, 130, 75–80.

[133] Jansson, P. J.; Sharpe, P. C.; Bernhardt, P. V.; Richardson, D. R. Novel Thiosemicarbazones of the ApT and DpT Series and Their Copper Complexes: Identification of Pronounced Redox Activity and Characterization of Their Antitumor Activity. J. Med. Chem. 2010, 53, 5759–5769.

[134] Kolberg, M.; Strand, K. R.; Graff, P.; Andersson, K. K. Structure, Function, and Mechanism of Ribonucleotide Reductases. Biochim. Biophys. Acta 2004, 1699, 1–34.

[135] Aye, Y.; Long, M. J. C.; Stubbe, J. Mechanistic Studies of Semicarbazone Triapine Targeting Human Ribonucleotide Reductase In Vitro and in Mammalian Cells: Tyrosyl Radical Quenching Not Involving Reactive Oxygen Species. J. Biol. Chem. 2012, 287, 35768–35778.

201

[136] Lane, D. J. R.; Mills, T. M.; Shafie, N. H.; Merlot, A. M.; Saleh Moussa, R.; Kalinowski, D. S.; Kovacevic, Z.; Richardson, D. R. Expanding Horizons in Iron Chelation and the Treatment of Cancer: Role of Iron in the Regulation of ER Stress and the Epithelial-Mesenchymal Transition. Biochim. Biophys. Acta - Rev. Cancer 2014, 1845, 166–181.

[137] Yuan, J.; Lovejoy, D. B.; Richardson, D. R. Novel Di-2-Pyridyl-Derived Iron Chelators with Marked and Selective Antitumor Activity: In Vitro and In Vivo Assessment. Blood 2004, 104, 1450–1458.

[138] Khandani, M.; Sedaghat, T.; Erfani, N.; Haghshenas, M. R.; Khavasi, H. R. Synthesis, Spectroscopic Characterization, Structural Studies and Antibacterial and Antitumor Activities of Diorganotin Complexes with 3-Methoxysalicylaldehyde Thiosemicarbazone. J. Mol. Struct. 2013, 1037, 136–143.

[139] Liu, T.; Sun, J.; Tai, Y.; Qian, H.; Li, M. Synthesis, Spectroscopic Characterization, Crystal Structure, and Biological Evaluation of a Diorganotin(IV) Complex with 2-Acetylpyridine N 4-Cyclohexylthiosemicarbazone. Inorg. Nano-Metal Chem. 2017, 47, 813–817.

[140] Salam, M. A.; Hussein, M. A.; Ramli, I.; Islam, S. Synthesis, Structural Characterization and Evaluation of Biological Activity of Organotin(IV) Complexes with 2-Hydroxy-5-methoxybenzaldehyde-N(4)-methylthiosemicarbazone. J. Organomet. Chem. 2016, 813, 71–77.

[141] Li, M. X.; Zhang, D.; Zhang, L. Z.; Niu, J. Y.; Ji, B. S. Diorganotin(IV) Complexes with 2-Benzoylpyridine and 2-Acetylpyrazine N(4)-Phenylthiosemicarbazones: Synthesis, Crystal Structures and Biological Activities. J. Organomet. Chem. 2011, 696, 852–858.

[142] Wang, F.; Yin, H.; Cui, J.; Zhang, Y.; Geng, H.; Hong, M. Synthesis, Structural Characterization, In Vitro Cytotoxicities, DNA-Binding and BSA Interaction of Diorganotin(IV) Complexes Derived from Hydrazone Schiff Base. J. Organomet. Chem. 2014, 759, 83–91.

[143] Rehman, W.; Badshah, A.; Khan, S.; Tuyet, L. T. A. Synthesis, Characterization, Antimicrobial and Antitumor Screening of Some Diorganotin(IV) Complexes of 2-[(9H-Purin-6-ylimino)]-phenol. Eur. J. Med. Chem. 2009, 44, 3981–3985.

[144] Vicente-Dueñas, C.; Romero-Camarero, I.; Cobaleda, C.; Sánchez-García, I. Function of Oncogenes in Cancer Development: A Changing Paradigm. EMBO J. 2013, 32, 1502–1513.

[145] Shawish, H. B.; Wong, W. Y.; Wong, Y. L.; Loh, S. W.; Looi, C. Y.; Hassandarvish, P.; Phan, A. Y. L.; Wong, W. F.; Wang, H.; Paterson, I. C.; Ea, C. K.; Mustafa, M. R.; Maah, M. J. Nickel(II) Complex of Polyhydroxybenzaldehyde N4-Thiosemicarbazone Exhibits Anti-Inflammatory Activity by Inhibiting NF-κB Transactivation. PLoS One 2014, 9, 1–13.

[146] Koch, B.; Baul, T. S. B.; Chatterjee, A. Cell Proliferation Inhibition and Antitumor Activity of Novel Alkyl Series of Diorganotin(IV) Compounds. J. Appl. Toxicol.

202

2008, 28, 430–438.

[147] Marinovich, M.; Viviani, B.; Galli, C. L. Reversibility of Tributyltin-Chloride-Induced Protein Synthesis Inhibition after ATP Recovery in HEL-30 Cells. Toxicol. Lett. 1990, 52, 311–317.

[148] Lee, R. F. Metabolism of Tributyltin Oxide by Crabs, Oysters and Fish. Mar. Environ. Res. 1985, 17, 145–148.

[149] Kimmel, E. C.; Fish, R. H.; Casida, J. E. Bioorganotin Chemistry. Metabolism of Organotin Compounds in Microsomal Monooxygenase Systems and in Mammals. J. Agric. Food Chem. 1977, 25, 1–9.

[150] Basu Baul, T. S.; Paul, A.; Pellerito, L.; Scopelliti, M.; Singh, P.; Verma, P.; Duthie, A.; de Vos, D.; Tiekink, E. R. T. Dibutyltin(IV) Complexes Containing Arylazobenzoate Ligands: Chemistry, In Vitro Cytotoxic Effects on Human Tumor Cell Lines and Mode of Interaction with Some Enzymes. Invest. New Drugs 2011, 29, 285–299.

[151] Sirajuddin, M.; Ali, S.; Tahir, M. N. Pharmacological Investigation of Mono-, Di- and Tri-organotin(IV) Derivatives of Carbodithioates: Design, Spectroscopic Characterization, Interaction with SS-DNA and POM Analyses. Inorganica Chim. Acta 2016, 439, 145–158.

[152] Zeglis, B. M.; Pierre, V. C.; Barton, J. K.; Zeglis, B. M.; Pierre, V. C.; Barton, J. K. Metallo-Intercalators and Metallo-Insertors. Chem. Commun. 2007, 44, 4549–4696.

[153] Arjmand, F.; Sharma, G. C.; Sayeed, F.; Muddassir, M.; Tabassum, S. De Novo Design of Chiral Organotin Cancer Drug Candidates : Validation of Enantiopreferential Binding to Molecular Target DNA and 5’-GMP by UV–visible, Fluorescence, 1H and 31P NMR. J. Photochem. Photobiol. B Biol. 2011, 105, 167–174.

[154] Arjmand, F.; Yousuf, I. Synthesis, Characterization and In Vitro DNA Binding of Chromone Schiff Base Organotin(IV) Complexes. J. Organomet. Chem. 2013, 743, 55–62.

[155] Rehman, W.; Yasmeen, R.; Rahim, F.; Waseem, M.; Guo, C. Y.; Hassan, Z.; Rashid, U.; Ayub, K. Synthesis Biological Screening and Molecular Docking Studies of Some Tin(IV) Schiff Base Adducts. J. Photochem. Photobiol. B Biol. 2016, 164, 65–72.

[156] Doonan, F.; Cotter, T. G. Morphological Assessment of Apoptosis. Methods 2008, 44, 200–204.

[157] Chen, F.; Vallyathan, V.; Castranova, V.; Shi, X. Cell Apoptosis Induced by Carcinogenic Metals. Mol. Cell. Biochem. 2001, 222, 183–188.

[158] Chauhan, M.; Banerjee, K.; Arjmand, F. DNA Binding Studies of Novel Copper(II) Complexes Containing L-Tryptophan as Chiral Auxiliary: In Vitro

203

Antitumor Activity of Cu-Sn2 Complex in Human Neuroblastoma Cells. Inorg. Chem. 2007, 46, 3072–3082.

[159] Esmail, S. A. A.; Shamsi, M.; Chen, T.; Al-asbahy, W. M. Design, Synthesis and Characterization of Tin-Based Cancer Chemotherapy Drug Entity: In Vitro DNA Binding, Cleavage, Induction of Cancer Cell Apoptosis by Triggering DNA Damage-Mediated p53 Phosphorylation and Molecular Docking. Appl. Organomet. Chem. 2019, 33, 1–14.

[160] Nakatsu, Y.; Kotake, Y.; Ohta, S. Concentration Dependence of the Mechanisms of Tributyltin-Induced Apoptosis. Toxicol. Sci. 2007, 97, 438–447.

[161] Grondin, M.; Marion, M.; Denizeau, F.; Averill-bates, D. A. Tributyltin Induces Apoptotic Signaling in Hepatocytes through Pathways Involving the Endoplasmic Reticulum and Mitochondria. Toxicol. Appl. Pharmacol. 2007, 222, 57–68.

[162] Tada-oikawa, S.; Kato, T.; Kuribayashi, K.; Nishino, K.; Murata, M.; Kawanishi, S. Critical Role of Hydrogen Peroxide in the Differential Susceptibility of Th1 and Th2 Cells to Tributyltin-Induced Apoptosis. Biochem. Pharmacol. 2008, 75, 552–561.

[163] Hayyan, M.; Hashim, M. A.; Alnashef, I. M. Superoxide Ion : Generation and Chemical Implications. Chem. Rev. 2016, 116, 3026–3085.

[164] Devasagayam, T.; Tilak, J.; Boloor, K.; Sane, K.; Ghaskadbi, S.; Lele, R. Free Radicals and Antioxidants in Human Health: Current Status and Future Prospects. J. Assoc. Physicians India 2004, 52, 794–804.

[165] Liu, X.; Kim, C. N.; Yang, J.; Jemmerson, R.; Wang, X. Induction of Apoptotic Program in Cell-Free Extracts: Requirement for dATP and Cytochrome c. Cell 1996, 86, 147–157.

[166] Chow, S. C.; Orrenius, S. Rapid Cytoskeleton Modification in Thymocytes Induced by the Immunotoxicant Tributyltin. Toxicol. Appl. Pharmacol. 1994, 127, 19–26.

[167] Smith, P. J. Chemistry of Tin; Blackie Academic & Professionall: Glasgow, U.K, 1998.

[168] Das, V. G. K.; Gielen, M. Chemistry and Technology of Silicon and Tin; Oxford University Press: Oxford, 1992.

[169] Das, V. G. K. Main Group Elements and Their Compounds; Narosa Publishing House: New Delhi, India, 1996.

[170] Reed, J. C. Mechanisms of Apoptosis. Am. J. Pathol. 2000, 157, 1415–1430.

[171] Sasmal, A.; Garribba, E.; Ugone, V.; Rizzoli, C.; Mitra, S. Synthesis, Crystal Structures, EPR and DFT Studies of First Row Transition Metal Complexes of Lignin Model Compound Ethylvanillin. Polyhedron 2016, 121, 107–114.

[172] Singh, H. L.; Varshney, A. K. Synthesis and Characterization of Coordination

204

Compounds of Organotin(IV) with Nitrogen and Sulfur Donor Ligands. Appl. Organomet. Chem. 2001, 15, 762–768.

[173] Singh, R. V.; Chaudhary, P.; Chauhan, S.; Swami, M. Microwave-Assisted Synthesis, Characterization and Biological Activities of Organotin(IV) Complexes with Some Thio Schiff Bases. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2009, 72, 260–268.

[174] Manan, M. A. F. A.; Tahir, M. I. M.; Crouse, K. a.; How, F. N.-F.; Watkin, D. J. Synthesis, Characterization and Antibacterial Activity of Schiff Base Derived from S-Methyldithiocarbazate and Methylisatin. J. Chem. Crystallogr. 2012, 42, 173–179.

[175] Niu, M. J.; Li, Z.; Chang, G. L.; Kong, X. J.; Hong, M.; Zhang, Q. Crystal Structure, Cytotoxicity and Interaction with DNA of Zinc(II) Complexes with O-Vanillin Schiff Base Ligands. PLoS One 2015, 10, 1–14.

[176] Basha, M. T.; Chartres, J. D.; Pantarat, N.; Ali, M. A.; HuqMirza, A.; Kalinowski, D. S.; Richardson, D. R.; Bernhardt, P. V. Heterocyclic Dithiocarbazate Iron Chelators: Fe Coordination Chemistry and Biological Activity. Dalt. Trans. 2012, 41, 6536–6548.

[177] Yusof, E. N. M.; Jotani, M. M.; Tiekink, E. R. T.; Ravoof, T. B. S. A. 2-[(1 E)-([(Benzylsulfanyl)methanethioyl]aminoimino)methyl]-6-Methoxyphenol: Crystal Structure and Hirshfeld Surface Analysis. Acta Crystallogr. Sect. E Crystallogr. Commun. 2016, 72, 516–521.

[178] Yusof, E. N. M.; Latif, M. A. M.; Tahir, M. I. M.; Sakoff, J. A.; Simone, M. I.; Page, A. J.; Veerakumarasivam, A.; Tiekink, E. R. T.; Ravoof, T. B. S. A. O-Vanillin Derived Schiff Bases and Their Organotin(IV) Compounds: Synthesis, Structural Characterisation, In-Silico Studies and Cytotoxicity. Int. J. Mol. Sci. 2019, 20, 854.

[179] Ali, M. A.; Tarafder, M. T. . Metal Complexes of Sulphur and Nitrogen-Containing Ligands: Complexes of S-Benzyldithiocarbazate and a Schiff Base Formed by Its Condensation with Pyridine-2-carboxaldehyde. J. Inorg. Nucl. Chem. 1977, 39, 1785–1791.

[180] Agilent. CrysAlis PRO. Agilent Technologies: Yarnton, England 2011.

[181] Sheldrick, G. M. A Short History of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, A64, 112–122.

[182] Sheldrick, G. M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, C71, 3–8.

[183] Farrugia, L. J. WinGX and ORTEP for Windows: An Update. J. Appl. Crystallogr. 2012, 45, 849–854.

[184] Brandenburg, K. DIAMOND, Crystal Impact GbR. Bonn, Germany 2006.

205

[185] Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; R., J. C.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Caricato, M., Li, Xiaosong, Hratchian, H. P., Izmaylov, Artur F., Bloino, Julien, Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery Jr., J. A., Peralta, J. E., Ogliaro, François, Bearpark, Michael J., Heyd, Jochen, Brothers, E. N., Kudin, K. N., Staroverov, V. N., Kobayashi, Rika, Normand, J., Raghavachari, Krishnan, Rendell, Alistair P., Burant, J. C., Iyengar, S. S., Tomasi, Jacopo, Cossi, M., Rega, N., Millam, N. J., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, Ödön, Foresman, J. B., Ortiz, J. V., Cioslowski, J., Fox, Douglas J. Gaussian 09, Revision D.01. Wallingford CT. 2013.

[186] Roy, D.; Todd, K.; John, M. GaussView, Ver 5.0.9. Semichem Inc: Shawnee Mission, KS, USA, 2009.

[187] Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789.

[188] Becke, A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652.

[189] Jeffrey, P. M.; Damian, M.; Radom, L. An Evaluation of Harmonic Vibrational Frequency Scale Factors. J. Phys. Chem. A 2007, 111, 11683–11700.

[190] Chen, K.-Y.; Tsai, H.-Y. Synthesis, X-Ray Structure, Spectroscopic Properties and DFT Studies of a Novel Schiff Base. Int. J. Mol. Sci. 2014, 15, 18706–18724.

[191] Scalmani, G.; Frisch, M. J.; Mennucci, B.; Tomasi, J.; Cammi, R.; Barone, V. Geometries and Properties of Excited States in the Gas Phase and in Solution: Theory and Application of a Time-Dependent Density Functional Theory Polarizable Continuum Model. J. Chem. Phys. 2006, 124, 94107.

[192] Cancès, E.; Mennucci, B.; Tomasi, J. A New Integral Equation Formalism for the Polarizable Continuum Model: Theoretical Background and Applications to Isotropic and Anisotropic Dielectrics. J. Chem. Phys. 1997, 107, 3032–3041.

[193] Tomasi, J.; Persico, M. Molecular Interactions in Solution: An Overview of Methods Based on Continuous Distributions of the Solvent. Chem. Rev. 1994, 94, 2027–2094.

[194] Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3093.

[195] Mosmann, T. Rapid Colorimetric Assay for Cellular Growth and Survival: Application to Proliferation and Cytotoxicity Assays. J. Immunol. Methods 1983, 65, 55–63.

206

[196] Sakoff, J. A.; Ackland, S. P. Thymidylate Synthase Inhibition Induces S-Phase Arrest, Biphasic Mitochondrial Alterations and Caspase-Dependent Apoptosis in Leukaemia Cells. Cancer Chemother. Pharmacol. 2000, 46, 477–487.

[197] Bergman, A. M.; van Haperen, V. W. T. R.; Veerman, G.; Kuiper, C. M.; Peters, G. J. Synergistic Interaction between Cisplatin and Gemcitabine In Vitro. Clin. Cancer Res. 1996, 2, 521–530.

[198] Ali, M. A.; Butcher, R. J.; Bryan, J. C. Synthetic, Spectroscopic and X-Ray Crystallographic Structural Studies on Some Copper(II) Complexes of the 6-Methylpyridine-2-carboxaldehyde Schiff Base of S-Methyldithiocarbazate. Inorganica Chim. Acta 1999, 287, 8–13.

[199] Anto, R. J.; Sukumaran, K.; Kuttan, G.; Rao, M. N. A.; Subbaraju, V.; Kuttan, R. Anticancer and Antioxidant Activity of Synthetic Chalcones and Related Compounds. Cancer Lett. 1995, 97, 33–37.

[200] Awidat, S. A. K. Biological Activities and Molecular Analysis of Novel Dithiocarbazate Complex Compounds on Glioma Cell Lines, Ph.D. Dissertation, Universiti Putra Malaysia, Malaysia, 2005.

[201] Bacchi, A.; Bonardi, A.; Carcelli, M.; Mazza, P.; Pelagatti, P.; Pelizzi, C.; Pelizzi, G.; Solinas, C.; Zani, F. Organotin Complexes with Pyrrole-2,5-Dicarboxaldehyde Bis(Acylhydrazones). Synthesis, Structure, Antimicrobial Activity and Genotoxicity. J. Inorg. Biochem. 1998, 69, 101–112.

[202] Jain, M.; Gaur, S.; Singh, V. P.; Singh, R. V. Organosilicon(IV) and Organotin(IV) Complexes as Biocides and Nematicides : Synthetic, Spectroscopic and Biological Studies of N∩N Donor Sulfonamide Imine and Its Chelates. Main Gr. Met. Compd. 2004, 18, 73–82.

[203] Malhotra, R.; Ravesh, A.; Singh, V. Synthesis, Characterization, Antimicrobial Activities and QSAR Studies of Organotin(IV) Complexes. Phosphorus. Sulfur. Silicon Relat. Elem. 2017, 192, 73–80.

[204] Singh, R.; Kaushik, N. K. Spectral and Thermal Studies with Anti-Fungal Aspects of Some Organotin(IV) Complexes with Nitrogen and Sulphur Donor Ligands Derived from 2-Phenylethylamine. Spectrochim. Acta - Part A Mol. Biomol. Spectrosc. 2008, 71, 669–675.

[205] Sirajuddin, M.; Ali, S.; Mckee, V.; Sohail, M.; Pasha, H. Potentially Bioactive Organotin(IV) Compounds: Synthesis, Characterization, In Vitro Bioactivities and Interaction with SS-DNA. Eur. J. Med. Chem. 2014, 84, 343–363.

[206] Tariq, M.; Ali, S.; Muhammad, N.; Shah, N. A.; Sirajuddin, M.; Tahir, M. N.; Khalid, N.; Khan, M. R. Biological Screening, DNA Interaction Studies, and Catalytic Activity of Organotin(IV) 2-(4-Ethylbenzylidene) Butanoic Acid Derivatives: Synthesis, Spectroscopic Characterization and X-ray Structure. J. Coord. Chem. 2014, 67, 323–340.

[207] Win, Y. F.; Choong, C. S.; Dang, J. C.; Iqbal, M. A.; Quah, C. K.; Majid, A. M. S.

207

A.; Teoh, S. G. Polymeric Seven-Coordinated Organotin(IV) Complexes Derived from 5-Amino-2-Chlorobenzoic Acid and In Vitro Anti-Cancer Studies. J. Coord. Chem. 2014, 67, 3401–3413.

[208] Basu Baul, T. S.; Masharing, C.; Ruisi, G.; Jirásko, R.; Holčapek, M.; de Vos, D.; Wolstenholme, D.; Linden, A. Self-Assembly of Extended Schiff Base Amino Acetate Skeletons, 2-[(2Z)-(3-Hydroxy-1-Methyl-2-Butenylidene)]aminophenylpropionate and 2-[(E)-1-(2-Hydroxyaryl)alkylidene]aminophenylpropionate Skeletons Incorporating Organotin(IV) Moieties: Synthesis, Sp. J. Organomet. Chem. 2007, 692, 4849–4862.

[209] Katsoulakou, E.; Tiliakos, M.; Papaefstathiou, G.; Terzis, A.; Raptopoulou, C.; Geromichalos, G.; Papazisis, K.; Papi, R.; Pantazaki, A.; Kyriakidis, D.; Cordopatis, P.; Manessi-Zoupa, E. Diorganotin(IV) Complexes of Dipeptides Containing the α-Aminoisobutyryl Residue (Aib): Preparation, Structural Characterization, Antibacterial and Antiproliferative Activities of [(N-Bu)2Sn(H-1L)] (LH = H-Aib-L-Leu-OH, H-Aib-L-Ala-OH). J. Inorg. Biochem. 2008, 102, 1397–1405.

[210] Hou, H.-N.; Qi, Z.-D.; Ouyang, Y.-W.; Liao, F.-L.; Zhang, Y.; Liu, Y. Studies on Interaction between Vitamin B12 and Human Serum Albumin. J. Pharm. Biomed. Anal. 2008, 47, 134–139.

[211] Sielecki, T. M.; Boylan, J. F.; Benfield, P. A.; Trainor, G. L. Cyclin-Dependent Kinase Inhibitors: Useful Targets in Cell Cycle Regulation. J. Med. Chem. 2000, 43, 1–18.

[212] Rudorf, W. Reactions of Carbon Disulfide with C- Nucleophiles. Sulfur Reports 1991, 11, 51–141.

[213] Chandrasekhar, V.; Nagendran, S.; Baskar, V. Organotin Assemblies Containing Sn-O Bonds. Coord. Chem. Rev. 2002, 235, 1–52.

[214] Nath, M.; Pokharia, S.; Yadav, R. Organotin(IV) Complexes of Amino Acids and Peptides. Coord. Chem. Rev. 2001, 215, 99–149.

[215] Song, X.; Zapata, A.; Eng, G. Organotins and Quantitative-Structure Activity/property Relationships. J. Organomet. Chem. 2006, 691, 1756–1760.

[216] Spek, A. L. Structure Validation in Chemical Crystallography. Acta Crystallogr. Sect. D Biol. Crystallogr. 2009, D65, 148–155.

[217] Hay, P. J.; Wadt, W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for the Transition Metal Atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283.

[218] Hay, P. J.; Wadt, W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 299–310.

208

[219] Wadt, W. R.; Hay, P. J. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for Main Group Elements Na to Bi. J. Chem. Phys. 1985, 82, 284–298.

[220] Morris, G. M.; Huey, R.; Lindstrom, W.; Sanner, M. F.; Belew, R. K.; Goodsell, D. S.; Olson, A. J. Autodock4 and AutoDockTools4: Automated Docking with Selective Receptor Flexiblity. J. Comput. Chem. 2009, 16, 2785–2791.

[221] Besler, B. H.; Merz, K. M.; Kollman, P. A. Atomic Charges Derived from Semiempirical Methods. J. Comput. Chem. 1990, 11, 431–439.

[222] ADL: Parameters for docking with metal ions in receptor http://autodock.1369657.n2.nabble.com/ADL-Parameters-fordocking-with-metal-ions-in-receptor-td2505649.html (accessed Sep 7, 2016).

[223] Sudan, S.; Rupasinghe, H. V. Antiproliferative Activity of Long Chain Acylated Esters of Quercetin-3-O-Glucoside in Hepatocellular Carcinoma HepG2 Cells. Exp. Biol. Med. 2015, 240, 1452–1464.

[224] Hong, M.; Chang, G.; Li, R.; Niu, M. Anti-Proliferative Activity and DNA/BSA Interactions of Five Mono- or Di-organotin(IV) Compounds derived from 2-Hydroxy-N’-[(2-Hydroxy-3-Methoxyphenyl)methylidene]-Benzohydrazone. New J. Chem. 2016, 40, 7889–7900.

[225] Ganeshpandian, M.; Loganathan, R.; Suresh, E.; Riyasdeen, A.; Akbarsha, M. A.; Palaniandavar, M. New Ruthenium(II) Arene Complexes of Anthracenyl-Appended Diazacycloalkanes: Effect of Ligand Intercalation and Hydrophobicity on DNA and Protein Binding and Cleavage and Cytotoxicity. Dalt. Trans. 2014, 43, 1203–1219.

[226] Loganathan, R.; Ramakrishnan, S.; Suresh, E.; Riyasdeen, A.; Akbarsha, M. A.; Palaniandavar, M. Mixed Ligand Copper(II) Complexes of N, N-Bis(benzimidazol-2-ylmethyl)amine (BBA) with Diimine Co-Ligands: Efficient Chemical Nuclease and Protease Activities and Cytotoxicity. Inorg. Chem. 2012, 51, 5512–5532.

[227] Malhotra, R.; Singh, J. P.; Dudeja, M.; Dhindsa, K. S. Ligational Behavior of N-Substituted Acid Hydrazides towards Transition Metals and Potentiation of Their Microbiocidal Activity. J. Inorg. Biochem. 1992, 46, 119–127.

[228] Ali, M. A.; Mirza, A. H.; Butcher, R. J. Synthesis and Characterization of Copper(II) Complexes of the Methylpyruvate Schiff Base of S-Methyldithiocarbazate (Hmpsme) and the X-Crystal Structures of Hmpsme and [Cu(mpsme)Cl]. Polyhedron 2001, 20, 1037–1043.

[229] Naqeebullah; Farina, Y.; Chan, K. M.; Mun, L. K.; Rajab, N. F.; Ooi, T. C. Diorganotin(IV) Derivatives of N-Methyl P-fluorobenzo-hydroxamic Acid: Preparation, Spectral Characterization, X-Ray Diffraction Studies and Antitumor Activity. Molecules 2013, 18, 8696–8711.

[230] Holeček, J.; Nádvorník, M.; Handlíř, K.; Lyčka, A. 13C and 119Sn NMR Spectra of

209

Di-N-butyltin(IV) Compounds. J. Organomet. Chem. 1986, 315, 299–308.

[231] Zhang, Y.-Y.; Zhang, R.-F.; Zhang, S.-L.; Cheng, S.; Li, Q.-L.; Ma, C.-L. Syntheses, Structures and Anti-Tumor Activity of Four New Organotin(IV) Carboxylates Based on 2-Thienylselenoacetic Acid. Dalt. Trans. 2016, 45, 8412–8421.

[232] Barreiro, S.; Durán-Carril, M. L.; Viqueira, J.; Sousa-Pedrares, A.; García-Vázquez, J. A.; Romero, J. Structural Studies and Bioactivity of Diorganotin(IV) Complexes of Pyridin-2-thionato Derivatives. J. Organomet. Chem. 2015, 791, 155–162.

[233] Ma, C.; Zhang, J.; Tian, G.; Zhang, R. Syntheses, Crystal Structures and Coordination Modes of Tri- and Di-Organotin Derivatives with 2-Mercapto-4-Methylpyrimidine. J. Organomet. Chem. 2005, 690, 519–533.

[234] Yusof, E. N. M.; Jotani, M. M.; Tiekink, E. R. T.; Ravoof, T. B. S. A. 2-[(1E)-([(Benzylsulfanyl)methanethioyl]amino-imino)methyl]-6-methoxyphenol: Crystal Structure and Hirshfeld Surface Analysis. Acta Crystallogr. Sect. E Crystallogr. Commun. 2016, 72, 516–521.

[235] Addison, A. W.; Rao, T. N. Synthesis, Structure, and Spectroscopic Properties of Copper(II) Compounds Containing Nitrogen-Sulphur Donor Ligands ; the Crystal and Molecular Structure of Aqua[l,7-bis(N-Methylbenzimidazol-2’-yl)-2,6-dithiaheptane]copper(II) Perchlorate. J. Chem. Soc. Dalt. Trans. 1984, 7, 1349–1356.

[236] Milčič, M. K.; Medakovič, V. B.; Sredojevič, D. N.; Juranič, N. O.; Zarič, S. D. Electron Delocalization Mediates the Metal-Dependent Capacity for CH/π Interactions of Acetylacetonato Chelates. Inorg. Chem. 2006, 45, 4755–4763.

[237] Tiekink, E. R. T. Supramolecular Assembly Based on “emerging” Intermolecular Interactions of Particular Interest to Coordination Chemists. Coord. Chem. Rev. 2017, 345, 209–228.

[238] Chen, D.; Lai, C. S.; Tiekink, E. R. T. Crystal Structures of 2,2’-bipyridine Adducts of Two Cadmium O-Alkyl Dithiocarbonates: Rationalisation of Disparate Coordination Geometries Based on Different Crystal Packing Environments. Zeitschrift fur Krist. 2002, 217, 747–752.

[239] Malenov, D. P.; Janjić, G. V.; Medaković, V. B.; Hall, M. B.; Zarić, S. D. Noncovalent Bonding: Stacking Interactions of Chelate Rings of Transition Metal Complexes. Coord. Chem. Rev. 2017, 345, 318–341.

[240] González-garcía, C.; Mata, A.; Zani, F.; Mendiola, M. A.; López-torres, E. Synthesis and Antimicrobial Activity of Tetradentate Ligands Bearing Hydrazone and / or Thiosemicarbazone Motifs and Their Diorganotin (IV) Complexes. J. Inorg. Biochem. 2016, 163, 118–130.

[241] Andree, H. A. M.; Reutelingsperger, C. P. M.; Hauptmann, R.; Hemker, H. C.; Hermens, W. T.; Willems, G. M. Binding of Vascular Anticoagulant Alpha (VAC

210

Alpha) to Planar Phospholipid Bilayers. J. Biol. Chem. 1990, 265 (9), 4923–4928.

[242] Gea, R.; Ma, W.-H.; Li, Y.-L.; Li, Q.-S. Apoptosis Induced Neurotoxicity of Di-N-butyl-di-(4-chlorobenzohydroxamato) Tin(IV) via Mitochondria-Mediated Pathway in PC12 Cells. Toxicol. Vitr. 2013, 27, 92–102.

[243] Poon, I. K. H.; Hulett, M. D.; Parish, C. R. Molecular Mechanisms of Late Apoptotic/necrotic Cell Clearance. Cell Death Differ. 2010, 17, 381–397.

[244] Prosser, K. E.; Chang, S. W.; Saraci, F.; Le, P. H.; Walsby, C. J. Anticancer Copper Pyridine Benzimidazole Complexes: ROS Generation, Biomolecule Interactions and Cytotoxicity. J. Inorg. Biochem. 2017, 167, 89–99.

[245] Fani, S.; Kamalidehghan, B.; Lo, K. M.; Hashim, N. M.; Chow, K. M.; Ahmadipour, F. Synthesis, Structural Characterization, and Anticancer Activity of a Monobenzyltin Compound against MCF-7 Breast Cancer Cells. Drug Des. Devel. Ther. 2015, 9, 6191–6201.

[246] Hussain, S.; Bukhari, I. H.; Ali, S.; Shahzadi, S.; Shahid, M.; Munawar, K. S. Synthesis and Spectroscopic and Thermogravimetric Characterization of Heterobimetallic Complexes with Sn(IV) and Pd(II); DNA Binding, Alkaline Phosphatase Inhibition and Biological Activity Studies. J. Coord. Chem. 2015, 68, 662–677.

[247] Hussain, S.; Ali, S.; Shahzadi, S.; Tahir, M. N.; Shahid, M. Synthesis, Characterization, Biological Activities, Crystal Structure and DNA Binding of Organotin(IV) 5-Chlorosalicylates. J. Coord. Chem. 2015, 68, 2369–2387.

[248] Yadav, S.; Yousuf, I.; Usman, M.; Ahmad, M.; Arjmand, F.; Tabassum, S. Synthesis and Spectroscopic Characterization of Diorganotin(IV) Complexes of N′-(4-Hydroxypent-3-en-2-ylidene)isonicotinohydrazide: Chemotherapeutic Potential Validation by In Vitro Interaction Studies with DNA/HSA, DFT, Molecular Docking and Cytotoxic Ac. RSC Adv. 2015, 5, 50673–50690.

[249] Aldridge, W. H. The Influence of Organotin Compounds on Mitochondrial Functions. In Organotin Compounds: New Chemistry and Applications; Zuckerman, J. J., Ed.; American Chemical Society: Washington, DC, 1976; pp 186–196.

[250] Ventrella, V.; Nesci, S.; Trombetti, F.; Bandiera, P.; Pirini, M.; Borgatti, A. R.; Pagliarani, A. Tributyltin Inhibits the Oligomycin-Sensitive Mg-ATPase Activity in Mytilus Galloprovincialis Digestive Gland Mitochondria. Comp. Biochem. Physiol. Part C 2011, 153, 75–81.

[251] Saxena, A. K.; Huber, F. Organotin Compounds and Cancer Chemotheraphy. Coord. Chem. Rev. 1989, 95, 109–123.

[252] Murdock, K. C.; Child, R. G.; Fabio, P. F.; Angier, R. B.; Wallace, R. E.; Durr, F. E.; Citarella, R. V. Antitumor Agents. 1. 1,4-Bis[(aminoalkyl)amino]-9,10-Anthracenediones. J. Med. Chem. 1979, 22, 1024–1030.

211

[253] Pereañez, J. A.; Núñez, V.; Patiño, A. C.; Londoño, M.; Quintana, J. C. Inhibitory Effects of Plant Phenolic Compounds on Enzymatic and Cytotoxic Activities Induced By a Snake Venom Phospholipase A2. Vitae 2011, 19, 295–304.

[254] Johnson, M. G.; Kiyokawa, H.; Tani, S.; Koyama, J.; Morris-natschke, S. L.; Bowers-daines, A. M. M. M.; Lange, C.; Lee, K. Antitumor Agents-CLXVII. Synthesis and Structure-Activity Correlations of the Cytotoxic Anthraquinone 1,4-Bis-(2,3-Epoxypropylamino)-9,10-Anthracenedione, and of Related Compounds. Bioorg. Med. Chem. 1997, 5, 1469–1479.

[255] Kim, K. J.; Kim, M. A.; Jung, J. H. Antitumor and Antioxidant Activity of Protocatechualdehyde Produced from Streptomyces Lincolnensis M-20. Arch. Pharm. Res. 2008, 31, 1572–1577.

[256] Lee, B. H.; Yoon, S. H.; Kim, Y. S.; Kim, S. K.; Moon, B. J.; Bae, Y. S. Apoptotic Cell Death through Inhibition of Protein Kinase CKII Activity by 3,4-Dihydroxybenzaldehyde Purified from Xanthium Strumarium. Nat. Prod. Res. 2008, 22, 1441–1450.

[257] Fayed, A. M.; Elsayed, S. A.; El-Hendawy, A. M.; Mostafa, M. R. Complexes of Cis-dioxomolybdenum(VI) and Oxovanadium(IV) with a Tridentate ONS Donor Ligand: Synthesis, Spectroscopic Properties, X-Ray Crystal Structure and Catalytic Activity. Spectrochim. Acta - Part A Mol. Biomol. Spectrosc. 2014, 129, 293–302.

[258] Rigaku Oxford Diffraction. CrysAlis PRO. Agilent Technologies Inc. Santa Clara, CA, USA p 2015.

[259] Yusof, E. N. M.; Tiekink, E. R. T.; Jotani, M. M.; Simone, M. I.; Page, A. J.; Ravoof, T. B. S. A. Synthesis, Characterisation and Structure Determination of 3-[(1Z)-2-ylidenemethyl]benzene-1,2-diol. J. Mol. Struct. 2018, 1171, 650–657.

[260] Mıhçıokur, Ö.; Özpozan, T. Molecular Structure, Vibrational Spectroscopic Analysis (IR & Raman), HOMO-LUMO and NBO Analysis of Anti-Cancer Drug Sunitinib Using DFT Method. J. Mol. Struct. 2017, 1149, 27–41.

[261] Groom, C. R.; Bruno, I. J.; Lightfoot, M. P.; Ward, S. C. The Cambridge Structural Database. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2016, 72, 171–179.

[262] Frija, L. M. T.; Pombeiro, A. J. L.; Kopylovich, M. N. Coordination Chemistry of Thiazoles, Isothiazoles and Thiadiazoles. Coord. Chem. Rev. 2016, 308, 32–55.

[263] Javed, F.; Ali, S.; Shahzadi, S.; Sharma, S. K.; Qanungo, K.; Munawar, K. S.; Khan, I. Synthesis, Characterization, and Biological Activity of Organotin(IV) Complexes with 4-Oxo-4-[3-(Trifluoromethyl)phenylamino]butanoic Acid. Russ. J. Gen. Chem. 2017, 87, 2409–2420.

[264] Saxena, A.; Tandon, J. P.; Molloy, K. C.; Zuckerman, J. J. Tin(IV) Complexes of Tridentate Schiff Bases Having ONS Donor Systems. Inorganica Chim. Acta 1982, 63, 71–74.

212

[265] Degaonkar, M. P.; Gopinathan, S.; Gopinathan, C. Preparation and Characterisation of Some New Dithioca Bazate Schiff Base Complexes of Titanium(IV), Tin(IV) and Lead(IV). Indian J. Chem. 1989, 28, 678–682.

[266] Degaonkar, M. P.; Gopinathan, S.; Gopinathan, C. Synthesis and Characterisation of Titanium(IV), Tin(IV) and Lead(IV) Complexes of Schiff Bases Derived from S-Benzyldithiocarbazate. Synth. React. Inorg. Met. Chem. 1989, 19, 613–625.

[267] Shi, Y. C.; Yang, H. M.; Song, H. Bin; Yan, C. G.; Hu, X. Y. Syntheses and Crystal Structures of the Potential Tridentate Ligand Formed from Condensation of Ferrocenoylacetone and S-Benzyldithiocarbazate and Its Bivalent Metal Complexes. Polyhedron 2004, 23, 567–573.

[268] Yin, H. D.; Hong, M.; Li, G.; Wang, D. Q. Synthesis, Characterization and Structural Studies of Diorganotin(IV) Complexes with Schiff Base Ligand Salicylaldehyde Isonicotinylhydrazone. J. Organomet. Chem. 2005, 690, 3714–3719.

[269] Wrackmeyer, B. l19Sn-NMR Parameters. Annu. Reports NMR Spectrosc. 1985, 16, 73–186.

[270] Sousa, G. F. de; Gatto, C. C.; Ellena, J.; Ardisson, J. D. Synthesis and Crystal Structures of Octahedral Sn(IV) Complexes Prepared from SnCl2·2H2O and 2-Hydroxyacetophenone (S-Benzydithiocarbazate) Ligand (H2L). J. Chem. Crystallogr. 2011, 41, 838–842.

[271] Nakanishi, T.; Masuda, A.; Suwa, M.; Akiyama, Y.; Hoshino-Abe, N.; Suzuki, M. Synthesis of Derivatives of NK109, 7-OH Benzo[c]phenanthridine Alkaloid, and Evaluation of Their Cytotoxicities and Reduction-Resistant Properties. Bioorganic Med. Chem. Lett. 2000, 10, 2321–2323.

[272] Maia, P. I. da S.; Pavan, F. R.; Leite, C. Q. F.; Lemos, S. S.; de Sousa, G. F.; Batista, A. A.; Nascimento, O. R.; Ellena, J.; Castellano, E. E.; Niquet, E.; Deflona, V. M. Vanadium Complexes with Thiosemicarbazones: Synthesis, Characterization, Crystal Structures and Anti-Mycobacterium Tuberculosis Activity. Polyhedron 2009, 28, 398–406.

[273] Wallace, R.; Richard Thomson, J.; Bell, M. J.; Skipper, H. E. Observations on the Antileukemic Activity of Pyridine- 2-Carboxaldehyde Thiosemicarbazone and Thiocarbohydrazone. Cancer Res. 1956, 16, 167–170.

[274] French, F. A.; Blanz, E. J. The Carcinostatic Activity of α-(N) Heterocyclic Carboxaldehyde Thiosemicarbazones. Cancer Res. 1965, 25, 1454–1458.

[275] West, D. X.; Liberta, A. E.; Padhye, S. B.; Chikate, R. C.; Sonawane, P. B.; Kumbhar, A. S.; Yerande, R. G. Thiosemicarbazone Complexes of Copper(II): Structural and Biological Studies. Coord. Chem. Rev. 1993, 123, 49–71.

[276] Shanmugapriya, A.; Jain, R.; Sabarinathan, D.; Kalaiarasi, G.; Dallemer, F.; Prabhakaran, R. Structurally Different Mono-, Bi- and Trinuclear Pd(II) Complexes and Their DNA/protein Interaction, DNA Cleavage, and Anti-Oxidant,

213

Anti-Microbial and Cytotoxic Studies. New J. Chem. 2017, 41, 10324–10338.

[277] Kalaiarasi, G.; Jain, R.; Shanmugapriya, A.; Puschman, H.; Kalaivani, P.; Prabhakaran, R. New Binuclear Ni(II) Metallates as Potent Antiproliferative Agents against MCF-7 and HeLa Cells. Inorganica Chim. Acta 2017, 462, 174–187.

[278] Kalaivani, P.; Prabhakaran, R.; Ramachandran, E.; Dallemer, F.; Paramaguru, G.; Renganathan, R.; Poornima, P.; Vijaya Padma, V.; Natarajan, K. Influence of Terminal Substitution on Structural, DNA, Protein Binding, Anticancer and Antibacterial Activities of palladium(II) Complexes Containing 3-Methoxy Salicylaldehyde-4(N) Substituted Thiosemicarbazones. Dalt. Trans. 2012, 41, 2486.

[279] Kalaiarasi, G.; Jain, R.; Puschman, H.; Poorna Chandrika, S.; Preethi, K.; Prabhakaran, R. New Binuclear Ni(II) Metallates Containing ONS Chelators: Synthesis, Characterisation, DNA Binding, DNA Cleavage, Protein Binding, Antioxidant Activity, Antimicrobial and In Vitro Cytotoxicity. New J. Chem. 2017, 41, 2543–2560.

[280] Swesi, A. T.; Farina, Y.; Baba, I. Synthesis and Characterization of Some diorganotin(IV) Complexes of 2,3-Dihydroxybenzaldehyde N(4)-Substituted Thiosemicarbazone. Sains Malaysiana 2007, 36, 21–26.

[281] Swesi, A. T.; Farina, Y.; Kassim, M.; Ng, S. W. 2,3-Dihydroxybenzaldehyde Thiosemicarbazone Hemihydrate. Acta Crystallogr. Sect. E Struct. Reports Online 2006, E62, o5457–o5458.

[282] Vrdoljak, V.; Cindrić, M.; Milić, D.; Matković-Čalogović, D.; Novak, P.; Kamenar, B. Synthesis of Five New Molybdenum(VI) Thiosemicarbazonato Complexes. Crystal Structures of Salicylaldehyde and 3-Methoxy-Salicylaldehyde 4-Methylthiosemicarbazones and Their molybdenum(VI) Complexes. Polyhedron 2005, 24, 1717–1726.

[283] Rocha, F. V.; Barra, C. V.; Mauro, A. E.; Carlos, I. Z.; Nauton, L.; El Ghozzi, M.; Gautier, A.; Morel, L.; Netto, A. V. G. Synthesis, Characterization, X-Ray Structure, DNA Cleavage and Cytotoxic Activities of Palladium(II) Complexes of 4-Phenyl-3-Thiosemicarbazide and Triphenylphosphane. Eur. J. Inorg. Chem. 2013, 25, 4499–4505.

[284] Matsuda, Y.; Ebata, T.; Mikami, N. Vibrational Spectroscopy of 2-Pyridone and Its Clusters in Supersonic Jets: Structures of the Clusters as Revealed by Characteristic Shifts of the NH and C=O Bands. J. Chem. Phys. 1999, 110, 8397.

[285] Shawish, H. B.; Paydar, M.; Looi, C. Y.; Wong, Y. L.; Movahed, E.; Halim, S. N. A.; Wong, W. F.; Mustafa, M. R.; Maah, M. J. Nickel(II) Complexes of Polyhydroxybenzaldehyde N4-Thiosemicarbazones: Synthesis, Structural Characterization and Antimicrobial Activities. Transit. Met. Chem. 2014, 39, 81–94.

[286] Shaheen, F.; Sirajuddin, M.; Ali, S.; Zia-ur-Rehman; Dyson, P. J.; Shah, N. A.;

214

Tahir, M. N. Organotin(IV) 4-(benzo[d][1,3]dioxol-5-ylmethyl)piperazine-1-carbodithioates: Synthesis, Characterization and Biological Activities. J. Organomet. Chem. 2018, 856, 13–22.

[287] Miskolci, C.; Labádi, I.; Kurihara, T.; Motohashi, N.; Molnár, J. Guanine–cytosine Rich Regions of Plasmid DNA Can Be the Target in Anti-Plasmid Effect of Phenothiazines. Int. J. Antimicrob. Agents 2002, 14, 243–247.

[288] Foye, W. O. Cancer Chemotherapeutic Agents; American Chemical Society: Washington, DC, 1995.

[289] Sishc, B. J.; Nelson, C. B.; Mckenna, M. J.; Battaglia, C. L. R.; Tanzarella, C. Telomeres and Telomerase in the Radiation Response: Implications for Instability, Reprograming, and Carcinogenesis. Front. Oncol. 2015, 5, 1–19.

[290] Veronese, F. M.; Pasut, G. PEGylation, Successful Approach to Drug Delivery. Drug Discov. Today 2005, 10, 1451–1458.

[291] Jong, J. H. De; Rodermond, H. M.; Zimberlin, C. D.; Lascano, V.; Melo, F. D. S. E.; Richel, D. J.; Medema, J. P.; Vermeulen, L. Fusion of Intestinal Epithelial Cells with Bone Marrow Derived Cells Is Dispensable for Tissue Homeostasis. Sci. Rep. 2012, 2, 271.

[292] Szakacs, G.; Varadi, A.; Ozvegy-Laczka, C.; Sarkadi, B. The Role of ABC Transporters in Drug Absorption, Distribution, Metabolism, Excretion and Toxicity (ADME–Tox). Drug Discov. Today 2008, 13, 379–393.

[293] Hill, A. J.; Teraoka, H.; Heideman, W.; Peterson, R. E. Zebrafish as a Model Vertebrate for Investigating Chemical Toxicity. Toxicol. Sci. 2005, 86, 6–19.

[294] Zon, L. I.; Peterson, R. T. In Vivo Drug Discovery in the Zebrafish. Nat. Rev. Drug Discov. 2005, 4, 35.

[295] Garberoglio, G. OBGMX: A Web-Based Generator of GROMACS Topologies for Molecular and Periodic Systems Using the Universal Force Field. J. Comput. Chem. 2012, 33, 2204–2208.

[296] Lindorff-Larsen, K.; Piana, S.; Palmo, K.; Maragakis, P.; Klepeis, J. L.; Dror, R. O.; Shaw, D. E. Improved Side-Chain Torsion Potentials for the Amber ff99SB Protein Force Field. Proteins Struct. Funct. Bioinforma. 2010, 78, 1950–1958.

[297] Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of Simple Potential Functions for Simulating Liquid Water. J. Chem. Phys. 1983, 79, 926–936.

215

APPENDICES

Copyright Permission for Figure 1.1:

216

Copyright Permission for Figure 1.3:

217

218

Copyright Permission for Figure 1.8:

219

Copyright Permission for Figure 1.19:

220

Copyright Permission for Figure 1.20:

221

Copyright Permission for Figure 1.21:

222

Copyright Permission Contents Presented in Chapter 2:

223

Copyright Permission Contents Presented in Chapter 2 and 3:

224

APPENDIX - ANALYSES

Melting Points

Melting points were determined using an Electrothermal digital melting point apparatus

at Department of Chemistry, Universiti Putra Malaysia.

Elemental (CHN) Analyses

CHNS was determined using a LECO CHNS-932 instrument at Department of

Chemistry, Universiti Putra Malaysia and Thermo Flash EA110 elemental analyser at

Department of Chemistry, Universiti Teknologi Mara, Malaysia. The samples were

weighed to around 1.8-2.1 mg into the tin capsule, crimped and then combusted under

excess of oxygen. Sulphamethazine was used as a calibration standard before analyses.

The presented results obtained are based on weight percent composition.

Molar conductivity Analyses

The tin(IV) compounds were diluted into 1x10-3 concentrations in DMSO. Molar

conductance was determined by immersing a dip-type cell with platinised electrode into

the tin(IV) compounds solution at 28°C using a 4310 Jenway Conductivity Meter at

Department of Chemistry, Universiti Putra Malaysia. The molar conductivity values were

determined by using the following equation:

Molar conductance, ΛM = 1000R/C[Λsolution - Λsolvent] x 10-6

Where; R = cell constant = 1 C = Concentration (0.001 M) Λsolution = Conductivity of solution Λsolvent = Conductivity of solvent

225

Fourier Transformed Infrared (FTIR) Spectroscopic Analyses

Infrared spectra were recorded using the Perkin Elmer Spectrum 100 with Universal ATR

Polarization in the range 4000-280 cm-1 at Department of Chemistry, Universiti Putra

Malaysia. All the spectra were recorded at room temperature.

Ultraviolet-visible (UV-vis) Spectral Analyses

The tin(IV) compounds were weighed accurately and dissolved in DMSO as a solvent.

The compounds were prepared in three different concentration (10-3 M, 10-4 M, and 10-5

M). Electronic spectra were recorded on a Shimadzu UV-1650 PC recording

spectrophotometer (1000–200 nm) at Department of Chemistry, Universiti Putra

Malaysia.

Mass spectroscopic (MS) Analyses

The mass spectra were recorded using a Shimadzu GC-MS QP2010Plus mass

spectrometer at Department of Chemistry, Universiti Putra Malaysia.

Multinuclear (1H, 13C and 119Sn) Nuclear Magnetic Resonance (NMR) spectroscopic

analyses

1H and 13C NMR spectra were recorded using NMR JNM ECA400 spectrometer at

Universiti Putra Malaysia and tetramethylsilane (TMS) was used as an internal standard.

At room temperature, 5-20 mg of compounds were weighed and dissolved in DMSO-d6

or CDCl3. 119Sn NMR of compounds 4-9 and 19-21 were recorded using a JOEL NMR

spectrophotometer at Universiti Malaya, Malaysia and 119Sn NMR of compounds 13-18

and 22-40 were measured using Bruker BioSpin Avance III (600MHz) spectrometer at

The University of Newcastle, Australia.

226

Cytotoxic Assay

The cytotoxic assay against bladder cancer cell lines (EJ-28 and RT-112) was evaluated

at the Molecular Genetics Laboratory, Department of Biomedical Sciences, Universiti

Putra Malaysia. For HT29 (colon), U87 and SJ-G2 (glioblastoma), MCF-7 (breast),

A2780 (ovarian), H460 (lung), A431 (skin), Du145 (prostate), BE2-C (neuroblastoma),

MIA (pancreas) cell lines and one normal breast cell line, MCF-10A (normal breast), the

MTT assays were performed at Department of Medical Oncology, Calvary Mater

Newcastle Hospital, Australia.

227

APPENDIX FOR CHAPTER 2 (A2)

Table A2.1: Experimental and calculated FTIR vibrations (cm-1) for the Schiff bases (1-3).

Compound Method IR bands (cm-1) v(o-OH) v(NH) v(C=N) v(N-N) v(C=S)

1

Experimental - 3084 1600 1117 1026 B3LYP/6-311G(d,p) 3419 3377 1596 1112 1024

2

Experimental - 3092 1598 1118 1030 B3LYP/6-311G(d,p) 3415 3378 1596 1112 1024

3

Experimental - 3090 1598 1125 1030 B3LYP/6-311G(d,p) 3419 3377 1596 1112 1025

Table A2.2: 1H NMR spectral data for the Schiff bases (1-3).

Compound 1H NMR Assignment, δ (ppm)

NH OH CH CH2 O-CH3 Ar- CH3 Aromatic Protons

1 13.34 (s, 1H)

9.57 (s, 1H)

8.51 (s, 1H)

4.40 (s, 2H) 3.76 (s, 3H) 2.30

(s, 3H) 6.75-7.34 (m,7H)

2 13.32 (s, 1H)

9.61 (s, 1H)

8.51 (s, 1H)

4.39 (s, 2H) 3.76 (s, 3H) 2.23

(s, 3H) 6.97-9.79 (m,7H)

3 13.34 (s, 1H)

9.58 (s, 1H)

8.52 (s, 1H)

4.45 (s, 2H) 3.77 (s, 3H) - 6.77-7.37 (m,8H)

228

Table A2.3: 13C1H NMR spectral data for the Schiff bases (1-3).

Compound Solvent 13C1H NMR Assignment, δ (ppm)

C=S/ C-S C=N O-CH3 CH2 Ar-CH3 Aromatic carbons

1 DMSO-d6 196.1 148.6 56.4 36.9 19.4 114.4, 118.8, 120.0, 126.7, 128.2, 130.7, 130.8, 134.4, 137.4, 144.9, 147.4, 148.6

2 DMSO-d6 196.2 148.6 56.4 38.0 21.2 114.4, 118.8, 120.0, 129.6, 129.7, 134.0, 137.0, 145.0, 147.4, 148.6

3 DMSO-d6 196.1 148.6 56.4 38.1 - 114.4, 118.8, 120.0, 127.8, 129.0, 129.8, 137.3, 144.9, 147.4, 148.6

229

Figure A2.1: GCMS spectra of (a) 1, (b) 2 and (c) 3.

S NH

NS

HOO

C16H16N2O2S2Mol. Wt.: 332.44

(c)

S NH

NS

HOO

C17H18N2O2S2Mol. Wt.: 346.47

S NH

NS

HOO

C17H18N2O2S2Mol. Wt.: 346.47

(a)

(b)

230

Table A2.4: Experimental and calculated UV-visible absorption data for the Schiff bases (1-3).

Compounds Wavelength (nm) Experimental B3LYP/6-311G(d,p)

1 371 377 344 334

2 389 378 348 334

3 371 378 340 334

231

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes. Table A2.5: Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) of 3. x y z Uiso */Ueq S1 0.22247 (2) 0.46299 (12) 0.61684 (2) 0.02185 (15) S2 0.17478 (3) 0.80851 (12) 0.54164 (2) 0.02258 (15) O1 0.35672 (7) −0.0373 (3) 0.63252 (4) 0.0239 (3) H1O 0.3342 (10) 0.086 (4) 0.6194 (7) 0.036* O2 0.43677 (8) −0.4115 (3) 0.66547 (5) 0.0305 (4) N1 0.27458 (8) 0.4630 (4) 0.54712 (5) 0.0203 (4) H1N 0.2819 (10) 0.519 (5) 0.5223 (4) 0.024* N2 0.31479 (8) 0.2682 (4) 0.56753 (5) 0.0194 (4) C1 0.22604 (9) 0.5777 (4) 0.56594 (6) 0.0179 (4) C2 0.15287 (10) 0.6579 (5) 0.63209 (6) 0.0237 (5) H2A 0.1579 0.8639 0.6268 0.028* H2B 0.1130 0.5907 0.6161 0.028* C3 0.14873 (10) 0.6059 (5) 0.67741 (6) 0.0223 (5) C4 0.18779 (13) 0.7499 (6) 0.70649 (8) 0.0395 (6) H4 0.2181 0.8838 0.6979 0.047* C5 0.18365 (13) 0.7029 (6) 0.74845 (8) 0.0432 (7) H5 0.2108 0.8059 0.7682 0.052* C6 0.14056 (12) 0.5086 (5) 0.76139 (7) 0.0356 (6) H6 0.1381 0.4735 0.7900 0.043* C7 0.10141 (14) 0.3670 (7) 0.73262 (8) 0.0502 (8) H7 0.0708 0.2346 0.7413 0.060* C8 0.10538 (13) 0.4127 (6) 0.69076 (8) 0.0404 (7) H8 0.0779 0.3096 0.6712 0.048* C9 0.35930 (10) 0.1553 (4) 0.54759 (6) 0.0192 (4) H9 0.3635 0.2118 0.5199 0.023* C10 0.40308 (9) −0.0564 (4) 0.56664 (6) 0.0188 (4) C11 0.40002 (9) −0.1423 (4) 0.60766 (6) 0.0191 (4) C12 0.44376 (10) −0.3469 (5) 0.62508 (7) 0.0236 (5) C13 0.48986 (10) −0.4610 (5) 0.60167 (7) 0.0282 (5) H13 0.5196 −0.5984 0.6134 0.034* C14 0.49292 (11) −0.3747 (5) 0.56070 (7) 0.0286 (5) H14 0.5246 −0.4541 0.5448 0.034* C15 0.45042 (10) −0.1762 (5) 0.54345 (7) 0.0239 (5) H15 0.4529 −0.1189 0.5156 0.029* C16 0.47583 (12) −0.6366 (5) 0.68359 (8) 0.0346 (6) H16A 0.4679 −0.8109 0.6674 0.052* H16B 0.4649 −0.6693 0.7119 0.052* H16C 0.5215 −0.5840 0.6839 0.052*

232

Table A2.6: Atomic displacement parameters (Å2) of 3. U11 U22 U33 U12 U13 U23 S1 S2 O1 O2 N1 N2 C1 C2 C3 C4 C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16

0.0221 (3) 0.0225 (3) 0.0258 (8) 0.0399 (9) 0.0206 (9) 0.0180 (8) 0.0181 (10) 0.0239 (11) 0.0234 (11) 0.0430 (15) 0.0523 (17) 0.0411 (14) 0.0549 (18) 0.0493 (16) 0.0225 (10) 0.0184 (10) 0.0185 (10) 0.0264 (11) 0.0239 (11) 0.0228 (11) 0.0223 (11) 0.0411 (14)

0.0271 (3) 0.0273 (3) 0.0281 (9) 0.0269 (9) 0.0253 (10) 0.0208 (9) 0.0179 (10) 0.0257 (12) 0.0231 (12) 0.0457 (16) 0.0504 (17) 0.0470 (16) 0.066 (2) 0.0480 (16) 0.0203 (11) 0.0170 (10) 0.0189 (11) 0.0194 (11) 0.0238 (12) 0.0301 (13) 0.0269 (12) 0.0237 (13)

0.0169 (3)0.0180 (3) 0.0183 (7) 0.0238 (8) 0.0154 (8) 0.0194 (9) 0.0176 (10) 0.0222 (11) 0.0212 (11) 0.0293 (13) 0.0258 (13) 0.0200 (11) 0.0310 (14) 0.0244 (12) 0.0150 (10) 0.0210 (10) 0.0197 (10) 0.0239 (11) 0.0355 (13) 0.0331 (13) 0.0228 (11) 0.0363 (14)

0.0071 (2) 0.0075 (2) 0.0072 (7) 0.0075 (7) 0.0044 (8) 0.0039 (7) −0.0018 (8) 0.0080 (9) 0.0093 (9) −0.0093 (13) −0.0002 (14) −0.0188 (16) −0.0172 (13) 0.0152 (12) −0.0017 (9) −0.0013 (9) −0.0013 (8) −0.0004 (9) 0.0068 (10) 0.0063 (10) 0.0012 (9) −0.0008 (11)

0.0051 (2) 0.0024 (2) 0.0044 (6) −0.0035 (7) 0.0040 (7) 0.0024 (7) 0.0016 (8) 0.0061 (9) 0.0059 (9) 0.0010 (12) −0.0034 (12) 0.0103 (11) 0.0104 (13) 0.0063 (12) 0.0028 (9) 0.0015 (8) 0.0004 (9) −0.0046 (9) −0.0067 (10) 0.0035 (10) 0.0041 (9) −0.0129 (12)

0.0040 (2) 0.0047 (2) 0.0034 (7) 0.0064 (7) 0.0058 (8) 0.0026 (8) 0.0000 (9) 0.0038 (10) 0.0005 (9) 0.0040 (12) −0.0029 (13) 0.0032 (12) 0.0078 (14) −0.0001 (12) 0.0010 (9) −0.0030 (9) −0.0039 (9) −0.0011 (9) −0.0046 (11) −0.0077 (11) −0.0030 (10) 0.0070 (11)

233

Table A2.7: Geometric parameters (Å, º) of 3. S1—C1 S1—C2 S2—C1 O1—C11 O1—H1O O2—C12 O2—C16 N1—C1 N1—N2 N1—H1N N2—C9 C2—C3 C2—H2A C2—H2B C3—C4 C3—C8 C4—C5 C4—H4 C5—C6 C5—H5 C1—S1—C2 C11—O1—H1O C12—O2—C16 C1—N1—N2 C1—N1—H1N N2—N1—H1N C9—N2—N1 N1—C1—S2 N1—C1—S1

1.749 (2) 1.817 (2) 1.670 (2) 1.354 (2) 0.836 (10) 1.368 (3) 1.429 (3) 1.338 (2) 1.371 (2) 0.874 (9) 1.290 (2) 1.504 (3) 0.9900 0.9900 1.370 (3) 1.376 (3) 1.392 (3) 0.9500 1.370 (4) 0.9500 101.75 (10) 108.6 (16) 117.13 (17) 119.96 (16) 120.0 (15) 120.0 (15) 117.70 (17) 121.37 (15) 113.89 (15)

C6—C7 C6—H6 C7—C8 C7—H7 C8—H8 C9—C10 C9—H9 C10—C11 C10—C15 C11—C12 C12—C13 C13—C14 C13—H13 C14—C15 C14—H14 C15—H15 C16—H16A C16—H16B C16—H16C C3—C8—C7 C3—C8—H8 C7—C8—H8 N2—C9—C10 N2—C9—H9 C10—C9—H9 C11—C10—C15 C11—C10—C9 C15—C10—C9 O1—C11—C10

1.359 (4) 0.9500 1.388 (3) 0.9500 0.9500 1.450 (3) 0.9500 1.401 (3) 1.408 (3) 1.408 (3) 1.383 (3) 1.400 (3) 0.9500 1.370 (3) 0.9500 0.9500 0.9800 0.9800 0.9800 120.5 (2) 119.7 119.7 121.21 (18) 119.4 119.4 119.1 (2) 121.72 (18) 119.15 (18) 123.47 (19)

234

S2—C1—S1 C3—C2—S1 C3—C2—H2A S1—C2—H2A C3—C2—H2B S1—C2—H2B H2A—C2—H2B C4—C3—C8 C4—C3—C2 C8—C3—C2 C3—C4—C5 C3—C4—H4 C5—C4—H4 C6—C5—C4 C6—C5—H5 C4—C5—H5 C7—C6—C5 C7—C6—H6 C5—C6—H6 C6—C7—C8 C6—C7—H7 C8—C7—H7

124.74 (11) 107.49 (14) 110.2 110.2 110.2 110.2 108.5 118.2 (2) 121.0 (2) 120.8 (2) 121.0 (2) 119.5 119.5 120.3 (3) 119.8 119.8 118.8 (2) 120.6 120.6 121.1 (3) 119.4 119.4

O1—C11—C12 C10—C11—C12 O2—C12—C13 O2—C12—C11 C13—C12—C11 C12—C13—C14 C12—C13—H13 C14—C13—H13 C15—C14—C13 C15—C14—H14 C13—C14—H14 C14—C15—C10 C14—C15—H15 C10—C15—H15 O2—C16—H16A O2—C16—H16B H16A—C16—H16B O2—C16—H16C H16A—C16—H16C H16B—C16—H16C

116.60 (18) 119.92 (18) 125.4 (2) 114.77 (18) 119.8 (2) 120.2 (2) 119.9 119.9 120.4 (2) 119.8 119.8 120.6 (2) 119.7 119.7 109.5 109.5 109.5 109.5 109.5 109.5

235

Table A2.8: Hydrogen-bond geometry (Å, º) of 3. D—H···A D—H H···A D···A D—H···A O1—H1O···N2 0.84 (2) 1.90 (2) 2.639 (2) 146 (2) N1—H1N···S2i 0.88 (2) 2.47 (2) 3.3351 (18) 168 (2) C6—H6···O1ii 0.95 2.51 3.453 (3) 170 Symmetry codes: (i) −x+1/2, −y+3/2, −z+1; (ii) −x+1/2, y+1/2, −z+3/2.

236

APPENDIX FOR CHAPTER 3 (A3)

Table A3.1: Experimental and calculated FTIR vibrations (cm-1) for the organotin(IV) compounds (4-9).

Compound Method IR bands (cm-1) v(C=N) v(N-N) v(C-S)

4 Experimental 1588 1076 963 B3LYP/LanL2DZ/6-311G(d,p) 1621 1050 968

5 Experimental 1589 1068 958 B3LYP/LanL2DZ/6-311G(d,p) 1625 1047 959

6 Experimental 1579 1019 958 B3LYP/LanL2DZ/6-311G(d,p) 1624 1048 959

7 Experimental 1580 1076 959 B3LYP/LanL2DZ/6-311G(d,p) 1570 1006 934

8 Experimental 1589 1071 958 B3LYP/LanL2DZ/6-311G(d,p) 1623 1041 966

9 Experimental 1581 1026 959 B3LYP/LanL2DZ/6-311G(d,p) 1619 1047 969

237

Table A3.2: 1H NMR spectral data for the organotin(IV) compounds (4-9).

Compound 1H NMR Assignment, δ (ppm)

CH CH2 O-CH3 Ar- CH3 Sn-CH3 Aromatic Protons 4 8.77 (s, 1H) 4.47 (s, 2H) 3.97 (s, 3H) 2.43 (s, 3H) - 6.69-7.94 (m,17H) 5 8.74 (s, 1H) 4.41 (s, 2H) 3.94 (s, 3H) 2.32 (s, 3H) - 6.69-7.93 (m,17H) 6 8.74 (s, 1H) 4.45 (s, 2H) 3.96 (s, 3H) - - 6.69-7.93 (m,18H) 7 8.76 (s,1H) 4.42 (s, 2H) 3.85 (s, 3H) 2.42 (s, 3H) 0.97 (s, 6H) 6.69-7.34 (m,7H) 8 8.73 (s, 1H) 4.36 (s, 2H) 3.84 (s, 3H) 2.32 (s, 3H) 0.95 (s, 6H) 6.69-7.27 (m,7H) 9 8.73 (s, 1H) 4.40 (s, 2H) 3.85 (s, 3H) - 0.95 (s, 6H) 6.69-7.40 (m,8H)

Table A3.3: 13C1H NMR spectral data for the organotin(IV) compounds (4-9).

Compound 13C1H NMR Assignment, δ (ppm)

C=S/C-S C=N O-CH3 CH2 Ar-CH3 Sn-CH3 Aromatic carbons

4 171.9 166.2 56.6 34.4 19.4 - 116.2, 117.2, 126.1, 126.3, 127.9, 128.9, 130.2, 130.4, 130.6, 136.1, 142.0, 152.0, 159.3

5 171.8 166.1 56.6 36.0 21.2 - 116.3, 117.2, 126.1, 128.9, 129.2, 129.4, 130.2, 133.5, 135.8, 136.0, 137.2, 142.0, 152.0, 159.3

6 171.6 166.2 56.6 36.1 - - 115.2, 116.2, 117.1, 117.2, 126.1, 127.5, 128.7, 128.9, 129.3, 130.3, 136.1, 136.7, 141.9, 152.0

7 173.9 166.3 56.3 34.6 19.4 7.1 115.9, 116.5, 116.9, 126.1, 126.3, 127.9, 130.4, 130.6, 134.1, 137.2, 151.5, 158.5

8 173.7 166.2 56.3 36.2 21.2 7.1 115.9, 116.4, 116.8, 119.5, 126.1, 129.2, 133.6, 137.1, 151.5,158.5

173.6 166.3 56.3 36.4 - 7.1 115.9, 116.5, 116.9, 126.1, 127.4, 128.7, 129.3, 136.8, 151.5, 158.5

238

Table A3.4: Geometric (Å, °) details of the specified intermolecular interactions for 4. Atoms Bond lengths (Å) Bond

angle (°)

Symmetry operation

C10a H10a O1 0.95 2.58 3.490(9) 160 x, -1+y, z C2a H2a1 Cg(C11-

C16) 0.99 2.78 3.568(10) 137 x, -1+y, z

C2 H2b Cg(C11a-C16a)

0.99 2.85 3.610(11) 134 x, y, z

C18a H18f Cg(C3-C8) 0.98 2.95 3.816(11) 148 -1+x, y, z Table A3.5: Geometric (Å, °) details of the intermolecular specified interactions for 5. Atoms Bond lengths (Å) Bond

angle (°)

Symmetry operation

C8a H8a O2a 0.95 2.58 3.341(4) 137 1-x, -y, -z C17a H17e O1a 0.98 2.57 3.526(4) 165 1-x, 1-y, -z C17a H17e O2a 0.98 2.58 3.306(3) 131 1-x, 1-y, -z C5 H5 Cg(chelate)* 0.95 2.78 3.555(3) 139 1-x, -y, 1-z C19a H19e Cg(C3-C8) 0.98 2.81 3.755(3) 163 1-x, -y, 1-z

*chelate ring defined by Sn1, O1, N2, C10-C12 Table A3.6: Geometric (Å, °) details of the intermolecular specified interactions for 6. Atoms Bond lengths (Å) Bond

angle (°) Symmetry operation

C8a H8a Cg(C10a-C15a)

0.95 2.87 3.700(5)

147 -x, -y, 1-z

Cg(chelate)* Cg(C10-C15) 4.086(2)

-x, 2-y, 2-z

Cg(chelate)** Cg(C10a-C15a) 4.019(2)

1-x, -y, 1-z

*chelate ring defined by Sn1, S1, N1, N2, C1; closest edge-to-edge contact: N1...C13 = 3.040(6) Å. *chelate ring defined by Sn1a, S1a, N1a, N2a, C1a; closest edge-to-edge contact: N1a...C13a = 3.060(6) Å.

239

Table A3.7: Experimental and calculated UV-visible absorption data for the organotin(IV) compounds (4-9).

Compounds Wavelength (nm)

Experimental B3LYP/LanLD2Z/6-311G(d,p)

4 444 426 371 366 315 304

5 443 423 373 358 306 312

6 433 423 364 357 307 313

7

444 426 372 358 308 320

8 447 425 373 357 308 313

9

443 429 373 366 315 307

240

Figure A3.1: (a) Electronic absorption spectra of (i) 7, (ii) 8 and (iii) 9 and (b) Plot of [DNA]/εa - εf vs [DNA] for absorption titration of DNA with (i) 7, (ii) 8 and (iii) 9. (The arrow indicates the change in absorbance in tandem with increasing DNA concentration).

241

APPENDIX FOR CHAPTER 4 (A4)

Table A4.1: Experimental and calculated vibrations (cm-1) of the Schiff bases (10-12) and their organotin(IV) compounds (13-18).

Compound Method IR bands (cm-1) v(m-OH) v(o-OH) v(NH) v(C=N) v(N-N) v(C=S)/v(C-S)

10 Experimental - 3496 3097 1608 1119 1013 B3LYP/6-311G(d,p) 3712 3368 3359 1597 1112 1025

11 Experimental - 3501 3106 1607 1114 1012 B3LYP/6-311G(d,p) 3711 3365 3358 1597 1113 1027

12 Experimental - 3488 3097 1608 1119 1018 B3LYP/6-311G(d,p) 3712 3363 3358 1614 1113 1026

13 Experimental 3465 - - 1610 1021 959 B3LYP/LanL2DZ/6-311G(d,p) 3722 - - 1624 1037 961

14 Experimental 3462 - - 1611 1006 956 B3LYP/LanL2DZ/6-311G(d,p) 3721 - - 1625 1035 967

15 Experimental 3452 - - 1610 1016 960 B3LYP/LanL2DZ/6-311G(d,p) 3720 - - 1625 1035 966

16 Experimental - - 1597 1081 984 B3LYP/LanL2DZ/6-311G(d,p) 3830 - - 1621 1069 962

17 Experimental 3452 - - 1613 1006 956 B3LYP/LanL2DZ/6-311G(d,p) 3830 - - 1621 1033 962

18 Experimental 3448 - - 1612 1016 960 B3LYP/LanL2DZ/6-311G(d,p) 3830 - - 1621 1070 962

242

Table A4.2: 1H NMR spectral data of the Schiff bases (10-12) and their organotin(IV) compounds (13-18).

Compound 1H NMR Assignment, δ (ppm)

NH OH CH CH2 Ar- CH3 Sn-CH3 Aromatic Protons

10 13.41 (s, 1H) 9.55 (s, 1H); 9.51 (s, 1H) 8.51 (s, 1H) 4.45 (s, 2H) 2.34 (s, 3H) - 6.69-7.39 (m, 7H)

11 13.39 (s, 1H) 9.54 (s, 1H); 9.53 (s, 1H) 8.51 (s, 1H) 4.44 (s, 2H) 2.29 (s, 3H) - 6.69-7.30 (m, 7H)

12 13.39 (s, 1H) 9.56 (s, 1H); 9.53 (s, 1H) 8.50 (s, 1H) 4.47 (s, 2H) - - 6.67-7.39 (m, 8H)

13 - - 8.81 (s, 1H) 4.44 (s, 2H) 2.40 (s,3H) - 6.40-7.80 (m, 17H) 14 - - 8.80 (s, 1H) 4.41 (s, 2H) 2.34 (s, 3H) - 6.43-7.81 (m, 17H) 15 - - 8.76(s, 1H) 4.44 (s, 2H) - - 6.43-7.80 (m, 18H) 16 - - 8.79 (s, 1H) 4.43 (s, 2H) 2.42 (s,3H) 0.93 (s, 6H) 6.27-7.34 (m, 7m) 17 - - 8.77 (s, 1H) 4.37 (s, 2H) 2.34 (s, 3H) 0.92 (s, 6H) 6.24-7.27 (m, 7H) 18 - - 8.76 (s, 1H) 4.41 (s 2H), - 0.92 (s, 6H) 6.24-7.38 (m, 8H)

243

Table A4.3: 13C NMR spectral data of the Schiff bases (10-12) and their organotin(IV) compounds (13-18).

Compound 13C1H NMR Assignment, δ (ppm)

-C=S C=N -CH2 Ar-CH3 Sn-CH3 Aromatic carbons

10 195.6 146.5 36.9 19.3 - 118.1, 118.5, 119.7, 120.0, 126.6, 128.2, 130.7, 130.8, 134.3, 137.3, 146.2, 146.2

11 195.8 146.5 38.0 21.2 - 118.1, 118.4, 119.8, 119.9, 129.6, 129.7, 133.8, 137.0, 146.1, 146.2 12 195.6 146.5 38.1 - - 118.1, 118.4, 119.7, 119.9, 127.7, 129.0, 129.7, 137.1, 146.1

13 172.3 166.1 34.3 19.3 - 114.7, 118.3, 118.5, 125.1, 126.3, 127.9, 129.1, 130.3, 130.6, 133.7, 135.6, 136.1, 137.1, 141.9, 147.9, 154.9

14 172.1 166.1 35.9 21.1 - 114.7, 118.2, 118.4, 125.1, 129.0, 129.1, 129.3, 130.4, 133.2, 135.6, 137.2, 141.8, 147.8, 154.9

15 172.0 166.2 36.0 - - 114.7, 118.3, 118.5, 125.1, 127.5, 128.6, 129.0, 129.2, 130.4, 135.6, 136.4, 141.8, 147.8, 154.9

16 173.3 166.1 34.6 19.5 6.9 114.5, 117.9, 119.0, 124.8, 126.1, 126.4, 128.0, 129.9, 130.4, 130.7, 137.2, 148.0

17 173.3 166.1 36.0 21.2 6.8 114.4, 117.7, 124.6, 129.1, 129.3, 133.4, 137.1, 147.8, 154.5 18 173.1 166.2 36.2 - 6.8 114.3, 117.7, 124.6, 127.4, 128.6, 129.2, 136.6, 147.8, 154.2, 154.7

244

Figure A4.1: GCMS spectra of (a) 10, (b) 11 and (c) 12.

S NH

SN

HOOH

C16H16N2O2S2Mol. Wt.: 332.44

(b)

(a)

S NH

SN

HOOH

C16H16N2O2S2Mol. Wt.: 332.44

S NH

SN

HOOH

C15H14N2O2S2Mol. Wt.: 318.41

(c)

245

Table A4.4: Experimental and theoretical UV-vis data of the Schiff bases (10-12) and their organotin(IV) compounds (13-18).

Compound Experimental B3LYP/6-311G(d,p) Wavelength (nm) Wavelength (nm)

10 381 372 326 335

11 392 386 343 345

12 390 384 343 343

13 426 419 360 362 308 331

14 429 418 360 364 312 316

15 433 418 362 363 306 331

16 433 424 358 359 311 328

17 433 424 358 360 308 313

18 434 424 360 359 304 312

246

Figure A4.2: Molecular packing in 12: a view of the unit cell contents in projection down the a-axis. The hydroxyl-O‒H…O(hydroxyl) and thioamide-N‒H…S(thione) hydrogen bonds are shown as orange and green dashed lines, respectively. The hydroxyphenyl-C‒H…π(benzyl-phenyl) interactions are shown as purple dashed lines. Intramolecular hydrogen bonding O1 H1o N2 0.84(6) 1.87(5) 2.635(6) 151(6) x, y, z O2 H2o O1 0.84(5) 2.31(6) 2.703(7) 109(5) x, y, z Intermolecular interactions N1 H1n S1 0.88(4) 2.53(4) 3.390(5) 169(5) 2-x, 2-y, 1-z O2 H2o O1 0.84(5) 2.07(6) 2.814(6) 148(5) 1-x, 2-y, 2-z C6 H6 Cg(C10-C15) 0.95 2.66 3.500(7) 147 1-x, 2-y, 2-z C6 H6 Cg(C10-C15) 0.95 2.76 3.550(7) 141 1½-x, ½+y, 1½-z

247

Figure A4.3: Molecular packing in 14: a view of the unit cell contents in projection down the a-axis. The tin-bound phenyl- and hydroxyphenyl-C‒H…π(Sn-bound phenyl) and π(chelate ring)…π(hydroxyphenyl) interactions are shown as purple and orange dashed lines, respectively. Intramolecular hydrogen bonding O2 H2o O1 0.84(4) 2.12(5) 2.640(4) 120(4) x, y, z O2a H3o O1a 0.84(4) 2.19(4) 2.666(4) 116(4) x, y, z Intermolecular interactions C7 H7 Cg(C17-C22) 0.95 2.80 3.685(5) 155 x, ½-y, ½+z C18a H18a Cg(C23-C28) 0.95 2.92 3.609(5) 130 1 -x, -y, 1-z Cg(Sn1,S1,C1,N1,N2) Cg(C3-C8) 3.874(2) 0.95(15) x,½-y,-½+z

248

(a) (b)

(c) Figure A4.4: Molecular packing in 15: (a) a view of the linear supramolecular chain comprising Sn1-molecules and sustained by benzyl-phenyl-C‒H…O(hydroxyl) interactions, (b) the centrosymmetric supramolecular dimer connecting Sn1a- molecules via tin-bound-phenyl-C‒H…O(hydroxyl) interactions, (c) a view of the unit cell contents in projection down the a-axis. The C‒H…O, C‒H…π and π…π interactions are shown as orange, purple and blue dashed lines, respectively. In the images of (a) and (b), the hydrogen atoms have been removed for reasons of clarity. Intramolecular hydrogen bonding O2 H2o O1 0.84(2) 2.16(4) 2.661(3) 119(3) x, y, z O2a H3o O1a 0.84(3) 2.15(3) 2.673(3) 120(3) x, y, z Intermolecular interactions C14 H14 O2 0.95 2.47 3.398(4) 165 1+x, -1+y, z C18a H18a O2a 0.95 2.49 3.424(4) 169 -1-x, 1-y, 1-z C8 H8 Cg(C16-C21) 0.95 2.68 3.540(3) 152 1-x, -y, 2-z C8a H8a Cg(C22a-C27a) 0.95 2.68 3.543(3) 152 -x, -y, 1-z Cg(C16-C21) Cg(C16a-C21a) 3.9121(18) 15.36(15) 1+x, y, z Cg(C22-C27) Cg(C22a-C27a) 3.8340(19) 8.32(16) x, y, z

249

Figure A4.5: Molecular packing in 17: a view of the unit cell contents in projection down the a-axis. The C‒H…O and C‒H…π interactions are shown as orange and blue dashed lines, respectively. Intramolecular hydrogen bonding O2 H2o O1 0.84(6) 2.16(7) 2.669(5) 119(5) x, y, z Intermolecular interactions C17 H17b Cg(C10-C15) 0.95 2.70 3.671(7) 171 x, 1½-y, -½+z C8 H8 O2 0.95 2.69 3.348(6) 127 -½+x, 1-y, z

250

Figure A4.6: Molecular packing in 18: a view of the unit cell contents in projection down the c-axis. The C‒H…O and C‒H…π interactions are shown as orange and blue dashed lines, respectively. Intramolecular hydrogen bonding O2 H2o O1 0.84(6) 2.21(7) 2.679(5) 115(5) x, y, z Intermolecular interactions C16 H16b Cg(C10-C15) 0.95 2.80 3.628(7) 142 ½+x, ½+y, -½+z C2 H2 O2 0.95 2.59 3.250(6) 127 -½+x, 1-y, z

251

Table A4.5: Selected calculated bond lengths (Å) and angles (°) by B3LYP level of theory.

Bond lengths 12 14 15 17 18 Sn‒S1 - 2.56236 2.56293 2.57621 2.57747 Sn‒O1 - 2.09197 2.09221 2.08185 2.08119 Sn‒N2 - 2.23866 2.23861 2.23856 2.23862 N1‒N2 1.35918 1.38671 1.38669 1.3882 1.38827 C1‒S1 1.65902 1.7542 1.7535 1.749 1.74883 C1‒S2 1.7751 1.77013 1.77134 1.77346 1.77356 C1‒N1 1.3676 1.29024 1.29006 1.29297 1.29304 C2‒N2 1.29003 1.30927 1.30954 1.31258 1.3123 Bond angles S1-Sn-O1 - 159.663 159.762 159.272 158.929 S1-Sn-N2 - 77.542 77.618 77.37 77.23 O1-Sn-N2 - 82.121 82.189 81.952 81.873 C-Sn-C - 123.563 123.452 123.801 123.931 C1-N1-N2 124.164 117.447 117.511 117.841 117.706 C2-N2-N1 118.233 112.916 112.81 112.574 112.666 S1-C1-S2 120.775 112.679 112.655 112.584 112.601 S1-C1-N1 119.99 128.158 128.249 127.961 127.921 S2-C1-N1 119.211 119.193 119.094 119.451 119.461

252

Figure A4.7: (a) Electronic absorption spectra of (i) 13, (ii) 14 and (iii) 15; (b) Plot of [DNA]/εa - εf vs [DNA] for absorption titration of DNA with (i) 13, (ii) 14 and (iii) 15. (The arrow indicates the change in absorbance in tandem with increasing DNA concentration).

253

APPENDIX FOR CHAPTER 5 (A5)

Table A5.1: Experimental and calculated vibrations (cm-1) of tin(IV) compounds (19-24).

Compound Method IR bands (cm-1) v(m-OH) v(C=N) v(N-N) v(C-S)

19 Experimental - 1587 1088 955 B3LYP/LanL2DZ/6-311G(d,p) - 1578 1024 929

20 Experimental - 1593 1083 958 B3LYP/LanL2DZ/6-311G(d,p) - 1623 1058 967

21 Experimental - 1582 1031 958 B3LYP/LanL2DZ/6-311G(d,p) - 1619 1056 949

22 Experimental 3384 1610 1035 964 B3LYP/LanL2DZ/6-311G(d,p) 3706 1623 1073 966

23 Experimental 3452 1620 1006 965 B3LYP/LanL2DZ/6-311G(d,p) 3708 1622 1074 962

24 Experimental 3033 1612 1016 960 B3LYP/LanL2DZ/6-311G(d,p) 3707 1628 1074 962

Table A5.2: 1H NMR spectral data for the tin(IV) compounds (19-24).

Compound 1H NMR Assignment, δ (ppm)

CH CH2 O-CH3 Bz- CH3 Aromatic Protons 19 8.81, 8.45 (s, 2H) 4.47, 4.35 (s, 4H) 3.85, 3.57 (s, 6H) 2.40, 2.36 (s, 6H) 6.77-7.33 (m, 14H) 20 8.77 (s, 2H) 4.36 (s, 4H) 3.56 (s, 6H) 2.33 (s, 6H) 6.76-7.24 (m, 14H) 21 8.78 (s, 2H) 4.45 (s, 4H) 3.56 (s, 6H) - 6.76-7.36 (m, 16H) 22 8.90 (s, 2H) 4.48 (s, 4H) - 2.42 (s, 6H) 6.02-7.34 (m, 14H) 23 8.65 (s, 2H) 4.41 (s, 4H) - 2.29 (s, 6H) 6.50-7.35 (m, 14H) 24 8.60 (s, 2H) 4.44 (s, 4H) - - 6.27-7.41 (m, 16H

254

Table A5.3: 13C1H NMR spectral data for the tin(IV) compounds (19-24).

Compound 13C1H NMR Assignment, δ (ppm)

-C-S C=N O-CH3 -CH2 Bz-CH3 Aromatic carbons

19 177.2, 171.0

165.4, 161.7 57.0 364.0 19.4

115.2, 117.2, 118.4, 119.0, 120.2, 121.9, 126.1, 126.2, 126.4, 126.7, 127.6, 128.0, 130.4, 130.4, 130.5, 130.6, 133.5, 134.5,

137.2, 137.3, 147.6, 150.4, 152.0, 156.7

20 170.8 165.4 56.9 35.5 21.2 117.2, 118.4, 119.0, 126.7, 129.2, 129.4, 133.1, 137.3, 152.0, 156.7

21 170.7 165.5 56.9 35.7 - 117.2, 118.4, 119.0, 126.7, 127.6, 128.7, 129.3, 136.3, 152.0, 156.7

22 171.3 165.6 - 34.3 19.5 115.1, 119.0, 119.6, 125.4, 126.5, 128.3, 130.5, 130.8, 133.2, 137.4, 148.2, 152.1

23 150.8 145.2 - 37.8 21.2 118.4, 129.5, 129.6, 134.2, 136.8 24 155.9 153.4 - 38.1 - 111.6, 114.3, 115.9, 116.9, 127.5, 128.9, 129.7, 137.6, 146.9

255

Table A5.4: Experimental and calculated UV-vis absorption data of tin(IV) compounds (19-24).

Compound Wavelength (nm)

Experimental B3LYP/LanLD2Z/6-311G(d,p)

19 415 436 345 348

20 419 437 355 361

21 419 438 348 353

22 417 428 345 352

23 427 432 345 352

24 424 432 343 351

256

Table A5.5: Least-squares plane data for 20a and 21. 20a S1a-chelate ring: envelope with Sn1a lying 0.307(4) Å above the remaining atoms (r.m.s. deviation = 0.0060 Å) S3a-chelate ring: planar with r.m.s. deviation = 0.0057 Å O1a-chelate ring: envelope with Sn1a lying 0.579(4) Å above the remaining atoms (r.m.s. deviation = 0.0279 Å) O3a-chelate ring: envelope with Sn1a lying 0.135(4) Å above the remaining atoms (r.m.s. deviation = 0.0234 Å) 21 S1-chelate ring: envelope with Sn1a lying 0.392(4) Å above the remaining atoms (r.m.s. deviation = 0.0066 Å) S3-chelate ring: envelope with Sn1a lying 0.309(4) Å above the remaining atoms (r.m.s. deviation = 0.0124 Å) O1-chelate ring: envelope with Sn1a lying 0.463(4) Å above the remaining atoms (r.m.s. deviation = 0.0392 Å) O3-chelate ring: envelope with Sn1a lying 0.271(4) Å above the remaining atoms (r.m.s. deviation = 0.0343 Å)

257

Figure A5.1: The molecular structure of the second independent molecule of 20, i.e. 20a, showing atom labelling scheme and 50% displacement ellipsoids.

258

Figure A5.2: Molecular packing in 20: (a) a side-on view of the supramolecular layer in the ab-plane, (b) plane view of the supramolecular layer (non-participating hydrogen atoms have been removed for clarity) and (c) unit cell contents in projection down the a-axis. The C–H…O, C–H…S, C–H…π and π…π interactions are shown as orange, green, purple and blue dashed lines, respectively.

259

Table A5.6: Geometric parameters (Å, °) characterising intermolecular interactions in the crystal of 20. A H B H…B A…B A–H…B Symmetry operation C2 H2 S3a 2.81 3.721(3) 162 -1+x, y, z C2A H2a S3 2.81 3.628(3) 145 1+x, y, z C7 H7 O2a 2.49 3.421(4) 166 ½-x, -½+y, ½-z C7a H7a O2 2.49 3.267(4) 139 1½-x, ½+y, ½-z C16a H16f O2a 2.56 3.425(4) 147 -1+x, y, z C19 H19 S1a 2.72 3.564(3) 148 x, y, z C19a H19a S1 2.77 3.699(3) 166 x, y, z C24 H24 O4a 2.45 3.267(4) 144 1½-x, - ½+y, ½-z C24a H24a O4 2.53 3.464(3) 169 ½-x, ½+y, ½-z C16 H16b Cg(C10-C15) 2.85 3.533(4) 127 -x, -y, -z C34 H34b Cg(C10-C15) 2.84 3.609(4) 136 -x, -y, -z Cg(C3-C8) Cg(C20a-C25a) 3.4564(19) ½-x, -½+y, ½-z Cg(C20-C25) Cg(C3a-C8a) 3.5716(19) 1½-x, - ½+y, ½-z

260

Figure A5.3: Molecular packing in 21: (a) a side-on view of the supramolecular layer in the ab-plane, (b) plane view of the supramolecular layer (non-participating hydrogen atoms have been removed for clarity) and (c) unit cell contents in projection down the b-axis. The C–H…O, C–H…S, C–H…π and π…π interactions are shown as orange, green, purple and blue dashed lines, respectively.

261

Table A5.7: Geometric parameters (Å, °) characterising intermolecular interactions in the crystal of 21. A H B H…B A…B A–H…B Symmetry operation C7 H7 O4 2.60 3.487(3) 156 1-x, -y, 2-z C23 H23 O2 2.37 3.140(3) 138 2-x, -y, 2-z C25 H25b S1 2.79 3.645(3) 145 2-x, 1-y, 2-z C16 H16a Cg(C10-C15) 2.95 3.354(4) 106 1-x, -y, 2-z Cg(C3-C8) Cg(C3-C8) 3.5623(15) 1-x, -y, 2-z

262

APPENDIX FOR CHAPTER 6 (A6) Table A6.1: Experimental and calculated frequencies (cm-1) of the thiosemicarbazone Schiff bases (25-28) and tin(IV) compounds (29-40).

Compound Method IR bands (cm-1)

v(m-OH) v(o-OH) v(NH) v(NH) v(C=N) v(N-N) v(C=S)/ v(C-S)

25 Experimental - 3337 3304 1610 1109 1037 B3LYP/6-311G(d,p) - 3647 3472 3410 1611 1109 1034

26 Experimental - - - 3300 1609 1103 908 B3LYP/6-311G(d,p) - 3646 3407 3399 1611 1099 942

27 Experimental 3418 3140 1601 1112 1035 B3LYP/6-311G(d,p) 3712 3354 3511 3407 1601 1131 1016

28 Experimental 3443 - 3129 1597 1047 1029 B3LYP/6-311G(d,p) 3712 3401 3480 3367 1601 1138 1020

29 Experimental - - 3299 - 1596 1066 973 B3LYP/LanL2DZ/6-311G(d,p) - - 3665 - 1629 1075 1001

30 Experimental - - 3222 - 1590 1066 973 B3LYP/LanL2DZ/6-311G(d,p) - - 3548 - 1575 1039 971

31 Experimental - - 3308 - 1590 1066 973 B3LYP/LanL2DZ/6-311G(d,p) - - 3665 - 1629 1085 1002

32 Experimental - - 1593 1006 953 B3LYP/LanL2DZ/6-311G(d,p) 3720 - 3668 - 1629 1062 822

33 Experimental - - 1596 1006 951 B3LYP/LanL2DZ/6-311G(d,p) 3706 - 3666 - 1598 1064 821

34 Experimental - - 1585 993 951 B3LYP/LanL2DZ/6-311G(d,p) 3702 - 3664 - 1570 1087 819

35 Experimental - - 3331 - 1586 1075 832 B3LYP/LanL2DZ/6-311G(d,p) - - 3619 - 1626 1081 836

263

36 Experimental - - 3294 - 1577 1059 824 B3LYP/LanL2DZ/6-311G(d,p) - - 3505 - 1572 1045 811

37 Experimental - - 3303 - 1581 1063 824 B3LYP/LanL2DZ/6-311G(d,p) - - 3503 - 1573 1055 809

38 Experimental - - 1587 998 957 B3LYP/LanL2DZ/6-311G(d,p) 3708 - 3616 - 1624 1090 852

39 Experimental - - 1575 1001 917 B3LYP/LanL2DZ/6-311G(d,p) 3707 - 3619 - 1623 1075 839

40 Experimental - - 1588 1001 939 B3LYP/LanL2DZ/6-311G(d,p) 3708 - 3616 - 1628 1084 852

264

Table A6.2: 1H NMR spectral data for the thiosemicarbazone Schiff bases (25-28) and tin(IV) compounds (29-40).

Compound 1H NMR Assignment, δ (ppm)

NH OH CH O-CH3 N-CH3 Sn-CH3 Aromatic Protons

25 11.42 (s, 1H); 8.39 (q, 1H) 9.18 (s, 1H) 8.37 (s, 1H) 3.79 (s, 3H) 2.99 (d, 3H) - 6.93-7.54 (m,3H)

26 11.78 (s, 1H); 8.50 (s, 1H) 10.02 (s, 1H) 9.26 (s, 1H) 3.80 (s, 3H) - - 6.77-7.69 (m,8H)

27 11.40 (s, 1H); 9.49 (s, 1H)

8.37 (s, 1H) 8.38 (s, 1H) 9.02 (s, 1H) - 3.00 (s, 3H) - 6.67-7.38 (m,3H)

28 11.75 (s, 1H); 10.01 (s, 1H)

9.52 (s, 1H) 8.96 (s, 1H) 8.49 (s, 1H) - - - 6.67-7.57 (m,8H)

29 8.59 (q, 1H) - 7.96 (s, 1H) 3.96 (s, 3H) 3.01 (d, 3H) - 7.36-8.12 (m,13H) 30 8.55 (q, 1H) - 7.26 (s, 1H) 3.94 (s, 3H) 2.97 (d, 3H) 0.91 (s, 6H) 6.66-6.87 (m,3H) 31 8.88 (q, 2H) - 7.60 (s, 2H) 3.82 (s, 6H) 2.36 (d, 6H) - 6.67-7.31 (m,6H) 32 11.27 (s, 1H) - 8.57 (s, 1H) - 3.00 (s, 3H) - 6.37-8.20 (m, 13H) 33 11.30 (s, 1H) 8.38 (s, 1H) 8.77 (s, 1H) - 3.00 (s, 3H) 0.62 (s, 6H) 6.32-7.10 (m, 3H) 34 11.27 (s, 1H) 8.34 (s, 1H) 8.77 (s, 1H) - 2.83 (s, 3H) - 6.65-7.52 (m, 6H) 35 8.70 (s, 1H) - 7.99 (s, 1H) 3.96 (s, 3H) - - 6.69-7.56 (m,18H) 36 8.65 (s, 1H) - 7.53 (s, 1H) 3.85 (s, 3H) - 0.96 (s, 6H) 6.69-7.32 (m,8H) 37 9.70 (s, 2H) - 9.08 (s, 2H) 3.58 (s, 6H) - - 6.80-7.73 (m,16H) 38 11.69 (s, 1H) - 9.87 (s, 1H) - - - 6.57-9.52 (m, 18H) 39 11.68 (s, 1H) 9.84 (s, 1H) 8.51 (s, 1H) - - 0.63 (s, 6H) 6.33-7.75 (m, 8H) 40 11.74 (s, 1H) 8.48 (s, 2H) 10.00 (s, 2H) - - - 6.66-7.55 (m, 16H)

265

Table A6.3: 13C 1H NMR spectral data for the thiosemicarbazone Schiff bases (25-28) and tin(IV) compounds (29-40).

Compound 13C NMR Assignment, δ (ppm)

-C=S/-C-S C=N O-CH3 N-CH3 Sn-CH3 Aromatic carbons 25 177.6 148.4 56.4 30.8 - 113.1, 118.2, 119.2, 121.2, 139.6, 146.7 26 176.1 148.6 56.3 - - 113.4, 118.8, 119.5, 121.2, 125.4, 126.4, 128.6, 139.7, 140.4, 146.6 27 178.0 146.0 - 31.3 - 116.7, 117.4, 119.4, 121.5, 140.1, 145.6 28 176.0 146.0 - - - 117.1, 117.9, 119.5, 121.3, 125.6, 126.0, 128.5, 139.6, 141.3, 145.9 29 160.3 157.1 56.5 29.8 115.7, 116.7, 117.2, 125.1, 128.6, 129.9, 135.9, 142.5, 151.6 30 178.0 156.5 56.4 31.3 8.7 115.7, 115.8, 117.8, 118.5, 125.8, 151.3

31 170.6 163.7 56.7 19.3 - 105.4, 116.7, 117.4, 126.5, 128.6, 129.0, 130.7, 130.8, 134.9, 149.7, 151.8, 158.3

32 177.0 154.3 - 30.9 - 112.8, 113.7, 115.7, 118.4, 128.3, 129.2, 136.0, 136.1, 141.1, 145.4, 153.3,

33 177.2 155.0 - 30.7 6.8 112.6, 113.7, 115.6, 118.6, 140.7, 153.9 34 177.5 150.6 - 30.6 - 113.8, 114.3, 116.7, 118.9, 140.0. 150.1

35 164.8 162.5 56.7 - - 115.2, 116.2, 116.9, 119.4, 120.7, 123.4, 124.1, 125.4, 128.7, 128.9, 130.0, 135.9, 139.3, 142.1, 148.3, 149.7, 151.7

36 163.9 162.5 56.2 - 6.5 115.3, 116.6, 116.7, 120.5, 123.3, 125.4, 128.9, 139.4, 151.3, 156.8 37 162.2 160.3 56.4 - - 117.9, 118.2, 121.0, 123.4, 126.8, 129.2, 140.4, 151.4, 154.8

38 175.2 148.8 - - - 120.5, 120.7, 125.4, 125.5, 125.6, 125.7, 128.5, 128.6, 128.7, 128.8, 128.9, 129.0, 129.1, 129.5, 134.6, 135.2, 135.5, 136.3,

136.4, 136.7, 139.7 39 167.5 153.8 - - 7.1 110.2, 114.7, 128.3, 128.4, 128.7, 128.8, 135.3, 152.1, 152.7, 153.6

40 175.7 176.1

145.9 146.0

- - - 117.1, 119.5, 121.2, 125.6, 125.7, 125.9, 126.0, 128.4, 128.5, 128.5, 139.6

266

Figure A6.1: GCMS spectra of (a) 25, (b) 26 (c) 27 and (d) 28.

NH

NH

SN

HOOH

C9H11N3O2SMol. Wt.: 225.27

(c)

NH

NH

SN

HOO

C10H13N3O2SMol. Wt.: 239.29

NH

NH

SN

HOO

C15H15N3O2SMol. Wt.: 301.36

(a)

(b)

267

Figure A6.1: continued

NH

NH

SN

HOOH

C14H13N3O2SMol. Wt.: 287.34

(d)

268

Table A6.4: Experimental and calculated absorption data for the thiosemicarbazone Schiff bases (25-28) and tin(IV) compounds (29-40). Compounds

Wavelength (nm) Experimental B3LYP/6-311G(d,p) or

B3LYP/LanLD2Z/6-311G(d,p)

25 353 335 328 318

26 374 348 334 330

27 368 348 328 337

28 375 361 330 326

29 406 403 343 348

30 412 405 333 348

31 406 410 345 356

32 408 395 328 353

33 401 402 328 353

34 406 408 330 347

35 413 414 350 365

36 419 420 354 368

37

417 428 360 376

38 414 414 352 370

39 414 414 352 373

40 413 424 339 380

269

APPENDIX B - FTIR SPECTRA OF SCHIFF BASES AND TIN(IV) COMPOUNDS

4000 3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

1 4 7 19

Figure B1: FTIR spectra of 1, 4, 7 and 19.

4000 3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

2 5 8 20

Figure B2: FTIR spectra of 2, 5, 8 and 20.

270

3600 3000 2400 1800 1200 600

wavenumber (cm-1)

3 6 9 21

Figure B3: FTIR spectra of 3, 6, 9 and 21.

4000 3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

10 13 16 22

Figure B4: FTIR spectra of 10, 13, 16 and 22.

271

4000 3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

11 14 17 23

Figure B5: FTIR spectra of 11, 14, 17 and 23.

4000 3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

12 15 18 24

Figure B6: FTIR spectra of 12, 15, 18 and 24.

272

4000 3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

25 29 30 31

Figure B7: FTIR spectra of 25, 29, 30 and 31.

4000 3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

26 35 36 37

Figure B8: FTIR spectra of 26, 35, 36 and 37.

273

4000 3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

27 32 33 34

Figure B9: FTIR spectra of 27, 32, 33 and 34.

4000 3500 3000 2500 2000 1500 1000 500

Wavenumber (cm-1)

28 38 39 40

Figure B10: FTIR spectra of 28, 38, 39 and 40

274

APPENDIX C - 1H and 13C NMR SPECTRA OF SCHIFF BASES AND TIN(IV) COMPOUNDS

Figure C1: 1H NMR spectrum of 1

Figure C2: 13C NMR spectrum of 1

275

Figure C3: 1H NMR spectrum of 2

Figure C4: 13C NMR spectrum of 2

276

Figure C5: 1H NMR spectrum of 3

Figure C6: 13C NMR spectrum of 3

277

Figure C7: 1H NMR spectrum of 4

Figure C8: 13C NMR spectrum of 4

278

Figure C9: 1H NMR spectrum of 5

Figure C10: 13C NMR spectrum of 5

279

Figure C11: 1H NMR spectrum of 6

Figure C12: 13C NMR spectrum of 6

280

Figure C13: 1H NMR spectrum of 7

Figure C14: 13C NMR spectrum of 7

281

Figure C15: 1H NMR spectrum of 8

Figure C16: 13C NMR spectrum of 8

282

Figure C17: 1H NMR spectrum of 9

Figure C18: 13C NMR spectrum of 9

283

Figure C19: 1H NMR spectrum of 10

Figure C20: 13C NMR spectrum of 10

284

Figure C21: 1H NMR spectrum of 11

Figure C22: 13C NMR spectrum of 11

285

Figure C23: 1H NMR spectrum of 12

Figure C24: 13C NMR spectrum of 12

286

Figure C25: 1H NMR spectrum of 13

Figure C26: 13C NMR spectrum of 13

287

Figure C27: 1H NMR spectrum of 14

Figure C28: 13C NMR spectrum of 14

288

Figure C29: 1H NMR spectrum of 15

Figure C30: 13C NMR spectrum of 15

289

Figure C31: 1H NMR spectrum of 16

Figure C32: 13C NMR spectrum of 16

290

Figure C33: 1H NMR spectrum of 17

Figure C34: 13C NMR spectrum of 17

291

Figure C35: 1H NMR spectrum of 18

Figure C36: 13C NMR spectrum of 18

292

Figure C37: 1H NMR spectrum of 19

Figure C38: 13C NMR spectrum of 19

293

Figure C39: 1H NMR spectrum of 20

Figure C40: 13C NMR spectrum of 20

294

Figure C41: 1H NMR spectrum of 21

Figure C42: 13C NMR spectrum of 21

295

Figure C43: 1H NMR spectrum of 22

Figure C44: 13C NMR spectrum of 22

296

Figure C45: 1H NMR spectrum of 23

Figure C46: 13C NMR spectrum of 23

297

Figure C47: 1H NMR spectrum of 24

Figure C48: 13C NMR spectrum of 24

298

Figure C49: 1H NMR spectrum of 25

Figure C50: 13C NMR spectrum of 25

299

Figure C51: 1H NMR spectrum of 26

Figure C52: 13C NMR spectrum of 26

300

Figure C53: 1H NMR spectrum of 27

Figure C54: 13C NMR spectrum of 27

301

Figure C55: 1H NMR spectrum of 28

Figure C56: 13C NMR spectrum of 28

302

Figure C57: 1H NMR spectrum of 29

Figure C58: 13C NMR spectrum of 29

303

Figure C59: 1H NMR spectrum of 30

Figure C60: 13C NMR spectrum of 30

304

Figure C61: 1H NMR spectrum of 31

Figure C62: 13C NMR spectrum of 31

305

Figure C63: 1H NMR spectrum of 32

Figure C64: 13C NMR spectrum of 32

306

Figure C65: 1H NMR spectrum of 33

Figure C66: 13C NMR spectrum of 33

307

Figure C67: 1H NMR spectrum of 34

Figure C68: 13C NMR spectrum of 34

308

Figure C69: 1H NMR spectrum of 35

Figure C70: 13C NMR spectrum of 35

309

Figure C71: 1H NMR spectrum of 36

Figure C72: 13C NMR spectrum of 36

310

Figure C73: 1H NMR spectrum of 37

Figure C74: 13C NMR spectrum of 37

311

Figure C75: 1H NMR spectrum of 38

Figure C76: 13C NMR spectrum of 38

312

Figure C77: 1H NMR spectrum of 39

Figure C78: 13C NMR spectrum of 39

313

Figure C79: 1H NMR spectrum of 40

Figure C80: 13C NMR spectrum of 40

314

APPENDIX D - 119Sn NMR SPECTRA OF TIN(IV) COMPOUNDS

Figure D1: 119Sn NMR spectrum of 4

Figure D2: 119Sn NMR spectrum of 5

315

Figure D3: 119Sn NMR spectrum of 6

Figure D4: 119Sn NMR spectrum of 7

316

Figure D5: 119Sn NMR spectrum of 8

Figure D6: 119Sn NMR spectrum of 9

317

Figure D7: 119Sn NMR spectrum of 13

Figure D8: 119Sn NMR spectrum of 14

318

Figure D9: 119Sn NMR spectrum of 15

Figure D10: 119Sn NMR spectrum of 16

319

Figure D11: 119Sn NMR spectrum of 17

Figure D12: 119Sn NMR spectrum of 18

320

Figure D13: 119Sn NMR spectrum of 19

Figure D14: 119Sn NMR spectrum of 20

321

Figure D15: 119Sn NMR spectrum of 21

Figure D16: 119Sn NMR spectrum of 22

322

Figure D17: 119Sn NMR spectrum of 23

Figure D18: 119Sn NMR spectrum of 24

323

Figure D19: 119Sn NMR spectrum of 29

Figure D20: 119Sn NMR spectrum of 30

324

Figure D21: 119Sn NMR spectrum of 31

Figure D22: 119Sn NMR spectrum of 32

325

Figure D23: 119Sn NMR spectrum of 33

Figure D24: 119Sn NMR spectrum of 34

326

Figure D25: 119Sn NMR spectrum of 35

Figure D26: 119Sn NMR spectrum of 36

327

Figure D27: 119Sn NMR spectrum of 37

Figure D28: 119Sn NMR spectrum of 38

328

Figure D29: 119Sn NMR spectrum of 39

Figure D30: 119Sn NMR spectrum of 40

329

APPENDIX E – UV-VIS SPECTRA OF SCHIFF BASES AND TIN(IV) COMPOUNDS IN DMSO AT THREE CONCENTRATIONS (10-3, 10-4 AND 10-5)

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E1: UV-vis spectrum of 1.

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E2: UV-vis spectrum of 2.

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E3: UV-vis spectrum of 3.

330

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E4: UV-vis spectrum of 4.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E5: UV-vis spectrum of 5.

400 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E6: UV-vis spectrum of 6.

331

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E7: UV-vis spectrum of 7.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E8: UV-vis spectrum of 8.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E9: UV-vis spectrum of 9.

332

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E10: UV-vis spectrum of 10.

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E11: UV-vis spectrum of 11.

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E12: UV-vis spectrum of 12.

333

300 400 500 6000

1

2

3

4

Abso

rban

ce

wavelength (nm)

10-3

10-4

10-5

Figure E13: UV-vis spectrum of 13.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E14: UV-vis spectrum of 14.

400 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E15: UV-vis spectrum of 15.

334

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E16: UV-vis spectrum of 16.

400 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E17: UV-vis spectrum of 17.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E18: UV-vis spectrum of 18.

335

300 400 500 6000

1

2

3

4

Absr

banc

e

Wavelength (nm)

10-3

10-4

10-5

Figure E19: UV-vis spectrum of 19.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E20: UV-vis spectrum of 20.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E21: UV-vis spectrum of 21.

336

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E22: UV-vis spectrum of 22.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E23: UV-vis spectrum of 23.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E24: UV-vis spectrum of 24.

337

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength(nm)

10-3

10-4

10-5

Figure E25: UV-vis spectrum of 25.

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength(nm)

10-3

10-4

10-5

Figure E26: UV-vis spectrum of 26.

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E27: UV-vis spectrum of 27.

338

300 400 5000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E28: UV-vis spectrum of 28.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E29: UV-vis spectrum of 29.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E30: UV-vis spectrum of 30.

339

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E31: UV-vis spectrum of 31.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E32: UV-vis spectrum of 32.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E33: UV-vis spectrum of 33.

340

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E34: UV-vis spectrum of 34.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength(nm)

10-3

10-4

10-5

Figure E35: UV-vis spectrum of 35.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength(nm)

10-3

10-4

10-5

Figure E36: UV-vis spectrum of 36.

341

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E37: UV-vis spectrum of 37.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E38: UV-vis spectrum of 38.

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E39: UV-vis spectrum of 39.

342

300 400 500 6000

1

2

3

4

Abso

rban

ce

Wavelength (nm)

10-3

10-4

10-5

Figure E40: UV-vis spectrum of 40.

343

Acknowledgement of Authorship

I hereby certify that the work embodied in this thesis contains scholarly work of which I am a joint author. I have included a written declaration below endorsed in writing by my supervisor, attesting to my contribution to the scholarly work.

By signing below I, Associate Professor Dr. Alister J. Page, confirm that Enis Nadia Md Yusof contributed to the following scholarly works:

1. 2-[(1E )-([(Benzylsulfanyl)methanethioyl]aminoimino)methyl]-6-methoxyphenol: Crystal Structure and Hirshfeld Surface Analysis.Yusof, E. N. M.; Jotani, M. M.; Tiekink, E. R. T.; Ravoof, T. B. S. A.Acta Crystallogr. Sect. E Crystallogr. Commun. 2016, 72, 516–521.

As leading author, Enis Nadia Md Yusof; Carried out experiments presented in the publication Performed data analysis and figure formatting for publication Wrote initial manuscript draft Collaborated with co-authors on manuscript revision, submission and reviewer

comments Provided corrections for proofs prior to publication

2. o-Vanillin Derived Schiff Bases and Their Organotin(IV) Compounds: Synthesis,Structural Characterisation, In-Silico Studies and CytotoxicityYusof, E. N. M.; Latif, M. A. M.; Tahir, M. I. M.; Sakoff, J. A.; Simone, M. I.;Page, A. J.; Veerakumarasivam, A.; Tiekink, E. R. T.; Ravoof, T. B. S. A.Int. J. Mol. Sci. 2019, 20, 854

As leading author, Enis Nadia Md Yusof; Carried out experiments presented in the publication Performed data analysis and figure formatting for publication Wrote initial manuscript draft Collaborated with co-authors on manuscript revision, submission and reviewer

comments Provided corrections for proofs prior to publication

Signed,

Alister J. Page (Date: 15/7/2019)

344

BIODATA OF STUDENT

Enis Nadia binti Md Yusof was born on 22nd October 1989 in Terengganu, Malaysia. She

received her primary education at Sekolah Kebangsaan Kuala Pilah from 1995-1996 and

changed to Sekolah Kebangsaan Cherating, Kuantan from 1997-2001. She continued her

secondary education at Sekolah Menengah Kebangsaan Chukai, Terengganu from 2002-

2006. She had completed her matriculation program in biological science at Malacca

Matriculation College in 2008. In 2011, she was awarded her first degree, Bachelor of

Science (Hons.) majoring in Chemistry Resource and minoring in Biotechnology from

Universiti Malaysia Sarawak, Malaysia. In 2015, she was awarded Master of Science

(Inorganic Synthesis) from Universiti Putra Malaysia, Malaysia under the supervision of

Dr. Thahira Begum. In the same year 2015, she started her Doctor of Philosophy in

Inorganic Synthesis at Universiti Putra Malaysia, Malaysia under supervision of

Associate Professor Dr. Thahira Begum, supported by the Ministry of Education under

MyBrain15 Program, MyPhD. In September 2017, she joined Joint Awarded Doctoral

Degree programme with The University of Newcastle, Australia under the supervision

Associate Professor Dr. Alister J. Page and Professor Adam McCluskey. During her study

in Australia, she was granted the University of Newcastle International Postgraduate

Research Scholarship (UNIPRS) and the University of Newcastle Research Scholarship

Central (UNRSC) for 20 months.

345

LIST OF PUBLICATIONS

1. Yusof, E. N. M.; Jotani, M. M.; Tiekink, E. R. T.; Ravoof, T. B. S. A. 2-[(1E)-([(Benzylsulfanyl)methanethioyl]aminoimino)methyl]-6-methoxyphenol: Crystal Structure and Hirshfeld Surface Analysis. Acta Crystallogr. Sect. E Crystallogr. Commun. 2016, 72, 516–521.

2. Yusof, E. N. M.; Ravoof, T. B. S. A.; Tahir, M. I. M.; Jotani, M. M.; Tiekink, E. R. T. Bis4-methylbenzyl 2-[4-(propan-2-yl)benzyl-idene]hydrazine carbodithioato-κ2N2,Snickel(II): Crystal Structure and Hirshfeld Surface Analysis. Acta Crystallogr. Sect. E Crystallogr. Commun. 2017, E73, 397–402.

3. Yusof, E. N. M; Tahir, M. I. M.; Ravoof, T. B. S. A.; Loon, S. A. Cinnamaldehyde

Schiff Base of S-(4-methylbenzyl)dithiocarbazate: Crystal Structure, Hirshfeld Surface Analysis and Computational Study. Acta Crystallogr. Sect. E Crystallogr. Commun. 2017, E73, 543–549.

4. Mokhtaruddin, N. S. M.; Yusof, E. N. M.; Ravoof, T. B. S. A.; Tiekink, E. R. T.;

Veerakumarasivam, A.; Tahir, M. I. M. Unusual Saccharin-N,O (Carbonyl) Coordination in Mixed-Ligand copper(II) Complexes: Synthesis, X-Ray Crystallography and Biological Activity. J. Mol. Struct. 2017, 1139, 1–9.

5. Ravoof, T. B. S. A.; Crouse, K. A.; Tiekink, E. R. T.; Tahir, M. I. M.; Yusof, E. N. M.;

Rosli, R. Synthesis, Characterisation and Biological Activities of S-2- or S-4-methylbenzyl-β-N-(di-2-pyridyl)methylenedithiocarbazate and Cu(II), Ni(II), Zn(II) and Cd(II) Complexes. Polyhedron 2017, 133, 383–392.

6. Yusof, E. N. M.; Tiekink, E. R. T.; Jotani, M. M.; Simone, M. I.; Page, A. J.; Ravoof, T. B. S. A. Synthesis, Characterisation and Structure Determination of 3-[(1Z)-2-ylidenemethyl]benzene-1,2-diol. J. Mol. Struct. 2018, 1171, 650–657.

7. Yusof, E. N. M.; Latif, M. A. M.; Tahir, M. I. M.; Sakoff, J. A.; Simone, M. I.; Page, A. J.; Veerakumarasivam, A.; Tiekink, E. R. T.; Ravoof, T. B. S. A. O-Vanillin Derived Schiff Bases and Their Organotin(IV) Compounds: Synthesis, Structural Characterisation, In-silico Studies and Cytotoxicity. Int. J. Mol. Sci. 2019, 20, 854.

8. Yusof, E. N. M.; Nasri, N. M.; Ravoof, T. B. S. A.; Tiekink, E. R. T. A Ternary

Nickel(II) Schiff Base Complex Containing Di-Anionic and Neutral Forms of a Dithiocarbazate Schiff Base. Molbank 2019, M1057, 1–7.

9. Yusof, E. N. M.; Nasri, N. M.; Ravoof, T. B. S. A.; Jotani, M. M.; Tiekink, E. R. T.

Bis[S-benzyl 3-(furan-2-ylmethylidene)dithiocarbazato-K2N3,S]copper(II): crystal structure and Hirshfeld surface analysis. Acta Crystallogr. Sect. E Crystallogr. Commun. 2019, E75, 794-799.

10. Yusof, E. N. M., Latif, M. A. M., Tahir, M. I. M., Sakoff, J. A., Abhi Veerakumarasivam, Page, A. J., Tiekink, E. R. T., Ravoof, T. B. S. A. Homoleptic tin(IV) compounds of dinegative ONS dithiocarbazate Schiff bases: Synthesis, X-ray crystallography, DFT and cytotoxicity studies. J. Mol. Struct. 2020, 1205, 127635.

346

11. Yusof, E. N. M., Ishak, N. M., Latif, M. A. M., Tahir, M. I. M., Sakoff, J. A., Page, A. J., Tiekink, E. R. T., Ravoof, T. B. S. A.. Selective cytotoxicity of organotin(IV) compounds with 2,3-dihydroxybenzyl dithiocarbazate Schiff bases. Res. Chem. Intermediat. 2020, 1-29.

347

LIST OF CONFERENCES/ WORKSHOP ATTENDED

E. N. M. Yusof, A. J. Page, M. I. Simone and T. B. S. A. Ravoof. Cytotoxicity of organotin(IV) complexes containing dithiocarbazate Schiff bases. ASEAN Emerging Researchers Conference 2019, Sunway University, Malaysia (9-10 December 2019). Poster Presentation.

E. N. M. Yusof, A. J. Page, M. I. Simone and T. B. S. A. Ravoof. Cytotoxicity of organotin(IV) complexes containing dithiocarbazate Schiff bases. Medicinal Chemistry and Chemical Biology (MCCB) 2018, Brisbane Convention & Exhibition Centre, Brisbane, Australia (18-21 November 2018). Poster Presentation.

E. N. M. Yusof and T. B. S. A. Ravoof. Synthesis, structural characterisation and

biological activities of Sn(IV) complexes derived from S-2-methylbenzyl-β-N-(2-hydroxy-3-methoxybenzylene) dithiocarbazate. 6th International Conference of Young Chemists (ICYC), St. Giles The Wembley Hotel, Penang, Malaysia (16-18 August 2017). Poster presentation.

E. N. M. Yusof and T. B. S. A. Ravoof. Sn(II) and Diphenyltin(IV) complexes of the

o-vanillin Schiff bases derived from S-benzyl- and S-4-methylbenzyldithiocarbazate: Synthesis, spectroscopic and structural characterization. Fundamental Science Congress (FSC), Universiti Putra Malaysia, Malaysia (9-10 August 2016). Oral presentation.

16th BCA/CCG Intensive Teaching School in X-Ray Structure Analysis, Trevelyan College, Durham, UK (25th March - 2nd April 2017), As Participant.

348

UNIVERSITI PUTRA MALAYSIA

STATUS CONFIRMATION FOR THESIS / PROJECT REPORT AND COPYRIGHT

ACADEMIC SESSION : FIRST SEMESTER 2019/2020

TITLE OF THESIS / PROJECT REPORT :

SYNTHESIS, STRUCTURAL CHARACTERISATION AND CYTOTOXICITY

STUDY OF TIN(IV) COMPOUNDS CONTAINING ONS SCHIFF BASES

NAME OF STUDENT: ENIS NADIA BINTI MD YUSOF

I acknowledge that the copyright and other intellectual property in the thesis/project report belonged to Universiti Putra Malaysia and I agree to allow this thesis/project report to be placed at the library under the following terms:

1. This thesis/project report is the property of Universiti Putra Malaysia.

2. The library of Universiti Putra Malaysia has the right to make copies for educational purposes only.

3. The library of Universiti Putra Malaysia is allowed to make copies of this thesis for academic exchange.

I declare that this thesis is classified as:

*Please tick (√ )

CONFIDENTIAL (Contain confidential information under Official Secret Act 1972).

RESTRICTED (Contains restricted information as specified by the

organization/institution where research was done).

OPEN ACCESS I agree that my thesis/project report to be published as hard

copy or online open access.

349

This thesis is submitted for:

PATENT Embargo from ____________ until ______________ (date) (date)

Approved by: ________________________ ________________________ (Signature of Student) (Signature of Chairman IC No/ Passport No.: of Supervisory Committee) 891022115274 / A38956205 Name: Thahira Begum / Alister J. Page Date : Date : [Note : If the thesis is CONFIDENTIAL or RESTRICTED, please attach with the letter from the organization/institution with period and reasons for confidentially or restricted. ]